Next Article in Journal
Glycerol Valorization over ZrO2-Supported Copper Nanoparticles Catalysts Prepared by Chemical Reduction Method
Next Article in Special Issue
TiO2@MOF Photocatalyst for the Synergetic Oxidation of Microcystin-LR and Reduction of Cr(VI) in Aqueous Media
Previous Article in Journal
From Cell-Free Protein Synthesis to Whole-Cell Biotransformation: Screening and Identification of Novel α-Ketoglutarate-Dependent Dioxygenases for Preparative-Scale Synthesis of Hydroxy-l-Lysine
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review on Metal Ions Modified TiO2 for Photocatalytic Degradation of Organic Pollutants

1
School of Environmental and Chemical Engineering, Shenyang University of Technology, Shenyang 110870, China
2
School of Materials Science and Engineering, Shenyang University of Technology, Shenyang 110870, China
3
Faculty of Bioengineering and Technology, Jeli Campus, Universiti Malaysia Kelantan, Jeli 17600, Malaysia
*
Author to whom correspondence should be addressed.
Submission received: 27 July 2021 / Revised: 19 August 2021 / Accepted: 26 August 2021 / Published: 27 August 2021
(This article belongs to the Special Issue Application of TiO2 Nanotube in Electrocatalysis/Photocatalysis)

Abstract

:
TiO2 is a semiconductor material with high chemical stability and low toxicity. It is widely used in the fields of catalysis, sensing, hydrogen production, optics and optoelectronics. However, TiO2 photocatalyst is sensitive to ultraviolet (UV) light; this is why its photocatalytic activity and quantum efficiency are reduced. To enhance the photocatalytic efficiency in the visible light range as well as to increase the number of the active sites on the crystal surface or inhibit the recombination rate of photogenerated electron–hole pairs electrons, various metal ions were used to modify TiO2. This review paper comprehensively summarizes the latest progress on the modification of TiO2 photocatalyst by a variety of metal ions. Lastly, the future prospects of the modification of TiO2 as a photocatalyst are proposed.

1. Introduction

In the past, industrial development has been accompanied by the discharge of a large number of organic pollutants, which has caused great harm to the environment and living beings in general [1,2]. Research on the purification of wastewater is currently available and ongoing. Various traditional methods including physical adsorption, biodegradation, catalytic oxidation, and high-temperature incineration have been proposed [3]. Table 1 presents the traditional treatment methods and their limitations. These treatment methods have some shortcomings, and the treatment effect on organic wastewater is not very ideal. Recent studies have been devoted to a promising approach, the advanced oxidation processes (AOP) for the degradation of organic pollutants from wastewaters [3,4,5,6,7,8,9,10], due to its ability to completely mineralize the targeted pollutants [11]. There are various types of AOPs including ozone (photo-ozonation, ozonation, O3+Fe2+/Fe3+, and ozonation + catalyst and O3+H2O2), photolysis (VUV or UV), hydrogen peroxide (photo-Fenton: H2O2+Fe2+/Fe3++UV, Fenton-like reagents: H2O2+Fe2+-solid/Fe3+-solid, Fenton: H2O2+Fe2+/Fe3+, and H2O2+UV), and photocatalysis [12,13].
The ordinary chemical oxidation process has limitations in degrading some harmful substances because of its poor oxidation ability. The intermediate products in the advanced oxidation process can react with hydroxyl radicals to oxidize harmful substances into carbon dioxide and water. AOPs are a chemical purification treatment used to remove inorganic/organic materials in wastewater and water via oxidation through reactions with hydroxyl radicals (·OH). AOPs can be categorized as heterogeneous systems (semiconductor photocatalytic process) and homogeneous systems (Fenton and photo-Fenton systems, ozone, UV/H2O2) [14]. Photocatalytic degradation is a part of AOP for the degradation of organic pollutants which has proven as an effective technology [15,16]. It is more effective in comparison with the AOP as it can easily mineralize various organic components and the semiconductors used are not expensive.

2. TiO2 Photocatalysis

Photocatalysis is a general term for a photoinduced reaction that is accelerated in the presence of a catalyst. Heterogeneous photocatalysis is the most widely used and effective process for the degradation of organic pollutants and does not produce harmful intermediates at ambient temperature and pressure [17,18]. This process starts with electron excitation, which is transferred from the valence band (VB) to the empty conduction band (CB). In the presence of light, the catalyst absorbs adequate energy to become excited from the light which is equal to its energy bandgap. The excited electrons move toward the conduction band via the holes acting as a positive charge in the valence band, owing to which photogenerated species, including e/h+ pair, are created in the photocatalytic system. The generated species react with -OH groups or oxygen molecules to further produce reactive oxygen species such as superoxide anion radicals (O2) and hydroxyl radicals (•OH). Thereafter, these reactive oxygen species attack the organic molecules, decomposing them via oxidation. Finally, the e-/h+ pairs are formed on the photoexcited catalyst’s surface [19,20].
These materials are generally N-type semiconductors with discontinuous band structures different from metals, usually composed of low-valence bands full of electrons. Common semiconductor photocatalysts include ZnO, TiO2, Bi20Ti20, Bi2WO6, Nb2O5, Fe2O3, SrTiO3, BiTiO3, CuS/ZnS, ZnWO4, WO3, ZnS, Ag2CO3, and SnO2, etc., but most photocatalysts have an impact on the photocatalytic performance due to the photocorrosion phenomenon. Among the aforementioned photocatalysts, Titanium dioxide (TiO2) has been widely studied by researchers due to its low cost, high efficiency, and stability. The disadvantage of TiO2 is that it cannot be activated by visible irradiation, but by UV. Its advantages over other semiconductors include being biologically and chemically inert, relatively easy to use and produce, photo-catalytically stable, relatively cheap, and able to efficiently catalyze reactions, with no risk to humans or the environment [11]. TiO2 exists in three polymorphs such as rutile, anatase, and brookite. Both rutile and anatase have tetragonal crystal lattices while brookite appears as orthorhombic crystal lattices [21]. Rutile form is more stable than the other polymorphs among these three phases. TiO2 has applications in various products including sunscreen lotions, capacitors, paint pigments, toothpaste, food coloring agents, solar cells, and electrochemical electrodes [22].
In 1967, Kenichi Honda and Akira Fujishima jointly discovered that light irradiating a titanium dioxide electrode can carry out the electrolysis of water. The ionized water or oxygen was converted into photo-living groups with oxidizing ability. Their energy was equivalent to a high temperature of 3600 K and is highly oxidizing. In 1972, Fujishima et al. first proposed the theory of photocatalytic water splitting on TiO2 electrodes, and then in 1973, they proposed the use of titanium dioxide photocatalyst for environmental remediation, thus promoting the development of photocatalytic technology [23]. In 1995, Fujishima et al. [24] discovered that a film containing a certain amount of TiO2 was super-hydrophilic under ultraviolet light, which promoted the derivation of the field of photocatalyst and film binding. This technology utilizes some special photocatalyst powder into sewage, which could decompose toxic metal substances in the water to obtain pure water under the irradiation with ultraviolet rays. This technology was also used to remediate polluted rivers and found harmless to the environment. Since then, researchers have found ways to modify the TiO2 photocatalyst in a large amount. By 2004, Sonawane reported that Fe-TiO2 film could degrade methyl orange solution within 4 h under sunlight, with a degradation efficiency (95%) [25].
The bandgap of pure nano-titanium dioxide powder is about 3.2 eV, which can only absorb ultraviolet light below 400 nm. In the natural environment, ultraviolet light is less than 10%, so pure nano-titanium dioxide does not have the function of a photocatalyst. This greatly affects the utilization rate of the catalyst to sunlight. Other key factors that limit the practical application of photocatalysts are the low quantum yield of the catalyst and the faster recombination of electrons and photogenerated holes. Generally, the recombination efficiency of photogenerated carriers determines the quantum quality of the photocatalyst, which affects carrier recombination. The main factors of the surface charge transfer process are the surface morphology, grain size, crystal phase structure, and surface lattice defects of the catalyst, as well as the intensity of light radiation.
To make the TiO2 photocatalyst absorb visible light and far-infrared light and improve the photocatalytic efficiency, the photocatalyst needs to be modified. At present, researchers have proposed two ways to enhance the photocatalytic performance of TiO2: one is to prevent photogenerated electrons from recombining with photogenerated holes so that they can effectively participate in the catalytic degradation reaction process; the other is to introduce other elements into TiO2 lattice, thereby reducing the bandgap energy of the catalyst and expanding its photoresponse range. Generally, the activated CB electrons and VB holes recombine into a neutral body to generate energy, which is lost in the form of light energy or heat. Therefore, the photogenerated electrons can be captured by the catalyst or the mobility of the surface charge of the catalyst can be improved. To slow down the recombination of hole–electron pairs, modification of TiO2 lattice with metal ions is an effective way to improve photocatalytic performance. In the subsequent sections, recent studies on metal ion modified TiO2 (with morphologies such as nanoparticles, microspheres, nanofibers, nanocrystals, nanosheets, nanotubes, nanopowders) were summarized and reported.

3. Mechanisms of TiO2 Photocatalysts for Organic Pollutants

The mechanism of the photocatalytic reaction in the presence of TiO2 photocatalysts consists of a free radical reaction initiated by light irradiation (photons) [26]. When the energy of solar radiation exceeds the bandgap of TiO2 (i.e., photon energy reaches or exceeds its bandgap energy), the surface of the photocatalyst becomes excited, and the electrons transit from the valence band (VB) to the conduction band (CB). In the CB, corresponding electron holes are derived in the VB at the same time, forming electron–hole pairs (i.e., generating electron ( e ) and hole ( H + ) pairs). VB holes have strong oxidation reaction activity (1.0~3.5 V) because they lose electrons and act as reducing agents, and electrons in the conduction band have good reducibility 0.5~1.5 V when they undergo reduction. Under light irradiation, positive holes and electrons are generated in the VB ( hv vb + ) and CB ( e cb ) of TiO2 as presented in Equation (1) [27]. These holes can either form hydroxyl radicals (Equation (3)) or react directly with organic molecules (Equation (5)) which subsequently oxidize the organic molecules (Equation (6)) [28]. The electrons can also react with organic compounds to produce reduction products (Equations (1)–(7)). The role of oxygen is important as it reacts with the photogenerated electrons. Organic compounds can then undergo oxidative degradation through their reactions with hydroxyl and peroxide radicals, VB holes, as well as reductive cleavage via reactions with electrons yielding various byproducts and finally mineral end-products [29].
hv +   TiO 2   hv vb + +   e cb
OH surface +   hv vb +     OH
H 2 O absorbed +   hv vb +     H + +   OH
O absorbed +   e cb     OH 2
Organic +   hv vb +   Oxidation   products
Organic +   OH   Degradation   products
Organic +   e cb   Reduction   products
Additionally, VB holes ( h vb + ) react with water ( H 2 O ) and the hydroxyl ion ( OH ) to form hydroxyl radicals ( HO ), while electrons ( e cb ) react with adsorbed oxygen molecule (O2), thus reducing it to superoxide radical anion ( O 2 ) which reacts with protons ( H + ) to form peroxide radicals ( HO 2 ) as shown in Equations (8)–(11) [30].
h vb + +   H 2 O     HO +   H +
h vb + +   OH     HO
e cb +   O 2     O 2
O 2 +   H +     HO 2
In the presence of light irradiation using metal ion modified TiO2 and because the metal ions create intermediate states in the TiO2 structure, the surface electrons in the intermediate states become excited, having sufficient energy to access more light absorption and transfer electrons to the TiO2 surface that stimulate more electrons under light, which as a consequence advances the redox reactions [31,32,33]. These properties being displayed by the metal ions help to improve the photocatalytic reactions. The transfer of electrons takes place at the VB to the CB via the creation of holes in the former and these holes subsequently react with H 2 O molecules present in the pollutant as well as helping to form the OH radicals [31,32,33,34]. The formed OH radicals are supportive and found significant success in degrading various dyes under light irradiation [35,36,37]. The photocatalytic mechanism of metal-ion modified TiO2 for organic pollutants is illustrated in Figure 1.

4. Modification with Metal Ions

Modification with metal-ion can act as electron/hole traps and alter e/h+ recombination rate, according to the below mechanisms [39]:
M n + +   e cb     M n 1 +   electron   trap
M n + +   h vb +     M n + 1 +   hole   trap
where energy level for M n + / M n 1 + lies below the CB edge ( e cb ) and the energy level for M n + / M n + 1 + above the VB edge ( e vb ).
The presence of metal ions allows visible light absorption and introduces trap/recombination sites within the TiO2 bands which may as a result increase the life span of photoinduced charge carriers as well as the reduction in quantum efficiency. Modifications with metal ions can influence TiO2 photocatalysis according to these principles: (i) improve the electron–hole separation (strength) by selective trapping [40,41]; (ii) due to their ability to act as recombination centers, they reduce carriers’ lifespan (weakness) [42,43], and (iii) enhance optical absorption in visible light range. The commonly used metal ions are the metals, transition metal, rare earth metal, or noble metal [44]. Modification of TiO2 by metal ions is generally carried out via sputtering [45], ion implantation [46], or via chemical processes (for example sol–gel) [47,48,49,50]. Figure 2 presents a table of commonly used metallic ions to enhance the photocatalytic degradation of organic pollutants.
In general, various types of metals (Transition, rare earth, or other metals) have been used to modify TiO2 photocatalysts. Table 2 presents the general properties and principal applications of the metals. In the subsequent sections, the modifications using metal ions TiO2 photocatalysts are described and discussed. The following section starts with the modification using transition metal, which is subsequently followed by lanthanides or rare earth metal modifications. At the end of this section, the modifications of other metals were presented.

4.1. Modification with Transition Metals

Modifying TiO2 with transition metals has shown great promise through the extension of spectral response and the ability to achieve visible light-activated photocatalysis [51]. Modifying with transition metal including Cu, Cr, Mn, Fe, V, and Zn can reduce the e/h+ recombination and decrease the bandgap by creating an intra-bandgap state, which shifts the light absorption into the visible light region [52,53]. The formation of additional energy levels in the bandgap of TiO2 can be described below [54]:
1   M + + hv     M n + 1 + +   e
2   M + + hv     M n 1 + +   h +
where M is the metal.
The energy necessary for reducing the metal ion should be less negative than the conduction band edge of titania while its oxidation energy should be less positive than the valence band edge of TiO2. Additionally, photocatalytic reactions can occur only if the trapped holes and electrons are transferred to the catalyst’s surface. Therefore, metal ions should be situated near the surface of TiO2 crystallites for a better charge transfer [55]. If the amount of metal ions is more than optimal, the metal ions tend to behave as recombination centers and the transfer of h+/e pairs to the interface is more complex. Generally, the metal ions are precipitated on the surface as oxides forming a MOx/TiO2 nanocluster and not incorporated in the structure of TiO2. If the energetic positions of CB and VB of components forming the composite are suitable (i.e., a type II heterojunction), beneficial charge transfer occurs within the photocatalyst where e is accumulated in the CB edge of one component and h + is accumulated in the VB edge of the other component. Therefore, the charge carriers are separated efficiently, leading to improved photocatalytic property [56,57].

4.1.1. Modification with Vanadium (V)

Vanadium exists in several oxidation states (V2+ to V5+) and a variety of species. The types of species and oxidation state are a function of the redox potential, pH, concentration, and other factors [58]. The radius of the V ion is nearly identical to that of titanium (Ti) and can be conveniently introduced into titania [59]. V ions with different valence states (V3+ to V5+) can transfer between V3+ to V5+ under oxidizing and reducing conditions. For example, the tetragonal crystal structure of VO2 is similar to that of titania, which is responsible for an increase in visible light absorption and photogenerated holes and electrons [60]. The photogenerated holes and electrons can be migrated, trapped, and released on the TiO2 surface by V4+ ions, and play a role in charge transfer species. It is very difficult for V5+ ions to enter the lattice of TiO2, making it appropriate to form V2O5 on the surface and thus responsible for h+ and e separation [61]. Subsequently, this photogenerated e and h+ can be accepted by the adsorbed O2 and surface hydroxyl group (OH), and therefore transform OH and O2 into hydroxyl radical (•OH) and superoxide radicals (•O2) active species, respectively.
Vanadium-TiO2 (V-TiO2) at low concentrations has been found to improve TiO2 photocatalytic behavior via the existence of the photogenerated charge and the more efficient expansion of absorption spectrum as compared to other metals [62,63]. V-TiO2 has an enhanced photocatalytic activity due to the following reasons: (i) improving the absorption in the visible light range, (ii) improved quantum efficiency due to effective h+ to e pair separation, and (iii) presence of both V5+ and V4+ species in the V-TiO2. Additionally, the second reason can contribute to the increased electron transfer and visible light absorption, whereas the latter enhances e–h+ separation and is a potential electron acceptor [64,65,66]. V-TiO2 have been extensively studied in experiments and theory due to evidence of ferromagnetism above room temperature or at room temperature [67,68]. Vanadium with a narrow bandgap can be coupled with TiO2 by various approaches such as metal-ion implantation, sol–gel, hydrothermal, and coprecipitation approach. Generally, the sol–gel method is frequently used for depositing the V-TiO2 films while the coprecipitation approach is used to prepare V-TiO2 powders. Among these approaches, (compared with the hydrothermal approach), the sol–gel and coprecipitation approach is quite harsh and complicated; however, the hydrothermal approach is easy to execute and involves low cost [69,70]. Table 3 presents the summary of recent progress on V-TiO2 photocatalysts for organic pollutants degradation.

4.1.2. Modification with Nickel (Ni)

Nickel (Ni) has good activity and it is less expensive than noble metals [78]. This element is used in many applications due to its physicochemical properties. Ni is a good candidate for substituting Ti atoms in the TiO2 structure [79]. The incorporation of Ni ions into the TiO2 lattice can actively modify the TiO2 physical properties via the creation of an impurity energy level. Upon modifying TiO2 with Ni, the recombination of photogenerated hole–electron pairs is suppressed effectively, thus resulting in an enhanced photocatalytic activity [79,80,81]. The structure of the Ni2+ valence band is 3d8. When the Ni ions trap the photogenerated hole–electron pairs, the valence layer (d) is converted from a high to low spin state, and thus results in a significant spin energy loss. Based on the crystal field theory, charge carriers that are being trapped by Ni ions tend to migrate to the H2O molecules (adsorbed on the surface) to restore their energy, and as a consequence will prevent the recombination of hole–electron pairs [82]. Furthermore, Ni2+ plays an important role in the improvement of thermal stability as well as controlling the morphology of TiO2 photocatalysts. Table 4 presents the summary of recent progress on Ni-TiO2 photocatalysts for organic pollutants degradation.

4.1.3. Modification with Copper (Cu)

The introduction of Cu ion can directly trap generated electrons by excitation of light, and thus prevents the rapid recombination of hole–electron pairs [95]. It can increase the surface diffusivity of the TiO2 [96], and thereby enables the interaction of holes and electrons with other compounds (e.g., H2O) more quickly [97], generating feasible reactive species, and subsequently bypassing the recombination of the hole–electron pair [98]. Cu with redox potentials of 0.16 V (Cu2+/Cu+) and 0.52 V (Cu2+/Cu) has been used as a suitable modifier for various visible light responsive photocatalysts. Ti4+ and Cu2+ have similar ionic radii and therefore Cu2+ can easily penetrate into TiO2 matrixes as a deep acceptor in conjunction with neighboring oxygen vacancies or substitute the positions of Ti4+ [99]. In addition, modification with Cu shifts the absorption edges of both the photocatalysts towards the visible region [100].
The behavior involved in modifying TiO2 with Cu is strongly linked to parameters of synthesis and diverse approaches used for fabrication of materials as well as analysis, and has led to various experimental findings in the past. Studies have observed the diverse speciation of Cu-TiO2 with the majority of studies reporting the presence of Cu in a Cu2+ valence. These species are frequently reported to substitute for Ti4+ to give a solution phase with the composition CuxO2Ti1-x and an increased lattice density of oxygen vacancies [101], and reportedly may also occupy interstices in anatase [102]. In addition to solid solutions, Cu2+ has been found to exist in amorphous or crystalline CuO nanoclusters as well as surface localized Cu(OH)2 in titania [103,104,105]. The presence of monovalent Cu + (less common as compared to Cu 2 + ) has also been reported, both in Cu2O nanoclusters and in substitutional positions [102,106]. The CB edges of CuO2 and CuO are suitable for the enhancement of TiO2 photocatalytical effect [107,108,109]. Generally, an efficient photocatalytic activity is understood to be driven by a reduced charge carrier recombination in Cu-TiO2 photocatalyst. This was found to arise as a result of photogenerated electrons which facilitates the reduction of Cu 2 + +   e     Cu + , and thus extends the valence band hole lifetimes at the surfaces, which are able to react with adsorbed species to form active radical species [103,104]. Alternatively, the presence of Cu2O or CuO phase in TiO2 may improve exciton lifetime through electron capture in the secondary phase [110] or improve the activity via the increment in surface area [111]. Table 5 presents the summary of recent progress on Cu-TiO2 photocatalysts for organic pollutants degradation.

4.1.4. Modification with Manganese (Mn)

Manganese (Mn) has a great potential to permit significant optical absorption in infrared solar light or visible light, via the introduction of intermediate bands within the forbidden gap as well as the combined effects of narrowed bandgap [129,130]. The intermediate bands provide adequate carrier mobility and hence significant curvature [131]. Mn ions can be easily incorporated into the lattice of TiO2 to obtain a deformed structure. It has been reported that Mn in 3d states has some contributions to the TiO2 conduction band, which will further impact the bandgap of TiO2. Modifying with Mn cations has been reported to favor the charge separation acting like electrons trap as Mn4+ or Mn3+ [129] and holes trap as Mn2+ [132,133], thus prolonging the separation of photoinduced carriers and increasing the photocatalytic activity [134].
In Mn-photocatalytic materials with a mixture of oxidation states could be seen as a limitation because it is impossible to easily control the amount of one specific oxidation state [135]. However, this can be approached as a benefit since the oxidation state of Mn can act as a charge separator, attracting photogenerated holes in the Mn2+/Mn3+ state or as a photogenerated electron in the Mn4+/Mn3+ state, the oxidation states mixture can improve the photocatalytic activity [133]. When TiO2 semiconductors are modified with Mn and applied in the photocatalytic degradation of organic dyes, the reduction in bandgap can be achieved, thereby improving the performance [136,137,138]. This is due to the following reasons: when Mn is introduced in the TiO2 network at different oxidation states (2+, 3+, or 4+), there is a deformation of the TiO2 crystalline structure and optimization of its optical properties [139,140,141]. Ref. [142] confirmed a high stability feature when TiO2 was modified with Mn1+ to Mn6+. Mn2+ can enter the lattice of TiO2, replacing the position of the original titanium atom while forming a new chemical bond. In addition, modifying with Mn2+ can also increase the number of hydroxyl groups on the surface of TiO2 in the aqueous solution, forming an active center for the adsorption of water [143].
Mn-TiO2 nanoparticles can be prepared by (i) sol–gel technique [131,144], (ii) hydrothermal treatment [145], (iii) ultrasonic synthesis [146], (iv) liquid flame aerosol synthesis technique [32], (v) impregnation method [147], (vi) calcination [148], and (vii) thermal treatment [149]. Table 6 presents the summary of recent progress on Mn-TiO2 photocatalysts for organic pollutants degradation.

4.1.5. Modification with Zirconium (Zr)

Zirconium (Zr) has a wide bandgap with a more positive (4.0 V vs NHE) VB and a more negative (−1.0 V vs NHE) CB than titania [155]. The tetravalent cations of both Ti and Zr have a comparable atomic radius, and their oxides have similar physicochemical properties. It has been previously reported that the modification of TiO2 with Zr can improve physicochemical characteristics including an increase in the specific surface area [156], decrease in crystallite size as a result of dissimilar coordination geometry and nuclei [157]; enhance surface acidity [158], the transition temperature between anatase and rutile [156,158], and the adsorption and the hydrophilic properties [159], and enhance thermal stability and imbalance of charge resulting from the formation of capture traps as well as the formation of Ti-O-Zr bonds. These features favor a higher efficiency while separating photogenerated carriers, thus improving the photocatalytic efficiency [160,161]. Moreover, the hydroxyl groups that are trapped on the catalyst surface by holes improve the quantum yield and reduce the recombination reactions [159,162]. Modifying with Zr conquers charge recombination by electron trapping [163].
Liu et al. [164] prepared mesoporous Zr-TiO2 NPs via the solvothermal approach and observed a higher photocatalytic activity in comparison with the commercialized P25 titania. Li et al. [165] reported a macro–mesoporous Zr-TiO2 material via the facile surfactant self-assembly approach and observed a remarkable photocatalytic activity towards RhB decomposition under UV light radiation as compared to pure ZrO2 and unmodified TiO2 materials. A study by [166] found that modifying with Zr4+ can increase the surface area, and suppress crystal growth. Table 7 presents the summary of recent progress on Zr-TiO2 photocatalysts for organic pollutants degradation.

4.1.6. Modification with Iron (Fe)

Iron (Fe) has a half-filled (3d5) electronic configuration, similar ionic radii of Fe3+ to that of Ti4+ and can be easily incorporated into the crystal lattice of TiO2 [174,175], and it has been documented that modifying the crystal lattice of TiO2 with Fe3+ ions weakens the surrounding oxygen atoms bonding. Consequently, the oxygen atoms are readily released from the lattice causing an oxygen vacancy, hence more oxygen/H2O/OH- can be adsorbed onto the TiO2 surface resulting in the decrease in CB electrons. In addition, Fe3+ ions can perform as a shallow charge trap in the TiO2 lattice, because the energy level (redox potential) of Fe2+/Fe3+ lies close to that of Ti3+/Ti4+, which favors the development of charge carrier separation, further causing a decrease in the e/h+ pair recombination and expanding the photoresponse of TiO2 into the visible light range [175,176].
Ali et al. [174] used the sol–gel method for the successful synthesis of Fe-TiO2 NPs which could be activated in the presence of visible light. Naghibi et al. [177] also studied the properties of TiO2 modified with four metals including Cu, Fe, Cd, and Ce, and observed that the Fe-TiO2 presents the smallest crystal size with forbidden bandwidth among these four metals-modified TiO2. [178] found that the Fe3+-TiO2 powder showed strong absorption characteristics in the visible region and [179] observed that introducing Fe3+ into the TiO2 lattice facilitates the formation of a redox site, enabling absorption of the visible portion. The semi-filled electronic structure of Fe3+ ions not only facilitates the separation of photoelectrons and holes but also reduces the bandgap of TiO2 [180]. In addition, modifying with Fe3+ can introduce more oxygen vacancies in TiO2 surface and lattice [181], which is beneficial for increasing the number of surface OH groups and improvement in photocatalytic degradation. Table 8 presents the summary of recent progress on Fe-TiO2 photocatalysts for organic pollutants degradation.

4.1.7. Modification with Chromium (Cr)

Chromium (Cr) has received many considerations owing to its partially filled d shell, and optically active nature. Cr3+ is suitable with an abundant electron shell structure [78]. Cr3+ ions have fewer valence electrons in comparison with Ti4+. TiO2 crystalline lattice deformation is evoked by substitution of Ti4+ by Cr3+, creating an extra energy level between the VB and the CB of TiO2, which permits those photons which have lower energy to accomplish the photocatalytic activity in the visible light range [218]. Because of the narrowing of the bandgap between valence and acceptor, the electron movement is better even at a lower temperature. In addition, such substitution restrains the crystal growth of TiO2 as well as reducing TiO2 crystallization, which may improve the absorption of light derived from the size effect [219]. Peng et al. [220] studied the influence of modification with Cr on the photocatalytic activity of TiO2 and observed a shift in absorption edges of TiO2 towards high wavelength region with the increasing Cr content. Various approaches are being employed to fabricate Cr-TiO2 such as the spray pyrolysis [221], sol–gel method [220], and sputtering [45]. Table 9 presents the summary of recent progress on Cr-TiO2 photocatalysts for organic pollutants degradation.

4.1.8. Modification with Molybdenum (Mo)

Modification of TiO2 using metals with a higher oxidation state (e.g., Mo6+ and Mo5+) increases the photocatalytic activity due to improved transfer and separation of photogenerated holes and electrons [225]. Moreover, Mo6+ is also a suitable transition metal to be introduced into the lattice of TiO2 due to similarity in the ionic radii of Ti4+ and Mo6+ [226], so Mo can easily be incorporated into TiO2 crystal structure, producing the impurity levels, which as a consequence narrows the bandgap of TiO2 [77]. In addition, Mo6+ has no electron in the d orbital and can also be reduced to lower oxidation states (Mo4+, Mo5+), implying its several oxidation states in the TiO2 matrix can act as a superficial potential trap for the photoinduced e to h+ pairs, thus lengthening the lifetime of carriers and increasing the photocatalytic activity. Furthermore, the redox potential of Mo6+/Mo5+ (vs. NHE) is 0.4 V, allowing Mo6+ to capture photoinduced electrons, inhibiting the recombination of charge carriers [227]. Mo-TiO2 photocatalysts can be synthesized using evaporation-induced self-assembly, hydrothermal method, and sol–gel process [228,229,230,231,232,233]. Table 10 presents the summary of recent progress on Mo-TiO2 photocatalysts for organic pollutants degradation.

4.1.9. Modification with Cobalt (Co)

Co2+ has a similar radius as Ti4+, which easily enters into lattice interstitial sites or replaces Ti4+ to broaden its photocatalytic activity [236,237]. Co-TiO2 is a capable candidate owing to its excellent properties such as stability, optically active nature, transparency, low cost and high n-type carrier mobility [238,239]. Furthermore, the presence of cobalt (Co2+ or Co3+) enhances crystallization due to the increased amounts of oxygen vacancies in the lattice of TiO2, created by replacing the Ti4+ sites of the TiO2 lattice [182,240,241]. Ebrahimian et al. [242] prepared Co-TiO2 NPs with a great absorption range in the visible light range. Iwasaki et al. [243] synthesized Co-TiO2 and observed that Co2+ could shift the edge of light absorption of TiO2 to the visible light range and enhance photocatalytic activity in both visible and UV regions. Additionally, [92,240] reported that modifying TiO2 lattice with Co enhances the photocatalytic activity of MO and carbamazepine under both visible and UV-A irradiations in comparison with unmodified TiO2.
Some studies have greatly emphasized the contradictory influence of cobalt: some authors have reported its incorporation as detrimental for photocatalytic activity, while some have observed a slight improvement in photodegradation or selectively for some organic compounds. Studies by [244,245] show that that the photocatalytic activity of Co-TiO2 is lower than unmodified TiO2 in the presence of UV light. Meanwhile, 0.5% Co-TiO2 presents the highest activity in the presence of visible light. Generally, most authors have concluded on the following facts (i) Co is present as the divalent form; (ii) Co (II) states are located within the bandgap; (iii) cobalt can be found in interstitial positions of TiO2 [246]. Table 11 presents the summary of recent progress on Co-TiO2 photocatalysts for organic pollutants degradation.

4.1.10. Modification with Niobium (Nb)

The incorporation of niobium (Nb) can act as electrons sink as well as minimize the recombination of e to h+ pairs, and thus make more carriers available for reduction or oxidation processes on TiO2 surface [258]. Modifying with Nb5+ enables the release of one electron for every Nb introduced: the Fermi level of TiO2 shifts upward into the CB and results in a typical n-type metallic characteristic in the electronic structure [259]. Furubayashi et al. [260] prepared an Nb5+-TiO2 film and their results show a room temperature resistivity of 2 to 3 × 10−4 Ω cm when the Nb concentration was >3%. Modifying of Nb5+ into TiO2 lattice can extend the absorption that spans from UV to visible region as well as to the mid-infrared region. This allows TiO2 to be active in the presence of both visible and UV light [261]. Some studies have also shown that charge transfer can be improved upon implantation with small Nb5+ content. Upon modifying with Nb, Nb5+ replaces Ti4+ and the donor is formed at the CB of TiO2, providing electron for Ti4+ as well as obtaining a high concentration of carrier [260]. Su et al. [262] have showed that modifying with Nb increases the electrical conductivity, decreases bandgap, and increases optoelectronic property. Khan et al. [263] also demonstrated that charge-compensating (NbTi-VTi)3− complexes serve as the dominant defect in Nb-TiO2 composite enabling the improvement in the photocatalytic activity via the formation of shallow defect level. Table 12 presents the summary of recent progress on Nb-TiO2 photocatalysts for organic pollutants degradation.

4.1.11. Modification with Tungsten (W)

Tungsten (W) can easily replace Ti in the TiO2 matrix due to the similarities in ionic radii of Ti4+ and W6+. W in its high oxidation state (+6) can act as an electron trap center and the substitutional modification of Ti4+ with W6+ can impart an almost small change in the TiO2 matrix [270]. Modifying with W can only reduce the recombination in the TiO2 bulk, whereas the defects present on the TiO2 surface can act as recombination centers [271]. Furthermore, modifying with W can lead to the placement of electrons in a state below the CB or donor state [272]. These electrons can receive energy from visible light and be transferred to the CB. Finally, these excited electrons can generate O2, OH•, and OH• which are responsible for the heterogeneous photocatalytic treatment of organic pollutants [273]. When the anatase TiO2 is modified with WO3, separation of charge is improved as a result of the coupling of both materials [274]. Moreover, the presence of WO3 can increase the TiO2 acidity, modifying the substrate’s affinity for TiO2 surface and thus affects the adsorption equilibrium and photooxidation activity of the catalyst [275,276]. Three major advantages of WO3 include the thermodynamically favorable position of VB edge for the oxidation of H2O (3 V vs NHE), visible light response with a bandgap energy of 2.6–2.8 eV, and photochemical stability in acidic media [277]. Superior photocorrosion resistance and chemical stability are the other advantages provided when TiO2 is modified with WO3. Various techniques including sol–gel, chemical vapor deposition, impregnation, mechanical blending, and liquid phase plasma (LPP) method have been also applied to the modification of host photocatalyst with transition metals [278,279]. Table 13 presents the summary of recent progress on W-TiO2 photocatalysts for organic pollutants degradation.

4.1.12. Modification with Zinc (Zn)

Zn2+ ion can easily substitute Ti4+ ion in TiO2 lattice without destroying the crystal structure, thereby stabilizing the anatase phase. Zn-TiO2 powder can be synthesized using various preparative methods such as sol-hydrothermal, homogeneous hydrolysis, sol–gel and solid-phase reaction, electrospinning, controlled hydrolysis of titanium (IV) butoxide, assembly process based on a Ligand exchange reaction; micro-emulsion and sol–gel method. Table 14 presents the summary of recent progress on Zn-TiO2 photocatalysts for organic pollutants degradation.

4.2. Modification with Noble Metals

4.2.1. Modification with Gold (Au)

Gold (Au) can act as an electron acceptor since the energy gap (Eg) of Au (5.1–5.47 eV) is greater than TiO2 (~4.7 eV) [293]. Owing to its plasmon resonance effect, modifying TiO2 with Au will allow the achievement of visible light-driven photocatalyst [294,295]. Au-TiO2 can also enhance the functional abilities of the material, causing alteration of the chemical or photocatalytic properties. Thus, the interaction between the photogenerated active species and the substrate of the Au-TiO2 interface is increased and easily distinguished [296,297,298]. The collective oscillating motion of CB electrons on the Au nanoparticles (NPs) surface is mainly responsible for the reductive properties of the NPs and originates from a well-known surface plasmon resonance (SPR) [299]. Furthermore, modifying with Au promotes the gradual injection of hot electrons into TiO2, and as a consequence increases the life span of the hole–electron pair and improve the photocatalytic activity [300]. The presence of Au in TiO2 lattice structure may cause the switching of Ti4+/Ti3+ in defects, which effectively minimizes the wide bandgap of TiO2. The presence of Ti4+/Ti3+ states inside the nanostructure enhances the electron-transfer mechanism inside the crystal lattice of TiO2, increasing the photocatalysis efficiency [301]. This property can overcome the limitations of using unmodified TiO2 photocatalyst due to its low recycling capability and complications of recovery from H2O.
The key parameters for the Au-TiO2 activity are the Au particle size and the crystal phase of TiO2. Additionally, Au-TiO2 possesses a high performance coupled with a longer life span [302]. On the other hand, the dependence Au-TiO2 activity is also influenced by the particle size of Au in the action mechanisms, which can be classified into the electron–hole transfer mechanism [303]. Deposition of Au occurred without changing the TiO2 crystal structure and also achieved effective separation of photogenerated electron–hole in the UV–visible (UV-vis) light conditions, due to the smaller Au particle size of 15 nm. Table 15 presents the summary of recent progress on Au-TiO2 photocatalysts for organic pollutants degradation.

4.2.2. Modification with Silver (Ag)

Compared to gold and platinum, silver (Ag) is the cheapest coinage metal having a high charge carrier density [315]. It also has other advantages such as higher antibacterial activity, lower cost, and toxicity [316]. Modifying with Ag ions can create a new impurity band about 0.7 eV below the CB of titania which will thus narrow the energy bandgap of titania [317]; subsequently, Ag enhances the electron–hole separation by acting as an electron trapper to capture the electrons while reacting with the contaminants [318]. Hence, Ag assists in charge separation via the formation of a Schottky barrier between the metal and TiO2 [319,320]. The incorporation of silver ion also extends the absorption of TiO2 under visible light due to SPR effect, excited by visible light [321,322,323,324]. The presence of Ag ion can also contribute in two ways: (i) reinforcement of the photoinduced charge carriers as well as enhancing the electromagnetic field at the interfaces, (ii) promotion of interfacial charge-transfer process which limits the recombination rates. The electronic band-alignment at the Ag-clusters-TiO2 interface favors the migrations of photogenerated electrons to metallic particles.
This process increases the lifespan of holes leading to redox reactions required for the photodegradation of the organic pollutants [119]. There are many approaches to synthesize Ag modified TiO2, including solvothermal [325,326], reduction using UV irradiation [327], hydrothermal [328], electrochemical [329], microwave [330], deposition–precipitation [331,332], sonochemical [333], and sol–gel methods [334,335].
Zhou et al. [336] reported Ag decorated Ti3+ self-doped porous black TiO2 pillar using a combined oil bath and wet impregnation approach. They claimed that the synergistic effect between Ti3+ self-doping and Ag decoration enhances photocatalytic activity. Liu et al. [337] presented the formation of Ag impurity band above the VB and observed the presence of AgO on the TiO2 surface as well as the presence of Ag in the bulk. They noted that this Ag system will fail to show the energy level matching concept as the presence of Ag-ions is required on the TiO2 surface. Contrary to various conventional reports, some reports have also confirmed the formation of impurity bands starting from 0.7 eV below CB [338]. Xiong et al. [339] also introduced 0.25 wt% Ag NPs into mesoporous TiO2 and observed enhanced degradation of RhB in the presence of UV light irradiation. In addition, the Ag-TiO2 catalysts present a higher photocatalytic efficiency for indigo carmine in the presence of visible light which was higher than the commercial P25 titania or unmodified TiO2 [339]. Table 16 presents the summary of recent progress on Ag-TiO2 photocatalysts for organic pollutants degradation.

4.2.3. Modification with Platinum (Pt)

Platinum (Pt) is a shiny silvery metal with low chemical activity. It is stable in air and a humid environment. The oxidation state of platinum is +2 to +6, which easily forms coordination compounds and has good ductility, thermal conductivity, and electrical conductivity. Pt is a noble metal that can increase the TiO2 photoefficiency [352]. The energy level of Pt is lower than the one of the TiO2 CB, so the energy required for the electron transition is lower and may be induced by the light with higher wavelengths. Therefore, the absorption spectrum can be broadened from the UV towards the VIS region. In the presence of UV light, the photogenerated electrons quickly transfer from the TiO2 surface to the Pt particles, leading to effective separation of electron–hole, and as a consequence improving the photocatalytic activity [353]. The deposition of Pt nanoclusters onto TiO2 surface serves as an electron sink, consequently slowing down charge–pair recombination as well as changing reaction pathways by providing catalytic sites where active intermediates can be stabilized [121]. Table 17 presents the summary of recent progress on Pt-TiO2 photocatalysts for organic pollutants degradation.

4.2.4. Modification with Ruthenium (Ru)

Modifying with ruthenium (Ru) can reduce the Egap of TiO2, allowing the visible light-driven photocatalysts, and acts as an electron acceptor/donor, which efficiently minimizes the recombination due to acceleration of electron transfer [360,361]. Ru-TiO2 nanotubes have been widely used for the photocatalytic degradation of organic pollutants under visible or UV light [362,363]. Table 18 presents the summary of recent progress on Ru-TiO2 photocatalysts for organic pollutants degradation.

4.2.5. Modification with Palladium (Pd)

Modification of TiO2 with Palladium (Pd) has also received some attention due to their good stability coupled with high effectivity [367]. Pd is one of the most active elements for interacting with the surface of oxides as support. It has been revealed that the absorbability in the visible light region notably increases with the Pd incorporation; moreover, this metal can act as trapping sites for electrons and hence increase the photocatalytic activity of TiO2. Table 19 presents the summary of recent progress on Pd-TiO2 photocatalysts for organic pollutants degradation.

4.3. Modification with Rare Earth Metals

Modifying with yttrium (Y) was found to be successful for improving the photocatalytic response of TiO2 under visible light irradiations [372,373]. Y-TiO2 can be synthesized using different techniques such as deposition–precipitation with urea, sol–gel, impregnation and chemical coprecipitation method. It has been reported that Y-TiO2 gives improved photocatalytic response attributed to the visible light absorption, electron–hole pairs separation, higher interfacial charge transfer, lower crystallite size, and high specific surface area. Niu et al. [372] found that introducing yttrium shifted the absorption edge of TiO2 towards visible light, reduced the crystallite size, inhibited anatase to rutile phase transformation and decreased the photogenerated electron–hole pair recombination. Table 20 presents the summary of recent progress on Y-TiO2 photocatalysts for organic pollutants degradation.
Modification with lanthanide ions can form complexes with various Lewis bases such as thiols amines, alcohols due to the interaction of their f-orbitals with different functional groups. Adding lanthanides in TiO2 lattice suppresses e to h+ recombination and also increases the adsorptive capacity of the model pollutants [376]. It also stabilizes the mesoporous structure, prevents agglomeration, and increases thermal stability [53,377]. The common lanthanides include lanthanum (La), erbium (Er), neodymium (Nd), gadolinium (Gd), thulium (Tm), ytterbium (Yb), holmium (Ho), terbium (Tb), praseodymium (Pr), samarium (Sm), and europium (Eu).
Among the lanthanide ions, modifying TiO2 with cerium (Ce) has received considerable attention due to the following: (1) it forms labile oxygen vacancies easily with the relatively high mobility of bulk oxygen species (2) the redox couple Ce3+/Ce4+ with the ability of ceria to shift between Ce2O3 and CeO2 under reducing and oxidizing conditions [378,379,380]. This enables Ce to reduce e- to h+ recombination within TiO2 via the trapping of an electron. The cerium ions (Ce3+/Ce4+) have variable valence states with multi-electron energy levels [381]. CeO2 has a strong absorption ability to UV light with a small bandgap [382], its electrons are easy to jump and the 5d and 4f orbitals of Ce4+ are without electrons, which leads to large deformation and strong polarization which allows for effective separation of photoinduced carriers and the improvement of the photocatalytic activity of TiO2 [383]. Ce-TiO2 materials can be prepared by hydrothermal and sol–gel methods [380,384,385,386].
Xie et al. [387] prepared the monodisperse Ce-TiO2 microspheres via a facile solvothermal process. Their results showed that the photocatalytic activity of TiO2 was improved activity for MB under visible light irradiation upon introduction with Ce. Aman et al. [388] presented a 5% Ce-TiO2 and their results show an improved activity for MB (50 ppm) with photodegradation of MB of ~100% within 60 min in the presence of visible light. Table 21 presents the summary of recent progress on lanthanides modified TiO2 photocatalysts for organic pollutants degradation.

4.4. Modification with Other Metal Ions

The other metal ions are also introduced in TiO2 crystal lattices to replace Ti ions, thus improving the transferring rate of interfacial electrons, lowering the recombination between e and h+, and adjusting the energy band structure [408]. Simultaneously, the energy level is introduced into the bandgap of TiO2 to broaden the absorption band edge to the visible light range, and finally, improve the activity of TiO2 under visible light [408,409].
Indium (In)-TiO2 has been observed to increase the photoelectron chemical responses and photocatalytic activity of TiO2 by enhancing the transfer of electrons and reduction in recombination rate of photoinduced charge carriers due to its CB/VB potential difference from TiO2 [410,411]. Indium has lower toxicity, is relatively cheap, and has multiple oxidation states (In°, In+1, In+3), which can help to improve charge mobility and electron trapping over the surface of TiO2 [412]. The result by [413] shows an enhanced photocatalytic activity for 2,4-dichlorophenoxyacetic acid in the presence of UV light irradiation. Pozynak et al. [414] reported an efficient separation of photogenerated charges in nanocrystalline In2O3/TiO2 photocatalyst during the degradation of 2-CP.
The presence of transition metal such as Gallium ion (Ga3+) at the Ti4+ sites can induce oxygen vacancies gaps [415,416] and build deficiencies near the CB in TiO2, which function as electron traps and increase the isolation of e- to h+ pairs [417,418]. Simultaneously, modifying with Ga shift the absorption edge towards visible regime as well as improving the separation between photoexcited charge carriers [419]. The synthesis of Ga–TiO2 can be achieved by laser pyrolysis [420], sol–gel method [417], traditional solid-state reaction [421,422], and hydrothermal method [419,423,424,425]. However, these methods often require either a high experimental temperature to promote the reactions [420] or post heat treatment for crystallization [417,423]. Generally, a higher heating temperature always leads to grain growth and agglomeration, which decreases the specific surface area and is detrimental to photocatalytic activity. Table 22 presents the summary of recent progress on other metals modified TiO2 photocatalysts for organic pollutants degradation.

5. Conclusions and Future Outlooks

TiO2 is the most widely used photocatalyst for the photodegradation of organic pollutants. This paper presents the recent development of metal ion modified TiO2 for the photodegradation of organic pollutants. Among the metal ions, metal such as Ni, Fe, Co, Cu, Au, Ag, Zr, W, and Mn have been widely explored and found to show beneficial influence on photocatalytic activity. However, the application of metal ion modified TiO2 photocatalyst for organic pollutants continues to be limited due to several obstacles (preparation method, security, cost, commercial use, and efficiency). Some of the recommendations drawn from this review are listed as follows:
The activity of TiO2 photocatalyst has a lot to do with its preparation method. The preparation methods are different, and the shape and size, surface, and structural properties of the catalyst are different. The main methods for preparing metal ion modified TiO2 include sol–gel, precipitation, immersion, and hydrothermal method. It was found that the samples prepared by sol–gel and hydrothermal methods produce better results. Its photocatalytic performance is much higher than that of photocatalysts prepared by other methods. This is mainly because the reaction process of these two methods is simple, the operation is controllable, and the prepared powder has a relatively small particle size, high purity, and good chemical uniformity. However, there is a limitation in the long preparation cycle. Therefore, future research should be devoted to new ways to find simple and effective ways to improve some defects of metal ions modified TiO2.
There is a need for continuous in-depth study on the use of sunlight irradiation, and the economical applicability of this approach for removing organic pollutants. To date, various metal ion modified TiO2 photocatalysts have been developed and reported. However, their preparations consume considerable chemicals which are expensive, time-consuming and complex. Traditional TiO2 is generally a massive particle, and the control of the morphology of TiO2 is an effective way to increase their contact with pollutants. For example, the preparation of hollow, porous, or larger specific surface area TiO2 nanoscale particles can effectively solve the traditional small contact surface of TiO2 and pollutants. In addition, TiO2 itself can be modified to add its hydrophilic functional groups. It can increase the compatibility of TiO2 with H2O, thereby promoting the working process of TiO2. Therefore, it is a necessity to develop a relatively scalable, inexpensive, environmentally friendly, and easy synthesis route.
There are also many defects in the application of TiO2 in daily life. The limiting factor for its development is the photocatalytic performance problem. In addition, if the traditional photocatalyst is unmodified, it cannot adapt to the change to other light sources (e.g., sunlight). The water body purified by TiO2 in daily life will take away part of the TiO2, and at the same time, the photocatalytic degradation of organic pollutants often produces some intermediate products. However, in most cases, the toxicity of the intermediate products tends to be stronger than the initial organic pollutants, and as a consequence poses greater harm to the environment and human beings. Therefore, there is a need for an extensive toxicological study of these intermediate products. Therefore, in terms of these issues, how to improve the flexibility of particles in practical applications and how to improve the safety of TiO2 also needs to be considered in future research.
The applicability of this method requires additional engineering amplification testing. However, it is hoped that through rapid and continuous evaluation of the pilot plant configuration, a large-scale solar-driven photocatalytic activity treatment process with low site area requirements and high efficacy can be realized in near future. Furthermore, the majority of the studies use artificial organic pollutants and studying the performance of TiO2 photocatalysts under real situations should be the focus of future work.
Finally, we hope that this paper can assist researchers to better understand the recent trend in the removal of organic pollutants using metal ion modified TiO2 photocatalysts and also hope that this field can flourish in the future.

Author Contributions

Conceptualization, D.J., T.A.O.; methodology, T.A.O., D.J.; software, N.F.S.; resources, S.L., T.A.O., D.J.; writing—original draft preparation, T.A.O., D.J., Y.O.; writing—review and editing, T.A.O., D.J., Y.O., N.F.S.; project administration, S.L., T.A.O.; funding acquisition, S.L.; supervision, S.L., T.A.O., S.W., A.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This review article was supported by the Key Laboratory for Catalyst Synthesis Technology of Polymer of Liaoning Province, China and the Discipline Team Building Program of Shenyang University of Technology. It was also supported by Shenyang University of Technology, China and Universiti Malaysia Kelantan, Malaysia under the postdoctoral scheme.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

OH: hydroxyl radicals; 2,4-CP: 2,4-dichlorophenol; 2-CP: 2-chlorophenol; 4-CP: 4-chlorophenol; AA: Aldrich anatase; Ag: silver; AO7: acid orange 7; AOPs: advanced oxidation processes; AR: Aldrich rutile; AR-27: acid red 27; ASA: acetylsalicylic acid; at. %: atomic %; Au: gold; BB-41: basic blue 41; BO2: basic orange 2; BPA: bisphenol A; CB: conduction band; CBN: carbofuran; CFS: ceftiofur sodium; CIP: ciprofloxacin; Co: cobalt; Cr: chromium; CR: Congo red; Cu: copper; CV: crystal violet; CVD: chemical vapor deposition; DB15: direct blue 15; DCA: dichloroacetic acid; DCF: diclofenac; DEP: diethyl phthalate; DES: diethyl sulfide; DXM: dexamethasone; E131 VF: food colorant; EB: eosin blue; ECR: Eriochrome cyanine red; EE2: 17-α-ethinylestradiol; EISA: evaporation induced self-assembly; ERY: erythromycin; Fe: iron; FSP: flame spray pyrolysis; h+: hole; HPT: highly porous TiO2; IBP: ibuprofen; LPP: liquid phase plasma; LP: laser pyrolysis; MB: methylene blue; MBE: molecular beam epitaxy technique; MEK: methyl ethyl ketone; MG: malachite green; Mn: manganese; MNZ: metronidazole; MO: methyl orange; Mo: molybdenum; MWASG: microwave-assisted sol–gel method; Nb: niobium; NB: nitrobenzene; NHE: normal hydrogen electrode; Ni: nickel; NOR: norfloxacin; NPs: nanoparticles; NPX: naproxen; O2•‒: superoxide anion radicals; OG: orange G; OH: hydroxyl radicals; OOH: hydroperoxyl radicals; P25: Evonik AEROXIDE P25; Pd: palladium; PEC: photoelectrocatalysis with UVA light; PEOx: plasma electrolytic oxidation; Pt: platinum; PNP: p-nitrophenol; RB: reactive blue; RB5: reactive black 5; RhB: rhodamine B; RR 198: reactive red 198; Ru: ruthenium; SMX: sulfamethoxazole; SMZ: sulfamethazine; TB: Terasil blue; TC: tetracycline; TNT: titania nanotubes; TiO2: titanium dioxide; TNAs: titanium nanotube arrays; UAS: ultrasonic aerosol spray; V: vanadium; VB: valence band; W: tungsten; wt.%: mass percent; Zn: zinc; Zr: zirconium.

References

  1. Otitoju, T.A.; Ooi, B.S.; Ahmad, A.L. Synthesis of 3-aminopropyltriethoxysilane-silica modified polyethersulfone hollow fiber membrane for oil-in-water emulsion separation. React. Funct. Polym. 2019, 136, 107–121. [Google Scholar] [CrossRef]
  2. Shoparwe, N.F.; Otitoju, T.A.; Ahmad, A.L. Fouling evaluation of polyethersulfone (PES)/sulfonated cation exchange resin (SCER) membrane for BSA separation. J. Appl. Polym. Sci. 2018, 135, 45854. [Google Scholar] [CrossRef]
  3. Zhong, Q.; Lan, H.; Zhang, M.; Zhu, H.; Bu, M. Preparation of heterostructure g-C3N4/ZnO nanorods for high photocatalytic activity on different pollutants (MB, RhB, Cr(VI) and eosin). Ceram. Int. 2020, 46, 12192–12199. [Google Scholar] [CrossRef]
  4. Wang, Y.; Li, J.; Wang, Q. The performance of daylight photocatalytic activity towards degradation of MB by the flower-like and approximate flower-like complexes of graphene with ZnO and Cerium doped ZnO. Optik 2019, 204, 164131. [Google Scholar] [CrossRef]
  5. Pirgholi Givi, G.; Farjami Shayesteh, S.; Azizian, Y. Investigation of The influence of irradiation intensity and stirring rate on the photocatalytic activity of titanium dioxide nanostructures prepared by the microwave-assisted method for photodegradation of MB from water. Phys. B Condens. Matter 2019, 578, 411886. [Google Scholar] [CrossRef]
  6. Lalliansanga; Tiwari, D.; Tiwari, A.; Shukla, A.; Shim, M.; Lee, S.-M. Facile synthesis and characterization of Ag(NP)/TiO2 nanocomposite: Photocatalytic efficiency of catalyst for oxidative removal of Alizarin Yellow. Catal. Today. [CrossRef]
  7. Dong, S.; Lee, G.J.; Zhou, R.; Wu, J. Synthesis of g-C3N4/BiVO4 Heterojunction Composites for Photocatalytic Degradation of Nonylphenol Ethoxylate. Sep. Purif. Technol. 2020, 250, 117202. [Google Scholar] [CrossRef]
  8. Tian, J.; Zhu, Z.; Liu, B. Novel Bi2MoO6/Bi2WO6/MWCNTs photocatalyst with enhanced photocatalytic activity towards degradation of RB-19 under visible light irradiation. Colloids Surf. A Physicochem. Eng. Asp. 2019, 581, 123798. [Google Scholar] [CrossRef]
  9. Rodrigues, J.; Hatami, T.; Rosa, J.; Tambourgi, E.; Innocentini-Mei, L. Photocatalytic degradation using ZnO for the treatment of RB 19 and RB 21 dyes in industrial effluents and mathematical modeling of the process. Chem. Eng. Res. Des. 2020, 153, 294–305. [Google Scholar] [CrossRef]
  10. Natesan, K.; Sinthiya, M.; Ramamurthi, K.; Ramraj, R.; Raman, S. Visible light driven photocatalytic activity of ZnO/CuO nanocomposites coupled with rGO heterostructures synthesized by solid-state method for RhB dye degradation. Arab. J. Chem. 2019, 13, 3910–3928. [Google Scholar] [CrossRef]
  11. Akpan, U.G.; Hameed, B.H. Parameters affecting the photocatalytic degradation of dyes using TiO2-based photocatalysts: A review. J. Hazard. Mater. 2009, 170, 520–529. [Google Scholar] [CrossRef] [PubMed]
  12. Kang, W.; Yu, H.; Xu, T.; Wu, S.; Wang, X.; Lu, N.; Quan, X.; Liang, H. Photocatalytic ozonation of organic pollutants in wastewater using a flowing through reactor. J. Hazard. Mater. 2020, 405, 124277. [Google Scholar] [CrossRef]
  13. Liao, X.; Wang, F.; Wang, Y.; Cai, Y.; Liu, H.; Wang, X.; Zhu, Y.; Chen, L.; Yao, Y.; Qinglan, H. Constructing Fe-based bi-MOFs for photo-catalytic ozonation of organic pollutants in Fischer-Tropsch waste water. Appl. Surf. Sci. 2020, 509, 145378. [Google Scholar] [CrossRef]
  14. Žener, B.; Matoh, L.; Carraro, G.; Miljevic, B.; Cerc Korošec, R. Sulfur-, nitrogen- and platinum-doped titania thin films with high catalytic efficiency under visible-light illumination. Beilstein J. Nanotechnol. 2018, 9, 1629–1640. [Google Scholar] [CrossRef] [Green Version]
  15. Hasanpour, M.; Hatami, M. Photocatalytic performance of aerogels for organic dyes removal from wastewaters: Review study. J. Mol. Liq. 2020, 309, 113094. [Google Scholar] [CrossRef]
  16. Otitoju, T.A.; Jiang, D.; Ouyang, Y.; Elamin, M.A.M.; Li, S. Photocatalytic degradation of Rhodamine B using CaCu3Ti4O12 embedded polyethersulfone hollow fiber membrane. J. Ind. Eng. Chem. 2020, 83. [Google Scholar] [CrossRef]
  17. Diao, Y.; Yan, Z.; Guo, M.; Wang, X. Magnetic multi-metal co-doped magnesium ferrite nanoparticles: An efficient visible light-assisted heterogeneous Fenton-like catalyst synthesized from saprolite laterite ore. J. Hazard. Mater. 2017, 344, 829–838. [Google Scholar] [CrossRef]
  18. Ensaldo-Rentería, M.K.; Ramírez-Robledo, G.; Sandoval-González, A.; Arellano, C.; Alvarez-Gallegos, A.; Zamudio-Lara, Á.; Silva Martínez, S. Photoelectrocatalytic oxidation of acid green 50 dye in aqueous solution using Ti/TiO2-NT electrode. J. Environ. Chem. Eng. 2018, 6, 1182–1188. [Google Scholar] [CrossRef]
  19. Alijani, M.; Kaleji, B.; Rezaee, S. Highly visible-light active with Co/Sn co-doping of TiO2 nanoparticles for degradation of methylene blue. J. Mater. Sci. Mater. Electron. 2017, 28, 15345–15353. [Google Scholar] [CrossRef]
  20. Wei, Z.; Liu, D.; Wei, W.; Chen, X.; Han, Q.; Yao, W.-Q.; Ma, X. Ultrathin TiO2(B) Nanosheets as the Inductive Agent for Transfering H2O2 into Superoxide Radicals. ACS Appl. Mater. Interfaces 2017, 9, 15533–15540. [Google Scholar] [CrossRef]
  21. Ayode Otitoju, T.; Ugochukwu Okoye, P.; Chen, G.; Li, Y.; Onyeka Okoye, M.; Li, S. Advanced ceramic components: Materials, Fabrication, and Applications. J. Ind. Eng. Chem. 2020, 85, 34–65. [Google Scholar] [CrossRef]
  22. Sangchay, W.; Kaewjang, S. SnO2-doped TiO2 nanostructured thin films with antibacterial properties. AIP Conf. Proc. 2016, 1775, 030023. [Google Scholar]
  23. Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at a Semiconductor Electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  24. Negishi, N.; Iyoda, T.; Hashimoto, K.; Fujishima, A. Preparation of Transparent TiO2 Thin Film Photocatalyst and Its Photocatalytic Activity. Chem. Lett. 1995, 9, 841–842. [Google Scholar] [CrossRef]
  25. Patil, N.; Uphade, B.; Jana, P.; Sonawane, R.; Bhargava, S.; Choudhary, V. Epoxidation of Styrene by Anhydrous t-Butyl Hydroperoxide over Au/TiO2 Catalysts. Catal. Lett. 2004, 94, 89–93. [Google Scholar] [CrossRef]
  26. Baran, W.; Makowski, A.; Wardas, W. The effect of UV radiation absorption of cationic and anionic dye solutions on their photocatalytic degradation in the presence TiO2. Dye. Pigment. 2008, 76, 226–230. [Google Scholar] [CrossRef]
  27. Poulopoulos, S.; Philippopoulos, C. Photo-assisted Oxidation of Chlorophenols in Aqueous Solutions Using Hydrogen Peroxide and Titanium Dioxide. J. Environ. Sci. Health. A Tox. Hazard. Subst. Environ. Eng. 2004, 39, 1385–1397. [Google Scholar] [CrossRef]
  28. Ahmad, R.; Ahmad, Z.; Khan, A.; Mastoi, N.; Aslam, M.; Kim, J. Photocatalytic Systems as an Advanced Environmental Remediation: Recent Developments, Limitations and new Avenues for Applications. J. Environ. Chem. Eng. 2016, 4, 4143–4164. [Google Scholar] [CrossRef]
  29. Mohite, V.S.; Mahadik, M.; Kumbhar, S.S.; Hunge, Y.; Kim, J.-H.; Moholkar, A.; Rajpure, K.; Bhosale, C. Photoelectrocatalytic degradation of benzoic acid using Au doped TiO2 thin films. J. Photochem. Photobiol. B 2014, 142C, 204–211. [Google Scholar] [CrossRef]
  30. Shinde, S.S.; Shinde, P.S.; Bhosale, C.H.; Rajpure, K.Y. Zinc oxide mediated heterogeneous photocatalytic degradation of organic species under solar radiation. J. Photochem. Photobiol. B Biol. 2011, 104, 425–433. [Google Scholar] [CrossRef] [PubMed]
  31. Murtaza, G.; Ahmad, R.; Rashid, M.S.; Hassan, M.; Hussnain, A.; Khan, M.A.; Ehsan ul Haq, M.; Shafique, M.A.; Riaz, S. Structural and magnetic studies on Zr doped ZnO diluted magnetic semiconductor. Curr. Appl. Phys. 2014, 14, 176–181. [Google Scholar] [CrossRef]
  32. Inturi, S.; Boningari, T.; Suidan, M.; Smirniotis, P. Visible-light-induced photodegradation of gas phase acetonitrile using aerosol-made transition metal (V, Cr, Fe, Co, Mn, Mo, Ni, Cu, Y, Ce, and Zr) doped TiO2. Appl. Catal. B Environ. 2014, 144, 333–342. [Google Scholar] [CrossRef]
  33. Choina, J.; Fischer, C.; Flechsig, G.-U.; Kosslick, H.; Tuan, V.; Nguyen, T.; Tuyen, N.A.; Schulz, A. Photocatalytic properties of Zr-doped titania in the degradation of the pharmaceutical ibuprofen. J. Photochem. Photobiol. A Chem. 2014, 274, 108–116. [Google Scholar] [CrossRef]
  34. Kim, T.-H.; Lee, Y.; Han, S.-H.; Hwang, S.-J. The effects of wavelength and wavelength mixing ratios on microalgae growth and nitrogen, phosphorus removal using Scenedesmus sp. for wastewater treatment. Bioresour. Technol. 2013, 130, 75–80. [Google Scholar] [CrossRef] [PubMed]
  35. Saravanan, R.; Sacari, E.; Gracia, F.; Khan, M.M.; Mosquera, E.; Gupta, V.K. Conducting PANI stimulated ZnO system for visible light photocatalytic degradation of coloured dyes. J. Mol. Liq. 2016, 221, 1029–1033. [Google Scholar] [CrossRef]
  36. Saravanan, R.; Gupta, V.K.; Mosquera, E.; Gracia, F.; Narayanan, V.; Stephen, A. Visible light induced degradation of methyl orange using β-Ag0.333V2O5 nanorod catalysts by facile thermal decomposition method. J. Saudi Chem. Soc. 2015, 19, 521–527. [Google Scholar] [CrossRef] [Green Version]
  37. Saravanan, R.; Gracia, F.; Khan, M.M.; Poornima, V.; Gupta, V.K.; Narayanan, V.; Stephen, A. ZnO/CdO nanocomposites for textile effluent degradation and electrochemical detection. J. Mol. Liq. 2015, 209, 374–380. [Google Scholar] [CrossRef]
  38. Gnanasekaran, L.; Hemamalini, R.; R, S.; Ravichandran, K.; Gracia, F.; Gupta, V. 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]
  39. Choi, W.; Termin, A.; Hoffmann, M. 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]
  40. Zhang, H.; Chen, G.; Bahnemann, D. Photoelectrocatalytic Materials for Environmental Applications. J. Mater. Chem. 2009, 19, 5089–5121. [Google Scholar] [CrossRef]
  41. Graetzel, M.; Howe, R. Electron paramagnetic resonance studies of doped titanium dioxide colloids. J. Phys. Chem. 1990, 94, 2566–2572. [Google Scholar] [CrossRef]
  42. Schneider, J.; Matsuoka, M.; Takeuchi, M.; Zhang, J.; Horiuchi, Y.; Anpo, M.; Bahnemann, D. Understanding TiO2 Photocatalysis: Mechanisms and Materials. Chem. Rev. 2014, 114, 919–9986. [Google Scholar] [CrossRef]
  43. Dvoranová, D.; Brezová, V.; Mazúr, M.; Malati, M. Investigations of Meta-Doped Titanium Dioxide Photocatalysis. Appl. Catal. B Environ. 2002, 37, 91–105. [Google Scholar] [CrossRef]
  44. Wang, Y.; Zhang, R.; Li, J.; Li, L.; Lin, S. First-principles study on transition metal-doped anatase TiO2. Nanoscale Res. Lett. 2014, 9, 46. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Jun, T.; Lee, K.-S. Cr-doped TiO2 thin films deposited by RF-sputtering. Mater. Lett. 2010, 64, 2287–2289. [Google Scholar] [CrossRef]
  46. Qin, L.Z.; Liang, H.; Liao, B.; Liu, A.D.; Wu, X.Y.; Sun, J. Photocatalytic performance of Fe-, Ni-, or Cu-ion implanted TiO2 films under UV light, visible light and sunlight irradiation. Nucl. Instrum. Methods Phys. Res. Sect. B Beam Interact. Mater. At. 2013, 307, 385–390. [Google Scholar] [CrossRef]
  47. Leyland, N.; Carroll (Podporska), J.; Browne, J.; Hinder, S.; Quilty, B.; Pillai, S. Highly Efficient F, Cu doped TiO2 anti-bacterial visible light active photocatalytic coatings to combat hospital-acquired infections. Sci. Rep. 2016, 6, 24770. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  48. Mathew, S.; Ganguly, P.; Rhatigan, S.; Vignesh, K.; Byrne, C.; Hinder, S.; Bartlett, J.; Nolan, M.; Pillai, S. Cu-Doped TiO2: Visible Light Assisted Photocatalytic Antimicrobial Activity. Appl. Sci. 2018, 8, 2067. [Google Scholar] [CrossRef] [Green Version]
  49. Ibram, G.; Polkampally, P.; Annapoorna, I.; Sumliner, J.; Ramakrishna, M.; Hebalkar, N.; Padmanabham, G.; Sundararajan, G. Preparation and characterization of Cu-doped TiO2 materials for electrochemical, photoelectrochemical, and photocatalytic applications. Appl. Surf. Sci. 2014, 293, 229–247. [Google Scholar] [CrossRef]
  50. Choi, J.; Park, H.; Hoffmann, M. Effects of Single Metal-Ion Doping on the Visible-Light Photoreactivity of TiO2. J. Phys. Chem. C 2009, 114, 783–792. [Google Scholar] [CrossRef] [Green Version]
  51. Murakami, N.; Chiyoya, T.; Tsubota, T.; Ohno, T. Switching redox site of photocatalytic reaction on titanium (IV) oxide particles modified with transition-metal ion controlled by irradiation wavelength. Appl. Catal. A Gen. 2008, 348, 148–152. [Google Scholar] [CrossRef] [Green Version]
  52. Rauf, M.A.; Meetani, M.A.; Hisaindee, S. An overview on the photocatalytic degradation of azo dyes in the presence of TiO2 doped with selective transition metals. Desalination 2011, 276, 13–27. [Google Scholar] [CrossRef]
  53. Nešić, J.; Manojlović, D.; Andelković, I.; Dojcinovic, B.; Vulić, P.; Krstic, J.; Roglic, G. Preparation, characterization and photocatalytic activity of lanthanum and vanadium co-doped mesoporous TiO2 for azo-dye degradation. J. Mol. Catal. A Chem. 2013, 378, 67–75. [Google Scholar] [CrossRef] [Green Version]
  54. Ni, M.; Leung, M.K.H.; Leung, D.; Sumathy, K. A Review and Recent Developments in Photocatalytic Water-Splitting Using TiO2 for Hydrogen Production. Renew. Sustain. Energy Rev. 2007, 11, 401–425. [Google Scholar] [CrossRef]
  55. Chang, S.-M.; Liu, W. Surface doping is more beneficial than bulk doping to the photocatalytic activity of vanadium-doped TiO2. Appl. Catal. B Environ. 2011, 101, 333–342. [Google Scholar] [CrossRef]
  56. Marschall, R. Photocatalysis: Semiconductor Composites: Strategies for Enhancing Charge Carrier Separation to Improve Photocatalytic Activity. Adv. Funct. Mater. 2014, 24, 2421–2440. [Google Scholar] [CrossRef]
  57. Dong, H.; Zeng, G.; Tang, L.; Fan, C.; Zhang, C.; He, X.; He, Y. An overview on limitations of TiO2-based particles for photocatalytic degradation of organic pollutants and the corresponding countermeasures. Water Res. 2015, 79, 128–146. [Google Scholar] [CrossRef]
  58. Mutlu, E.; Cristy, T.; Graves, S.; Hooth, M.; Waidyanatha, S. Characterization of aqueous formulations of tetra- and pentavalent forms of vanadium in support of test article selection in toxicology studies. Environ. Sci. Pollut. Res. 2017, 24, 405–416. [Google Scholar] [CrossRef]
  59. Ntobeng, M.; Imoisili, P.; Jen, T.-C. Facile fabrication of vanadium sensitized silver titanium oxides (V–Ag/TiO2) photocatalyst nanocomposite for pollutants removal in river water. Mater. Sci. Semicond. Process. 2020, 123, 105569. [Google Scholar] [CrossRef]
  60. Mitoraj, D.; Kisch, H. Surface Modified Titania Visible Light Photocatalyst Powders. Solid State Phenom. 2010, 162, 49–75. [Google Scholar] [CrossRef]
  61. Phung, H.; Truong, N.; Lâm, N.T.; Phan Thi, K.L.; Duong, P.; Le, H. Enhancement of the visible light photocatalytic activity of vanadium and nitrogen co-doped TiO2 thin film. J. Nonlinear Opt. Phys. Mater. 2015, 25, 1550052. [Google Scholar] [CrossRef]
  62. Weng, K.-W.; Han, S. Photocatalytic performance of TiVOx/TiO2 thin films prepared by bipolar pulsed magnetron sputter deposition. J. Vac. Sci. Technol. B Nanotechnol. Microelectron. 2017, 35, 41201–41207. [Google Scholar] [CrossRef]
  63. Shao, G.; Imran, S.; Jeon, S.J.; Kang, S.J.; Haider, M.; Kim, H.T. Sol–gel synthesis of vanadium doped titania: Effect of the synthetic routes and investigation of their photocatalytic properties in the presence of natural sunlight. Appl. Surf. Sci. 2015, 351, 1213–1223. [Google Scholar] [CrossRef]
  64. Zhou, J.; Takeuchi, M.; Ray, A.; Anpo, M.; Zhao, X.S. Enhancement of photocatalytic activity of P25 TiO2 by vanadium-ion implantation under visible light irradiation. J. Colloid Interface Sci. 2007, 311, 497–501. [Google Scholar] [CrossRef]
  65. Bettinelli, M.; Dallacasa, V.; Falcomer, D.; Fornasiero, P.; Gombac, V.; Montini, T.; Romanò, L.; Speghini, A. Photocatalytic activity of TiO2 doped with boron and vanadium. J. Hazard. Mater. 2007, 146, 529–534. [Google Scholar] [CrossRef]
  66. Kubacka, A.; Fuerte, A.; Martínez-Arias, A.; Fernández-García, M. Nanosized Ti–V mixed oxides: Effect of doping level in the photo-catalytic degradation of toluene using sunlight-type excitation. Appl. Catal. B Environ. 2007, 74, 26–33. [Google Scholar] [CrossRef]
  67. Patel, S.; Gajbhiye, N. Intrinsic room-temperature ferromagnetism of V-doped TiO2 (B) nanotubes synthesized by the hydrothermal method. Solid State Commun. 2011, 151, 1500–1503. [Google Scholar] [CrossRef]
  68. Khan, M.; Song, Y.; Chen, N.; Cao, W. Effect of V doping concentration on the electronic structure, optical and photocatalytic properties of nano-sized V-doped anatase TiO2. Mater. Chem. Phys. 2013, 142, 148–153. [Google Scholar] [CrossRef]
  69. Gu, D.-E.; Yang, B.-C.; Hu, Y.-D. A Novel Method for Preparing V-doped Titanium Dioxide Thin Film Photocatalysts with High Photocatalytic Activity Under Visible Light Irradiation. Catal. Lett. 2007, 118, 254–259. [Google Scholar] [CrossRef]
  70. Iketani, K.; Sun, R.-D.; Toki, M.; Hirota, K.; Yamaguchi, O. Sol–gel-derived VXTi1−XO2 films and their photocatalytic activities under visible light irradiation. Mater. Sci. Eng. B 2004, 108, 187–193. [Google Scholar] [CrossRef]
  71. Lu, D.; Zhao, B.; Fang, P.; Zhai, S.; Li, D.; Chen, Z.; Wu, W.; Chai, W.; Wu, Y.; Qi, N. Facile one-pot fabrication and high photocatalytic performance of vanadium doped TiO2-based nanosheets for visible-light-driven degradation of RhB or Cr(VI). Appl. Surf. Sci. 2015, 359, 435–448. [Google Scholar] [CrossRef]
  72. Camposeco, R.; Castillo, S.; Hinojosa Reyes, M.; Mejía-Centeno, I.; Zanella, R. Effect of incorporating vanadium oxide to TiO2, Zeolite-ZM5, SBA and P25 supports on the photocatalytic activity under visible light. J. Photochem. Photobiol. A Chem. 2018, 367, 178–187. [Google Scholar] [CrossRef]
  73. Vasilic, R.; Stojadinović, S.; Radic, N.; Stefanov, P.; Dohčević-Mitrović, Z.; Grbic, B. One-step preparation and photocatalytic performance of vanadium doped TiO2 coatings. Mater. Chem. Phys. 2015, 151, 337. [Google Scholar] [CrossRef]
  74. Wang, B.; Zhang, G.; Sun, Z.; Zheng, S. A comparative study about the influence of metal ions (Ce, La and V) doping on the solar-light-induced photodegradation toward rhodamine B. J. Environ. Chem. Eng. 2015, 3. [Google Scholar] [CrossRef]
  75. Sinirtas, E.; Isleyen, M.; Pozan, G. Photocatalytic degradation of 2,4-dichlorophenol with V2O5-TiO2 catalysts: Effect of catalyst support and surfactant additives. Chinese J. Catal. 2016, 37, 607–615. [Google Scholar] [CrossRef]
  76. Palanivel, M.; Lokeshkumar, E.; Amruthaluru, S.; Govardhanan, B.; Mahalingam, A.; Rameshbabu, N. Visible light photocatalytic activity of metal (Mo/V/W) doped porous TiO2 coating fabricated on Cp-Ti by plasma electrolytic oxidation. J. Alloys Compd. 2020, 825, 154092. [Google Scholar] [CrossRef]
  77. Khan, H.; Berk, D. Characterization and mechanistic study of Mo+6 and V+5 codoped TiO2 as a photocatalyst. J. Photochem. Photobiol. A Chem. 2014, 294, 96–109. [Google Scholar] [CrossRef]
  78. Murakami, E.; Mizoguchi, R.; Yoshida, Y.; Kitashoji, A.; Nakashima, N.; Yatsuhashi, T. Multiple strong field ionization of metallocenes: Applicability of ADK rates to the production of multiply charged transition metal (Cr, Fe, Ni, Ru, Os) cations. J. Photochem. Photobiol. A Chem. 2018, 369, 16–24. [Google Scholar] [CrossRef]
  79. Elahifard, M.; Ahmadvand, S.; Mirzanejad, A. Effects of Ni-doping on the photo-catalytic activity of TiO2 anatase and rutile: Simulation and experiment. Mater. Sci. Semicond. Process. 2018, 84, 10–16. [Google Scholar] [CrossRef]
  80. Kwak, B.; Vignesh, K.; Park, N.-K.; Ryu, H.-J.; Baek, J.-I.; Kang, M. Methane formation from photoreduction of CO2 with water using TiO2 including Ni ingredient. Fuel 2015, 143, 570–576. [Google Scholar] [CrossRef]
  81. Ranjitha, A.; Muthukumarasamy, N.; Thambidurai, M.; Velauthapillai, D.; Balasundaraprabhu, R.; Agilan, S. Fabrication of Ni-doped TiO2 thin film photoelectrode for solar cells. Sol. Energy 2014, 106, 159–165. [Google Scholar] [CrossRef]
  82. Devi, L.; Kottam, N.; Murthy, B.; Kumar, S. Enhanced photocatalytic activity of transition metal ions Mn2+, Ni2+ and Zn2+ doped polycrystalline titania for the degradation of Aniline Blue under UV/solar light. J. Mol. Catal. A Chem. 2010, 328, 44–52. [Google Scholar] [CrossRef]
  83. Shaban, M.; Ahmed, A.; Shehata, N.; Betiha, M.; Rabie, A. 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]
  84. Chaudhari, S.; Gawal, P.; Sane, P.; Sontakke, S.; Nemade, P. Solar light-assisted photocatalytic degradation of methylene blue with Mo/TiO2: A comparison with Cr- and Ni-doped TiO2. Res. Chem. Intermed. 2018, 44, 3115–3134. [Google Scholar] [CrossRef]
  85. Anju, K.R.; Radhik, T.; Raja, J.; Al-Lohedan, H. Hydrothermal synthesis of nanosized (Fe, Co, Ni)-TiO2 for enhanced visible light photosensitive applications. Optik 2018, 165. In Press. [Google Scholar] [CrossRef]
  86. Li, X.; Wu, Y.; Shen, Y.; Sun, Y.; Yang, Y.; Xie, A. A novel bifunctional Ni-doped TiO2 inverse opal with enhanced SERS performance and excellent photocatalytic activity. Appl. Surf. Sci. 2017, 427, 739–744. [Google Scholar] [CrossRef]
  87. Blanco-Vega, M.P.; Guzmán Mar, J.; Villanueva, M.; Maya-Treviño, L.; Garza-Tovar, L.L.; Hernandez-Ramírez, A.; Reyes, L. Photocatalytic elimination of bisphenol A under visible light using Ni-doped TiO2 synthesized by microwave assisted sol-gel method. Mater. Sci. Semicond. Process. 2017, 71, 275–282. [Google Scholar] [CrossRef]
  88. 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]
  89. Manjakuppam, M.; Rao, C.; Das, R.; Giri, A.; Golder, A. Evaluation of bimetal doped TiO2 in dye fragmentation and its comparison to mono-metal doped and bare catalysts. Appl. Surf. Sci. 2016, 368, 316–324. [Google Scholar] [CrossRef]
  90. Baygi, N.J.; Saghir, A.V.; Beidokhti, S.M.; Khaki, J.V. Modified auto-combustion synthesis of mixed-oxides TiO2/NiO nanoparticles: Physical properties and photocatalytic performance. Ceram. Int. 2020, 46, 15417–15437. [Google Scholar] [CrossRef]
  91. Mohseni-Salehi, M.; Taheri-Nassaj, E.; Zori, M. Effect of dopant (Co, Ni) concentration and hydroxyapatite compositing on photocatalytic activity of titania towards dye degradation. J. Photochem. Photobiol. A Chem. 2017, 356, 57–70. [Google Scholar] [CrossRef]
  92. Elmragui, A.; Zegaoui, O.; Daou, I. Synthesis, characterization and photocatalytic properties under visible light of doped and co-doped TiO2-based nanoparticles. Mater. Today Proc. 2019, 13, 857–865. [Google Scholar] [CrossRef]
  93. Hinojosa Reyes, M.; Solis, R.; Ruiz, F.; Glez, V.; Moctezuma, E. Promotional effect of metal doping on nanostructured TiO2 during the photocatalytic degradation of 4-chlorophenol and naproxen sodium as pollutants. Mater. Sci. Semicond. Process. 2019, 100, 130–139. [Google Scholar] [CrossRef]
  94. Kerkez, Ö.; Kibar, E.; Dayıoğlu, K.; Gedik, F.; Akin, A.; Özkara-Aydınoğlu, A.P.Ş. A comparative study for removal of different dyes over M/TiO2 (M=Cu, Ni, Co, Fe, Mn and Cr) photocatalysts under visible light irradiation. J. Photochem. Photobiol. A Chem. 2015, 311, 176–185. [Google Scholar] [CrossRef]
  95. Yang, X.-J.; Wang, S.; Sun, H.; Wang, X.-B.; Lian, J.-S. Preparation and photocatalytic performance of Cu-doped TiO2 nanoparticles. Trans. Nonferrous Met. Soc. China 2015, 25, 504–509. [Google Scholar] [CrossRef]
  96. Tsai, C.-Y.; Hsi, H.-C.; Kuo, T.-H.; Chang, Y.-M.; Liou, J.-H. Preparation of Cu-Doped TiO2 Photocatalyst with Thermal Plasma Torch for Low-Concentration Mercury Removal. Aerosol Air Qual. Res. 2013, 13, 639–648. [Google Scholar] [CrossRef]
  97. Yang, J.; Ganesan, P.; Teuscher, J.; Moehl, T.; Kim, Y.J.; Yi, C.; Comte, P.; Pei, K.; Holcombe, T.; Nazeeruddin, M.; et al. Influence of the Donor Size in D−π–A Organic Dyes for Dye-Sensitized Solar Cells. J. Am. Chem. Soc. 2014, 136, 5722–5730. [Google Scholar] [CrossRef]
  98. Pava-Gómez, B.; Vargas-Ramírez, X.; Díaz-Uribe, C. Physicochemical study of adsorption and photodegradation processes of methylene blue on copper-doped TiO2 films. J. Photochem. Photobiol. A Chem. 2018, 360, 13–25. [Google Scholar] [CrossRef]
  99. Yusoff, R.; Shamiri, A.; Aroua, M.; Ahmady, A.; Shafeeyan, M.S.; Lee, W.; Lim, S.; Burhanuddin, S.N.M. Physical properties of aqueous mixtures of N-methyldiethanolamine (MDEA) and ionic liquids. J. Ind. Eng. Chem. 2014, 20, 3349–3355. [Google Scholar] [CrossRef]
  100. Mohan, R.; Krishnamoorthy, K.; Kim, S.-J. Enhanced photocatalytic activity of Cu-doped ZnO nanorods. Solid State Commun. 2012, 152, 375–380. [Google Scholar] [CrossRef]
  101. Hanaor, D.; Sorrell, C. Review of the Anatase to Rutile Phase Transformation. J. Mater. Sci. 2011, 46, 855–874. [Google Scholar] [CrossRef] [Green Version]
  102. López, R.; Gomez, R.; Llanos, M. Photophysical and photocatalytic properties of nanosized copper-doped titania sol–gel catalysts. Catal. Today 2009, 148, 103–108. [Google Scholar] [CrossRef]
  103. Xia, X.H.; Gao, Y.; Wang, Z.; Jia, Z.J. Structure and photocatalytic properties of copper-doped rutile TiO2 prepared by a low-temperature process. J. Phys. Chem. Solids 2008, 69, 2888–2893. [Google Scholar] [CrossRef]
  104. Li, G.; Dimitrijevic, N.; Chen, L.; Rajh, T.; Gray, K. Role of Surface/Interfacial Cu2+ Sites in the Photocatalytic Activity of Coupled CuO-TiO2 Nanocomposites. J. Phys. Chem. C 2008, 112, 19040–19044. [Google Scholar] [CrossRef]
  105. Morikawa, T.; Irokawa, Y.; Ohwaki, T. Enhanced photocatalytic activity of TiO2−xNx loaded with copper ions under visible light irradiation. Appl. Catal. A Gen. 2006, 314, 123–127. [Google Scholar] [CrossRef]
  106. Colón, G.; Maicu, M.; Hidalgo, M.C.; Navio, J.A. Cu-doped TiO2 systems with improved photocatalytic activity. Appl. Catal. B Environ. 2006, 67, 41. [Google Scholar] [CrossRef]
  107. Jiang, Y.; Yang, Y.; Qiang, L.; Fan, R.; Ning, H.; Li, L.; Ye, T.; Yang, B.; Cao, W. Based on Cu(II) silicotungstate modified photoanode with long electron lifetime and enhanced performance in dye sensitized solar cells. J. Power Sources 2015, 278, 527–533. [Google Scholar] [CrossRef]
  108. Kim, C.-S.; Shin, J.-W.; Cho, Y.-H.; Jang, H.D.; Byun, H.-S.; Kim, T. Synthesis and characterization of Cu/N-doped mesoporous TiO2 visible light photocatalysts. Appl. Catal. A Gen. 2013, 455, 211–218. [Google Scholar] [CrossRef]
  109. Habibi, M.; Karimi, B.; Zendehdel, M.; Habibi, M. Fabrication, characterization of two nano-composite CuO-ZnO working electrodes for dye-sensitized solar cell. Spectrochim. Acta. A Mol. Biomol. Spectrosc. 2013, 116C, 374–380. [Google Scholar] [CrossRef]
  110. Liu, M.; Qiu, X.; Miyauchi, M.; Hashimoto, K. Cu(II) Oxide Amorphous Nanoclusters Grafted Ti3+ Self-Doped TiO2: An Efficient Visible Light Photocatalyst. Chem. Mater. 2011, 23, 5282–5286. [Google Scholar] [CrossRef]
  111. Hu, Q.; Huang, J.; Li, G.; Chen, J.; Zhaojun, Z.; Deng, Z.; Jiang, Y.; Guo, W.; Cao, Y. Effective water splitting using CuOx/TiO2 composite films: Role of Cu species and content in hydrogen generation. Appl. Surf. Sci. 2016, 369, 201–206. [Google Scholar] [CrossRef]
  112. Ravichandra, D.; Rao, T.; Kim, H.-S.; Pammi, S.V.N.; Prabhakarrao, N.; Imandi, M.R. Hybrid copper doped titania/polythiophene nanorods as efficient visible light-driven photocatalyst for degradation of organic pollutants. J. Asian Ceram. Soc. 2017, 5, 436–443. [Google Scholar] [CrossRef]
  113. Khairy, M.; Zakaria, W. Effect of metal-doping of TiO2 nanoparticles on their photocatalytic activities toward removal of organic dyes. Egypt. J. Pet. 2014, 23, 419–426. [Google Scholar] [CrossRef] [Green Version]
  114. Mahy, J.; Tilkin, R.; Douven, S.; Lambert, S. TiO2 nanocrystallites photocatalysts modified with metallic species: Comparison between Cu and Pt doping. Surfaces Interfaces 2019, 66, 100366. [Google Scholar] [CrossRef]
  115. Jaiswal, R.; Bharambe, J.; Patel, N.; Dashora, A.; Kothari, D.C.; Miotello, A. Copper and Nitrogen co-doped TiO2 photocatalyst with enhanced optical absorption and catalytic activity. Appl. Catal. B Environ. 2015, 168–169, 333–341. [Google Scholar] [CrossRef]
  116. Makki, F.; El Hajj Hassan, M.; El Jamal, M.; Tabatabai-Yazdi, F.-S.; Ebrahimian, A. Kinetic evaluation of photocatalytic degradation of food colorant E 131 VF by copper doped TiO2 nanophotocatalysts prepared at different calcination temperatures. Environ. Technol. Innov. 2020, 19, 100981. [Google Scholar] [CrossRef]
  117. Qian, S.; Zhang, Y.; Wang, P.; Bai, Y.; Lai, B. New insights on the enhanced non-hydroxyl radical contribution under copper promoted TiO2/GO for the photodegradation of tetracycline hydrochloride. J. Environ. Sci. 2021, 100, 99–109. [Google Scholar] [CrossRef]
  118. Medina Ramirez, I.; Rodil, S.; Pedroza-Herrera, G.; Lozano-Alvarez, J. Evaluation of the Photocatalytic Activity of Copper Doped TiO2 nanoparticles for the Purification and/or Disinfection of Industrial Effluents. Catal. Today 2019, 341, 37–48. [Google Scholar] [CrossRef]
  119. Hernández, J.; Coste, S.; García Murillo, A.; Carrillo-Romo, F.; Kassiba, A. Effects of metal doping (Cu, Ag, Eu) on the electronic and optical behavior of nanostructured TiO2. J. Alloys Compd. 2017, 710, 355–363. [Google Scholar] [CrossRef]
  120. Reda, S.M.; Khairy, M.; Mousa, M. Photocatalytic Activity of Nitrogen and Copper Doped TiO2 Nanoparticles Prepared by Microwave-Assisted Sol-Gel Process. Arab. J. Chem. 2017, 13, 86–95. [Google Scholar] [CrossRef]
  121. Hamadanian, M.; Karimzadeh, S.; Jabbari, V.; Villagran, D. Synthesis of cysteine, cobalt and copper-doped TiO2 nanophotocatalysts with excellent visible-light-induced photocatalytic activity. Mater. Sci. Semicond. Process. 2016, 41, 168–176. [Google Scholar] [CrossRef]
  122. Lin, J.; Sopajaree, K.; Jitjanesuwan, T.; Lu, M.-C. Application of visible light on copper-doped titanium dioxide catalyzing degradation of chlorophenols. Sep. Purif. Technol. 2017, 191, 233–243. [Google Scholar] [CrossRef]
  123. Ahmad, S.; Yasin, A. Photocatalytic degradation of deltamethrin by using Cu/TiO2/bentonite composite. Arab. J. Chem. 2020, 13, 8481–8488. [Google Scholar] [CrossRef]
  124. Dorraj, M.; Alizadeh, M.; Sairi, A.; Basirun, W.; Goh, B.T.; Woi, P.; Alias, Y. Enhanced Visible Light Photocatalytic Activity of Copper-doped Titanium Oxide–Zinc Oxide Heterojunction for Methyl Orange Degradation. Appl. Surf. Sci. 2017, 414, 251–261. [Google Scholar] [CrossRef]
  125. Bensouici, F.; Bououdina, M.; Tala-Ighil, R.; Tounane, M.; Iratni, A.; Souier, M.; Shengwen, L.; Cai, W. Optical, Structural and Photocatalysis Properties of Cu-Doped TiO2 Thin Films. Appl. Surf. Sci. 2016, 395, 110–116. [Google Scholar] [CrossRef]
  126. Reddam, H.; Elmail, R.; Lloria, S.; Monros, G.; Reddam, Z.; Coloma-Pascual, F. Synthesis of Fe, Mn and Cu modified TiO2 photocatalysts for photodegradation of Orange II. Boletín la Soc. Española Cerámica y Vidr. 2019, 59, 138–148. [Google Scholar] [CrossRef]
  127. Mahy, J.; Lambert, S.; Leonard, G.; Zubiaur, A.; Olu, P.-Y.; Mahmoud, A.; Boschini, F.; Heinrichs, B. Towards a large scale aqueous sol-gel synthesis of doped TiO2: Study of various metallic dopings for the photocatalytic degradation of p-nitrophenol. J. Photochem. Photobiol. A Chem. 2016, 329, 189–202. [Google Scholar] [CrossRef]
  128. Periyakaruppan, K.; Nisha, N.; Pugazhendhi, A.; Kandasamy, S.; Pitchaimuthu, S. An investigation of transition metal doped TiO2 photocatalysts for the enhanced photocatalytic decoloration of methylene blue dye under visible light irradiation. J. Environ. Chem. Eng. 2021, 9, 105254. [Google Scholar] [CrossRef]
  129. Paul, S.; Chetri, P.; Choudhury, A. Effect of manganese doping on the optical property and photocatalytic activity of nanocrystalline titania: Experimental and theoretical investigation. J. Alloys Compd. 2014, 583, 578–586. [Google Scholar] [CrossRef]
  130. Duan, J.; Wang, H.; Wang, H.; Zhang, J.; Wu, S.; Wang, Y. Mn-doped ZnO nanotubes: From facile solution synthesis to room temperature ferromagnetism. CrystEngComm 2012, 14, 1330–1336. [Google Scholar] [CrossRef]
  131. Deng, Q.; Xia, X.; Guo, M.; Gao, Y. Mn-doped TiO2 nanopowders with remarkable visible light photocatalytic activity. Mater. Lett. 2011, 65, 2051–2054. [Google Scholar] [CrossRef]
  132. Liu, M.; Du, Y.; Ma, L.; Jing, D.; Guo, L. Manganese doped cadmium sulfide nanocrystal for hydrogen production from water under visible light. Int. J. Hydrogen Energy 2012, 37, 730. [Google Scholar] [CrossRef]
  133. Pérez Larios, A.; Lartundo, L.; Mantilla, A.; Mendoza, G.; Gomez, R.; Hernández-Gordillo, L. Enhancing the H2 evolution from water-methanol solution using Mn2+–Mn+3–Mn4+ redox species of Mn-doped TiO2 sol-gel photocatalysts. Catal. Today 2016. [Google Scholar] [CrossRef]
  134. Shu, Y.; Xu, Y.; Huang, H.; Ji, J.; Liang, S.; Wu, M.; Leung, D. Catalytic oxidation of VOCs over Mn/TiO2/activated carbon under 185 nm VUV irradiation. Chemosphere 2018, 208, 550–558. [Google Scholar] [CrossRef] [PubMed]
  135. Devi, L.; Kumar, S.; Murthy, B.; Kottam, N. Influence of Mn2+ and Mo6+ dopants on the phase transformations of TiO2 lattice and its photo catalytic activity under solar illumination. Catal. Commun. 2009, 10, 794–798. [Google Scholar] [CrossRef]
  136. Faouzi, A.; Corbel, S.; Balan, L.; Mozet, K.; Girot, E.; Medjahdi, G.; Ben Said, M.; Ghrabi, A.; Schneider, R. Porous Mn-doped ZnO nanoparticles for enhanced solar and visible light photocatalysis. Mater. Des. 2016, 101, 309–316. [Google Scholar] [CrossRef]
  137. Reddy, C.; Reddy, I.; Akkinepally, B.; Harish, V.; Reddy, R.; Jaesool, S. Mn-doped ZrO2 nanoparticles prepared by a template-free method for electrochemical energy storage and abatement of dye degradation. Ceram. Int. 2019, 45, 15298–15306. [Google Scholar] [CrossRef]
  138. Bharati, B.; Mishra, N.C.; Sinha, A.S.K.; Rath, C. Unusual Structural Transformation and Photocatalytic Activity of Mn Doped TiO2 Nanoparticles under Sunlight. Mater. Res. Bull. 2019, 123, 110710. [Google Scholar] [CrossRef]
  139. Xu, Z.; Li, C.; Fu, N.; Li, W.; Zhang, G. Facile synthesis of Mn-doped TiO2 nanotubes with enhanced visible light photocatalytic activity. J. Appl. Electrochem. 2018, 48, 1197–1203. [Google Scholar] [CrossRef]
  140. Choudhury, B.; Paul, S.; Ahmed, G.; Choudhury, A. Adverse effect of Mn doping on the magnetic ordering in Mn doped TiO2 nanoparticles. Mater. Res. Express 2015, 2, 96104. [Google Scholar] [CrossRef]
  141. Shu, Y.; Ji, J.; Xu, Y.; Deng, J.; Huang, H.; He, M.; Leung, D.; Wu, M.; Liu, S.; Liu, S.; et al. Promotional role of Mn doping on catalytic oxidation of VOCs over mesoporous TiO2 under Vacuum Ultraviolet (VUV) irradiation. Appl. Catal. B Environ. 2017, 220, 78–87. [Google Scholar] [CrossRef]
  142. Islam, M.; Bredow, T. Rutile Band-Gap States Induced by Doping with Manganese in Various Oxidation States. J. Phys. Chem. C 2015, 119, 5534–5541. [Google Scholar] [CrossRef]
  143. Zhang, L.; Zhang, Y.; Ma, Z. Effect of manganese on photocatalysis. China Acad. J. 2011, 4–6, 15. [Google Scholar]
  144. Binas, V.; Sambani, K.; Maggos, T.; Katsanaki, A.; Kiriakidis, G. Synthesis and photocatalytic activity of Mn-doped TiO2 nanostructured powders under UV and visible light. Appl. Catal. B Environ. 2012, 113–114, 79–86. [Google Scholar] [CrossRef]
  145. Esparza, P.; Hernández, T.; Borges, M.E.; Alvarez-Galvan, M.; Ruiz-Morales, J.C.; Fierro, J.L.G. TiO2 modifications by hydrothermal treatment and doping to improve its photocatalytic behaviour under visible light. Catal. Today 2013, 210, 135–141. [Google Scholar] [CrossRef]
  146. Feng, H.; Yu, L.; Zhang, M.-H. Ultrasonic synthesis and photocatalytic performance of metal-ions doped TiO2 catalysts under solar light irradiation. Mater. Res. Bull. 2013, 48, 672–681. [Google Scholar] [CrossRef]
  147. Park, J.; Anh Le, H.; Kim, Y.-S.; Chin, S.-M.; Bae, G.-N.; Jurng, J. Synthesis and enhanced photocatalytic activity of Mn/TiO2 mesoporous materials using the impregnation method through CVC process. J. Porous Mater. 2012, 19, 877–881. [Google Scholar] [CrossRef]
  148. Khraisheh, M.; Wu, L.; Al-Muhtaseb, A.; Albadarin, A.; Walker, G. Phenol degradation by powdered metal ion modified titanium dioxide photocatalysts. Chem. Eng. J. 2012, 213, 125–134. [Google Scholar] [CrossRef] [Green Version]
  149. Zou, X.; Li, G.-D.; Guo, M.; Li, X.-H.; Liu, D.-P.; Su, J.; Chen, J. Heterometal Alkoxides as Precursors for the Preparation of Porous Fe- and Mn-TiO2 Photocatalysts with High Efficiencies. Chemistry 2008, 14, 11123–11131. [Google Scholar] [CrossRef] [PubMed]
  150. Kamble, R.; Sabale, S.; Chikode, P.; Puri, V.; Mahajan, S. Corrigendum: Structural and photocatalytic studies of hydrothermally synthesized Mn2+–TiO2 nanoparticles under UV and visible light irradiation (2016 Mater. Res. Express 3 115005). Mater. Res. Express 2018, 5, 129502. [Google Scholar] [CrossRef] [Green Version]
  151. Tbessi, I.; Benito, M.; Molins, E.; LIorca, J.; Touati, A.; Sayadi, S.; Najjar, W. Effect of Ce and Mn co-doping on photocatalytic performance of sol-gel TiO2. Solid State Sci. 2018, 88, 20–28. [Google Scholar] [CrossRef]
  152. Binas, V.; Stefanopoulos, V.; Kiriakidis, G.; Papagiannakopoulos, P. Photocatalytic oxidation of gaseous benzene, toluene and xylene under UV and visible irradiation over Mn-doped TiO2 nanoparticles. J. Mater. 2018, 5, 56–65. [Google Scholar] [CrossRef]
  153. Sudrajat, H.; Babel, S.; Ta, A.; Nguyen, T.K. Mn-doped TiO2 photocatalysts: Role, chemical identity, and local structure of dopant. J. Phys. Chem. Solids 2020, 144, 109517. [Google Scholar] [CrossRef]
  154. Wu, M.; Leung, D.; Zhang, Y.; Huang, H.; Xie, R.; Szeto, W.; Li, F. Toluene degradation over Mn-TiO2/CeO2 composite catalyst under vacuum ultraviolet (VUV) irradiation. Chem. Eng. Sci. 2018, 195, 985–994. [Google Scholar] [CrossRef]
  155. Pirzada, B.; Mir, N.; Qutub, N.; Mehraj, O.; Sabir, S.; Muneer, M. Synthesis, characterization and optimization of photocatalytic activity of TiO2/ZrO2 nanocomposite heterostructures. Mater. Sci. Eng. B 2014, 193, 137–145. [Google Scholar] [CrossRef]
  156. Ranjbar, S.; Saberyan, K.; Parsayan, F. A new process for preparation of Zr doped TiO2 nanopowders using APCVS method. Mater. Chem. Phys. 2018, 214, 337–344. [Google Scholar] [CrossRef]
  157. Kambur, A.; Pozan, G.; Boz, I. Preparation, characterization and photocatalytic activity of TiO2–ZrO2 binary oxide nanoparticles. Appl. Catal. B Environ. 2012, 115–116, 149–158. [Google Scholar] [CrossRef]
  158. Mohammadi, M.R.; Fray, D. Synthesis and characterization of nanosized TiO2-ZrO2 binary system prepared by an aqueous sol-gel process: Physical and sensing properties. Sensors Actuators B Chem. 2011, 155, 568–576. [Google Scholar] [CrossRef]
  159. T V L, T.; Deivasigamani, P.; Maheswari, M.A. Synthesis of mesoporous worm-like ZrO2–TiO2 monoliths and their photocatalytic applications towards organic dye degradation. J. Photochem. Photobiol. A Chem. 2017, 344, 212–222. [Google Scholar] [CrossRef]
  160. Goswami, P.; Ganguli, J. Tuning the band gap of mesoporous Zr-doped TiO2 for effective degradation of pesticide quinalphos. Dalton Trans. 2013, 42, 14480–14490. [Google Scholar] [CrossRef] [PubMed]
  161. Loganathan, K.; Azhagapillai, P.; Palanichamy, M.; Arumugam, E.; Murugesan, V. Synthesis and characterization of Zr4+, La3+ and Ce3+ doped mesoporous TiO2: Evaluation of their photocatalytic activity. J. Hazard Mater. 2011, 186, 1183–1192. [Google Scholar] [CrossRef]
  162. Babaei, M.; Dehghanian, C.; Taheri, P.; Babaei, M. Effect of duty cycle and electrolyte additive on photocatalytic performance of TiO2-ZrO2 composite layers prepared on CP Ti by micro arc oxidation method. Surf. Coatings Technol. 2016, 307, 554–564. [Google Scholar] [CrossRef] [Green Version]
  163. Pantelides, S. The electronic structure of impurities and other point defects in semiconductors. Rev. Mod. Phys. 1978, 50, 797–858. [Google Scholar] [CrossRef]
  164. Hamukwaya, S.L.; Zhao, Z.; Wang, N.; Liu, H.; Umar, A.; Zhang, J.; Wu, T.; Guo, Z. Enhanced Photocatalytic Activity of B, N-Codoped TiO2 by a New Molten Nitrate Process. J. Nanosci. Nanotechnol. 2019, 19, 839–849. [Google Scholar] [CrossRef]
  165. Li, M.; Li, X.; JIANG, G.; He, G. Hierarchically macro-mesoporous ZrO2-TiO2 composites with Enhanced photocatalytic activity. Ceram. Int. 2015, 41, 5749–5757. [Google Scholar] [CrossRef]
  166. Huang, Q.; Ma, W.; Yan, X.; Chen, Y.; Zhu, S.; Shen, S. Photocatalytic decomposition of gaseous HCHO by ZrxTi1−xO2 catalysts under UV–vis light irradiation with an energy-saving lamp. J. Mol. Catal. A Chem. 2013, 366, 261–265. [Google Scholar] [CrossRef]
  167. Mbiri, A.; Taffa, D.; Gatebe, E.; Wark, M. Zirconium doped mesoporous TiO2 multilayer thin films: Influence of the zirconium content on the photodegradation of organic pollutants. Catal. Today 2019, 328, 71–78. [Google Scholar] [CrossRef]
  168. Huang, C.; Ding, Y.; Chen, Y.; Li, P.; Zhu, S.; Shen, S. Highly efficient Zr doped-TiO2/glass fiber photocatalyst and its performance in formaldehyde removal under visible light. J. Environ. Sci. 2017, 60, 61–69. [Google Scholar] [CrossRef]
  169. Mahy, J.; Lambert, S.; Tilkin, R.; Wolfs, C.; Poelman, D.; Devred, F.; Gaigneaux, E.; Douven, S. Ambient temperature ZrO2-doped TiO2 crystalline photocatalysts: Highly efficient powders and films for water depollution. Mater. Today Energy 2019, 13, 312–322. [Google Scholar] [CrossRef]
  170. Belver, C.; Bedia, J.; Rodriguez, J.J. Zr-doped TiO2 supported on delaminated clay materials for solar photocatalytic treatment of emerging pollutants. J. Hazard Mater. 2016, 322, 233–242. [Google Scholar] [CrossRef]
  171. Zhao, J.; Ge, S.; Pan, D.; Shao, Q.; Lin, J.; Wang, Z.; Hu, Z.; Wu, T.; Guo, Z. Solvothermal Synthesis, Characterization and Photocatalytic Property of Zirconium Dioxide Doped Titanium Dioxide Spinous Hollow Microspheres with Sunflower Pollen as Bio-templates. J. Colloid Interface Sci. 2018, 529, 111–121. [Google Scholar] [CrossRef]
  172. Prabhakarrao, N.; Ravichandra, D.; Rao, T. Synthesis of Zr doped TiO2/reduced Graphene Oxide (rGO) nanocomposite material for efficient photocatalytic degradation of dyes under visible light irradiation. J. Alloys Compd. 2016, 694, 596–606. [Google Scholar] [CrossRef]
  173. Yasin, S.A.; Obaid, M.; El-Newehy, M.; Al-Deyab, S.; Barakat, N. Influence of TixZr(1−x)O2 nanofibers composition on the photocatalytic activity toward organic pollutants degradation and water splitting. Ceram. Int. 2015, 41, 11876–11885. [Google Scholar] [CrossRef]
  174. Ali, T.; Tripathi, P.; Azam, A.; Raza, W.; Shareef Ahmed, A.; Ahmed, A.; Muneer, M. Photocatalytic performance of Fe-doped TiO2 nanoparticles under visible-light irradiation. Mater. Res. Express 2017, 4, 15022. [Google Scholar] [CrossRef]
  175. Khan, H.; Swati, I. Fe3+-doped Anatase TiO2 with d-d Transition, Oxygen Vacancies and Ti3+ Centers: Synthesis, Characterization, UV-vis Photocatalytic and Mechanistic Studies. Ind. Eng. Chem. Res. 2016, 55, 6619–6633. [Google Scholar] [CrossRef]
  176. Sayyar, Z.; Babaluo, A.; Rahbar Shahrouzi, J. Kinetic Study of Formic Acid Degradation by Fe3+ doped TiO2 Self-Cleaning Nanostructure Surfaces Prepared by Cold Spray. Appl. Surf. Sci. 2015, 335, 1–10. [Google Scholar] [CrossRef]
  177. Naghibi, S.; Gharagozlou, M. Solvothermal Synthesis of M-doped TiO2 Nanoparticles for Sonocatalysis of Methylene Blue and Methyl Orange (M = Cd, Ag, Fe, Ce, and Cu): Sonocatalytic Performance of M-doped TiO2 NPs. J. Chin. Chem. Soc. 2017, 64, 640–650. [Google Scholar] [CrossRef]
  178. Li, J.; Ren, D.; Wu, Z.; Xu, J.; Bao, Y.; He, S.; Chen, Y. Flame retardant and visible light-activated Fe-doped TiO2 thin films anchored to wood surfaces for the photocatalytic degradation of gaseous formaldehyde. J. Colloid Interface Sci. 2018, 530, 78–87. [Google Scholar] [CrossRef] [PubMed]
  179. Craciun, E.; Predoana, L.; Atkinson, I.; Jitaru, I.; Anghel, E.M.; Bratan, V.; Gifu, C.; Anastasescu, C.; Rusu, A.; Raditoiu, V.; et al. Fe3+-doped TiO2 nanopowders for photocatalytic mineralization of oxalic acid under solar light irradiation. J. Photochem. Photobiol. A Chem. 2018, 356, 18–28. [Google Scholar] [CrossRef]
  180. Liu, M.; Qiu, X.; Miyauchi, M.; Hashimoto, K. Energy-Level Matching of Fe(III) Ions Grafted at Surface and Doped in Bulk for Efficient Visible-Light Photocatalysts. J. Am. Chem. Soc. 2013, 135, 10064–10072. [Google Scholar] [CrossRef]
  181. Han, F.; Kambala, V.; Srinivasan, M.; Dharmarajan, R.; Naidu, R. Tailored Titanium Dioxide Photocatalysts for the Degradation of Organic Dyes in Wastewater Treatment: A Review. Appl. Catal. A General 2009, 359, 25–40. [Google Scholar] [CrossRef]
  182. Jaihindh, D.; Verma, A.; Chen, C.-C.; Huang, Y.-c.; Dong, C.; Fu, Y. Study of oxidation states of Fe- and Co-doped TiO2 photocatalytic energy materials and their visible-light-driven photocatalytic behavior. Int. J. Hydrog. Energy 2018, 44, 15892–15906. [Google Scholar] [CrossRef]
  183. 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]
  184. Mancuso, A.; Sacco, O.; Sannino, D.; Pragliola, S.; Vaiano, V. Enhanced visible-light-driven photodegradation of Acid Orange 7 azo dye in aqueous solution using Fe-N co-doped TiO2. Arab. J. Chem. 2020, 13, 8347–8360. [Google Scholar] [CrossRef]
  185. Moradi, V.; Ahmed, F.; Jun, M.; Blackburn, A.; Herring, R. Acid-treated Fe-doped TiO2 as a high performance photocatalyst used for degradation of phenol under visible light irradiation. J. Environ. Sci. 2019, 83, 183–194. [Google Scholar] [CrossRef] [PubMed]
  186. Rodríguez, P.; Pecchi, G.; Casuscelli, S.; Elías, V.; Eimer, G. A simple synthesis way to obtain iron-doped TiO2 Nanoparticles as photocatalytic surfaces. Chem. Phys. Lett. 2019, 732, 136643. [Google Scholar] [CrossRef]
  187. Sood, S.; Umar, A.; Mehta, S.; Kansal, S. Highly effective Fe-doped TiO2 nanoparticles photocatalysts for visible-light driven photocatalytic degradation of toxic organic compounds. J. Colloid Interface Sci. 2015, 450, 213–223. [Google Scholar] [CrossRef] [PubMed]
  188. Komaraiah, D.; Radha, E.; Sivakumar, J.; Reddy, M.V.R.; Sayanna, R. Influence of Fe3+ ion doping on the luminescence emission behavior and photocatalytic activity of Fe3+, Eu3+-codoped TiO2 thin films. J. Alloys Compd. 2021, 868, 159109. [Google Scholar] [CrossRef]
  189. Kanakaraju, D.; Ya, M.; Jasni, M.; Endra, M.; Lim, Y.C. Fe Doped Titania Photocatalyst for Degradation of Methyl Orange. Mater. Today Proc. 2019, 19, 1657–1662. [Google Scholar] [CrossRef]
  190. Crişan, M.; Mardare, D.; Ianculescu, A.; Drăgan, N.; Niţoi, I.; Crişan, D.; Voicescu, M.; Todan, L.; Oancea, P.; Adomniţei, 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]
  191. Ahadi, S.; Moalej, N.; 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]
  192. Li, G.; Yi, L.; Wang, J.; Song, Y. Hydrodynamic cavitation degradation of Rhodamine B assisted by Fe3+-doped TiO2: Mechanisms, geometric and operation parameters. Ultrason. Sonochem. 2019, 60, 104806. [Google Scholar] [CrossRef] [PubMed]
  193. Velázquez-Martínez, S.; Silva Martínez, S.; Arellano, C.; González, A.; Salgado, I.; Morales-Pérez, A.A.; Peña-Cruz, M. Modified sol-gel/hydrothermal method for the synthesis of microsized TiO2 and iron-doped TiO2, its characterization and solar photocatalytic activity for an azo dye degradation. J. Photochem. Photobiol. A Chem. 2018, 359, 93–101. [Google Scholar] [CrossRef]
  194. Indah Anwar, D.; Mulyadi, D. Synthesis of Fe-TiO2 Composite as a Photocatalyst for Degradation of Methylene Blue. Procedia Chem. 2015, 17, 49–54. [Google Scholar] [CrossRef] [Green Version]
  195. Ambati, R.; Gogate, P. Photocatalytic degradation of Acid Blue 80 using iron doped TiO2 catalyst: Understanding the effect of operating parameters and combinations for synergism. J. Water Process Eng. 2017, 20, 217–225. [Google Scholar] [CrossRef]
  196. Tabasideh, S.; Maleki, A.; Shahmoradi, B.; Ghahremani, E.; Mckay, G. Sonophotocatalytic degradation of diazinon in aqueous solution using iron-doped TiO2 nanoparticles. Sep. Purif. Technol. 2017, 189, 186–192. [Google Scholar] [CrossRef]
  197. Khan, M.; Siwach, R.; Kumar, S.; Alhazaa, A. 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]
  198. Han, F.; Kambala, V.; Dharmarajan, R.; Liu, Y.; Naidu, R. Photocatalytic degradation of azo dye acid orange 7 using different light sources over Fe3+ -doped TiO2 nanocatalysts. Environ. Technol. Innov. 2018, 12, 27–42. [Google Scholar] [CrossRef]
  199. Zafar, Z.; Kim, J.-O. Optimization of hydrothermal synthesis of Fe-TiO2 nanotube arrays for enhancement in visible light using an experimental design methodology. Environ. Res. 2020, 189, 109908. [Google Scholar] [CrossRef]
  200. Bensouici, F.; Mostafa, B.; Tala-Ighil, R.; Bououdina, M.; Tounane, M. Surface, structural and optical properties dependence of Fe-doped TiO2 films deposited onto soda–lime–glass. Surf. Interfaces 2020, 21, 100682. [Google Scholar] [CrossRef]
  201. Komaraiah, D.; Radha, E.; Sivakumar, J.; Reddy, M.V.; 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]
  202. Moradi, V.; Jun, M.; Blackburn, A.; Herring, R. Significant improvement in visible light photocatalytic activity of Fe doped TiO2 using an acid treatment process. Appl. Surf. Sci. 2017, 427, 791–799. [Google Scholar] [CrossRef]
  203. Kakavandi, B.; Jorfi, S.; Fattahi, M.; Isari, A. Photocatalytic degradation of rhodamine B and real textile wastewater using Fe-doped TiO2 anchored on reduced graphene oxide (Fe-TiO2/rGO): Characterization and feasibility, mechanism and pathway studies. Appl. Surf. Sci. 2018, 462, 549–564. [Google Scholar]
  204. Crisan, M.; Răileanu, M.; Drăgan, N.; Crisan, D.; Ianculescu, A.; Nitoi, I.; Oancea, P.; Somacescu, S.; Stanica, N.; Vasile, B.; et al. Sol-gel iron-doped TiO2 nanopowders with Photocatalytic Activity. Appl. Catal. A Gen. 2014, 504, 130–142. [Google Scholar] [CrossRef]
  205. Adyani, S.M.; Ghorbani, M. A comparative study of physicochemical and photocatalytic properties of visible light responsive Fe, Gd and P single and tri-doped TiO2 nanomaterials. J. Rare Earths 2017, 36, 72–85. [Google Scholar] [CrossRef]
  206. Kaur, T.; Sraw, A.; Toor, A.; Wanchoo, R. Utilization of solar energy for the degradation of carbendazim and propiconazole by Fe doped TiO2. Sol. Energy 2016, 125, 65–76. [Google Scholar] [CrossRef]
  207. 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]
  208. Lian, J.-Z.; Tsai, C.-T.; Chang, S.-H.; Lin, N.-H.; Hsieh, Y.-H. Iron waste as an effective depend on TiO2 for photocatalytic degradation of dye waste water. Opt. Int. J. Light Electron Opt. 2017, 140, 197–204. [Google Scholar] [CrossRef]
  209. Lin, H.-J.; Yang, T.-S.; Hsi, C.-S.; Wang, M.-C.; Lee, K.-C. Optical and photocatalytic properties of Fe3+-doped TiO2 thin films prepared by a sol–gel spin coating. Ceram. Int. 2014, 40, 10633–10640. [Google Scholar] [CrossRef]
  210. May-Lozano, M.; Mendoza-Escamilla, V.; Rojas Garcia, E.; López-Medina, R.; Rivadeneyra-Romero, G.; Delgadillo, S. Sonophotocatalytic degradation of Orange II dye using low cost photocatalyst. J. Clean. Prod. 2017, 148, 836–844. [Google Scholar] [CrossRef]
  211. Chen, C.-C.; Hu, S.-H.; Fu, Y.-P. Effects of surface hydroxyl group density on the photocatalytic activity of Fe3+-doped TiO2. J. Alloys Compd. 2015, 632, 326–334. [Google Scholar] [CrossRef]
  212. Lin, H.-J.; Yang, T.-S.; Moo-Chin, W.; Hsi, C.-S. Structural and photodegradation behaviors of Fe3+-doping TiO2 thin films prepared by a sol–gel spin coating. J. Alloys Compd. 2014, 610, 478–485. [Google Scholar] [CrossRef]
  213. Nitoi, I.; Oancea, P.; Raileanu, M.; Crisan, M.; Constantin, L.A.; Cristea, I. UV–VIS photocatalytic degradation of nitrobenzene from water using heavy metal doped titania. J. Ind. Eng. Chem. 2015, 21, 677–682. [Google Scholar] [CrossRef]
  214. Xiang, G.; Yu, Z.; Hou, Y.; Chen, Y.; Peng, Z.; Sun, L.; Sun, L. Simulated solar-light induced photoelectrocatalytic degradation of bisphenol-A using Fe3+-doped TiO2 nanotube arrays as a photoanode with simultaneous aeration. Sep. Purif. Technol. 2016, 161, 144–151. [Google Scholar] [CrossRef]
  215. Soo, C.W.; Juan, J.C.; Wei Lai, C.; Abd Hamid, S.B.; Yusop, R. Fe-doped mesoporous anatase-brookite titania in the solar-light-induced photodegradation of Reactive Black 5 dye. J. Taiwan Inst. Chem. Eng. 2016, 68, 153–161. [Google Scholar] [CrossRef]
  216. Maki, L.; Maleki, A.; Rezaee, R.; Daraei, H. LED-activated immobilized Fe-Ce-N tri-doped TiO2 nanocatalyst on glass bed for photocatalytic degradation organic dye from aqueous solutions. Environ. Technol. Innov. 2019, 15, 100411. [Google Scholar] [CrossRef]
  217. Hosseini, M.; Ebratkhahan, M.; Shayegan, Z.; Niaei, A.; Salari, D.; Rostami, A.; Raeisipour, J. Investigation of the effective operational parameters of self-cleaning glass surface coating to improve methylene blue removal efficiency; application in solar cells. Sol. Energy 2020, 207, 398–408. [Google Scholar] [CrossRef]
  218. Choudhury, B.; Choudhury, A. Dopant induced changes in structural and optical properties of Cr3+ doped TiO2 nanoparticles. Mater. Chem. Phys. 2012, 132, 1112–1118. [Google Scholar] [CrossRef]
  219. Wang, Z.-M.; Yang, G.; Biswas, P.; Bresser, W.; Boolchand, P. Processing of Iron-Doped Titania Powders in Flame Aerosol Reactors. Powder Technol. 2001, 114, 197–204. [Google Scholar] [CrossRef]
  220. Peng, Y.-H.; Huang, G.; Huang, W.-Q. Visible-light absorption and photocatalytic activity of Cr-doped TiO2 nanocrystal films. Adv. Powder Technol. 2012, 23, 8–12. [Google Scholar] [CrossRef]
  221. Mardare, D.; Iacomi, F.; Cornei, N.; Girtan, M.; Luca, D. Undoped and Cr-doped TiO2 thin films obtained by spray pyrolysis. Thin Solid Films 2010, 518, 4586–4589. [Google Scholar] [CrossRef] [Green Version]
  222. Pei wen, K.; Mohd Hatta, M.H.; Ong, S.; Yuliati, L.; Lee, S. Photocatalytic degradation of photosensitizing and non-photosensitizing dyes over chromium doped titania photocatalysts under visible light. J. Photochem. Photobiol. A Chem. 2017, 332, 215–223. [Google Scholar] [CrossRef]
  223. Yamazaki, S.; Fujiwara, Y.; Yabuno, S.; Adachi, K.; Honda, K. Preparation of porous metal-ion-doped titanium dioxide and the photocatalytic degradation of 4-chlorophenol under visible light irradiation. Appl. Catal. B Environ. 2012, 121–122, 148–153. [Google Scholar] [CrossRef]
  224. Inturi, S.; Suidan, M.; Smirniotis, P. Influence of synthesis method on leaching of the Cr-TiO2 catalyst for visible light liquid phase photocatalysis and their stability. Appl. Catal. B Environ. 2016, 180, 351–361. [Google Scholar] [CrossRef]
  225. Park, S.; Joo, H.; Kang, J.-W. Effect of impurities in TiO2 thin films on trichloroethylene conversion. Sol. Energy Mater. Sol. Cells 2004, 83, 39–53. [Google Scholar] [CrossRef]
  226. Khan, H.; Jiang, Z.; Berk, D. Molybdenum doped graphene/TiO2 hybrid photocatalyst for UV/visible photocatalytic applications. Sol. Energy 2018, 162, 420–430. [Google Scholar] [CrossRef]
  227. Du, Y.; Gan, Y.; Yang, P.; Zhao, F.; Hua, N.; Jiang, L. Improvement in the Heat-Induced Hydrophilicity of TiO2 Thin Films by Doping Mo(VI) Ions. Thin Solid Films 2005, 491, 133–136. [Google Scholar] [CrossRef]
  228. Stengl, V.; Bakardjieva, S. Molybdenum-Doped Anatase and Its Extraordinary Photocatalytic Activity in the Degradation of Orange II in the UV and Vis Regions. J. Phys. Chem. C 2010, 114, 19308–19317. [Google Scholar] [CrossRef]
  229. Shahmoradi, B.; Ibrahim, I.; Sakamoto, N.; Ananda, S.; Row, T.; Soga, K.; Byrappa, K.; Parsons, S.; Shimizu, Y. In situ surface modification of molybdenum-doped organic-inorganic hybrid TiO2 nanoparticles under hydrothermal conditions and treatment of pharmaceutical effluent. Environ. Technol. 2010, 31, 1213–1220. [Google Scholar] [CrossRef]
  230. Cui, M.; Pan, S.; Tang, Z.; Chen, X.; Qiao, X.; Xu, Q. Physiochemical properties of n-n heterostructured TiO2/Mo-TiO2 composites and their photocatalytic degradation of gaseous toluene. Chem. Speciat. Bioavailab. 2017, 29, 60–69. [Google Scholar] [CrossRef] [Green Version]
  231. Zhan, C.; Chen, F.; Dai, H.; Yang, J.; Zhong, M. Photocatalytic activity of sulfated Mo-doped TiO2@fumed SiO2 composite: A mesoporous structure for methyl orange degradation. Chem. Eng. J. 2013, 225, 695–703. [Google Scholar] [CrossRef]
  232. Moghadam, R.; Vankova, S.; Monteverde Videla, A.; Specchia, S. Innovative carbon-free low content Pt catalyst supported on Mo-doped titanium suboxide (Ti3O5-Mo) for stable and durable oxygen reduction reaction. Appl. Catal. B Environ. 2016, 201, 419. [Google Scholar] [CrossRef]
  233. Avilés, O.; Espino-Valencia, J.; Romero, R.; Rico-Cerda, J.; Arroyo-Albiter, M.; Natividad, R. W and Mo doped TiO2: Synthesis, characterization and photocatalytic activity. Fuel 2017, 198, 31–41. [Google Scholar] [CrossRef]
  234. Sun, P.; Lu, Q.; Zhang, J.; Xiao, T.; Liu, W.; Ma, J.; Yin, S.; Cao, W. Mo-ion doping evoked visible light response in TiO2 nanocrystals for highly-efficient removal of benzene. Chem. Eng. J. 2020, 397, 125444. [Google Scholar] [CrossRef]
  235. Yang, H.; Li, X.; Wang, A.; Wang, Y.; Chen, Y. Photocatalytic degradation of methylene blue by MoO3 modified TiO2 under visible light. Chinese J. Catal. 2014, 35, 140–147. [Google Scholar] [CrossRef]
  236. Liu, S.; Zhao, X.; Wang, Y.; Shao, H.; Qiao, M.; Wang, Y.; Zhao, S. Peroxymonosulfate enhanced photoelectrocatalytic degradation of phenol activated by Co3O4 loaded carbon fiber cathode. J. Catal. 2017, 355, 167–175. [Google Scholar] [CrossRef]
  237. Yue, X.; Jiang, S.; Ni, L.; Wang, R.; Qiu, S.; Zhang, Z. The highly efficient photocatalysts of Co/TiO2: Photogenerated charge-transfer properties and their applications in photocatalysis. Chem. Phys. Lett. 2014, 615, 111–116. [Google Scholar] [CrossRef]
  238. Cao, C.; Hu, C.; Shen, W.; Wang, S.; Wang, J.; Tian, Y. Fabrication of a novel heterostructure of Co3O4-modified TiO2 nanorod arrays and its enhanced photoelectrochemical property. J. Alloys Compd. 2013, 550, 137–143. [Google Scholar] [CrossRef]
  239. Naseem, S.; Pinchuk, I.; Luo, Y.K.; Kawakami, R.; Khan, S.; Husain, S.; Khan, W. Epitaxial growth of cobalt doped TiO2 thin films on LaAlO3(100) substrate by Molecular Beam Epitaxy and their opto-magnetic based applications. Appl. Surf. Sci. 2019, 493, 691–702. [Google Scholar] [CrossRef]
  240. Elmragui, A.; Logvina, Y.; Pinto da Silva, L.; Zegaoui, O.; Silva, J. Synthesis of Fe- and Co-Doped TiO2 with Improved Photocatalytic Activity Under Visible Irradiation Toward Carbamazepine Degradation. Materials 2019, 12, 3874. [Google Scholar] [CrossRef] [Green Version]
  241. Mi, C.; Han, E.; Sun, L.; Zhu, L. Effect of Ti4+ doping on LiNi0.35Co0.27Mn0.35Fe0.03O2. Solid State Ionics 2019, 340, 114976. [Google Scholar] [CrossRef]
  242. Ebrahimian, A.; Monazzam, P.; Fakhari Kisomi, B. Co/TiO2 nanoparticles: Preparation, characterization and its application for photocatalytic degradation of methylene blue. Desalin. Wtare Treat. 2017, 63, 283–292. [Google Scholar] [CrossRef] [Green Version]
  243. Iwasaki, M.; Hara, M.; Kawada, H.; Tada, H.; Ito, S. Cobalt Ion-Doped TiO2 Photocatalyst Response to Visible Light. J. Colloid Interface Sci. 2000, 224, 202–204. [Google Scholar] [CrossRef]
  244. Hamadanian, M.; Reisi-Vanani, A.; Majedi, A. Sol-gel preparation and characterization of Co/TiO2 nanoparticles: Application to the degradation of methyl orange. Chem. Soc 2010, 7, 52–58. [Google Scholar] [CrossRef]
  245. Shifu, C.; Wei, L.; Sujuan, Z.; Yinghao, C. Preparation and activity evaluation of relative p-n junction photocatalyst Co-TiO2/TiO2. J. Sol-Gel Sci. Technol. 2010, 54, 258–267. [Google Scholar] [CrossRef]
  246. Amadelli, R.; Samiolo, L.; Andrea, M.; Molinari, A.; Mario, V.; Gazzoli, D. Preparation, Characterisation, and Photocatalytic Behaviour of Co-TiO2 with Visible Light Response. Int. J. Photoenergy 2008, 2008, 853753. [Google Scholar] [CrossRef] [Green Version]
  247. Nesfchi, M.; Ebrahimian, A.; Saraei, F.; Rojaee, F.; Mahdavi, F.; Rastegar, S. Fabrication of plasmonic nanoparticles/cobalt doped TiO2 nanosheets for degradation of tetracycline and modeling the process by artificial intelligence techniques. Mater. Sci. Semicond. Process. 2021, 122, 105465. [Google Scholar] [CrossRef]
  248. Elmragui, A.; Zegaoui, O.; Silva, J. Elucidation of the photocatalytic degradation mechanism of an azo dye under visible light in the presence of cobalt doped TiO2 nanomaterials. Chemosphere 2020, 266, 128931. [Google Scholar] [CrossRef] [PubMed]
  249. Pirbazari, A. Sensitization of TiO2 Nanoparticles With Cobalt Phthalocyanine: An Active Photocatalyst for Degradation of 4-Chlorophenol under Visible Light. Procedia Mater. Sci. 2015, 11, 622–627. [Google Scholar] [CrossRef] [Green Version]
  250. Alyani, S.; Ebrahimian, A.; Khalilsaraei, F.; Asasian, N.; Gilani, N. Growing Co-doped TiO2 nanosheets on reduced graphene oxide for efficient photocatalytic removal of tetracycline antibiotic from aqueous solution and modeling the process by artificial neural network. J. Alloys Compd. 2019, 799, 169–182. [Google Scholar] [CrossRef]
  251. Miao, Y.; Zhai, Z.; Jiang, L.; Shi, Y.; Yan, Z.; Duan, D.; Zhen, K.; Wang, J. Facile and new synthesis of cobalt doped mesoporous TiO2 with high visible-light performance. Powder Technol. 2014, 266, 365–371. [Google Scholar] [CrossRef]
  252. Hosaini, N.; Kamran-Pirzaman, A.; Aroon, M.; Ebrahimian, A. Photocatalytic degradation of 2,4-dichlorophenol by Co-doped TiO2(Co/TiO2) nanoparticles and Co/TiO2 containing mixed matrix membranes. J. Water Process Eng. 2017, 2017, 124–134. [Google Scholar] [CrossRef]
  253. Crisan, M.; Drăgan, N.; Crisan, D.; Ianculescu, A.; Nitoi, I.; Oancea, P.; Todan, L.; Stan, C.; Stanica, N. The effects of Fe, Co and Ni dopants on TiO2 structure of sol-gel nanopowders used as photocatalysts for environmental protection: A comparative study. Ceram. Int. 2015, 42, 3088–3095. [Google Scholar] [CrossRef]
  254. Preethi, T.; Abarna, B.; Vidhya, K.N.; Rajarajeswari, G.R. Sol-gel derived cobalt doped nano-titania photocatalytic system for solar light induced degradation of crystal violet. Ceram. Int. 2014, 40, 13159. [Google Scholar] [CrossRef]
  255. Lee, B.-K. Rapid photo-degradation of 2-chlorophenol under visible light irradiation using cobalt oxide-loaded TiO2/reduced graphene oxide nanocomposite from aqueous media. J. Environ. Manag. 2016, 165, 1–10. [Google Scholar] [CrossRef]
  256. Drăgan, N.; Crisan, M.; Răileanu, M.; Crisan, D.; Ianculescu, A.; Oancea, P.; Somacescu, S.; Todan, L.; Stanica, N.; Vasile, B. The effect of Co dopant on TiO2 structure of sol–gel nanopowders used as photocatalysts. Ceram. Int. 2014, 40, 12273–12284. [Google Scholar] [CrossRef]
  257. Sevim, A. Synthesis and characterization of Zn and Co monocarboxy-phthalocyanines and investigation of their photocatalytic efficiency as TiO2 composites. J. Organomet. Chem. 2017, 832, 18–26. [Google Scholar] [CrossRef]
  258. Dong, F.; Guo, S.; Wang, H.; Li, X.; Wu, Z. Enhancement of the Visible Light Photocatalytic Activity of C-Doped TiO2 Nanomaterials Prepared by a Green Synthetic Approach. J. Phys. Chem. C 2011, 115, 13285–13292. [Google Scholar] [CrossRef]
  259. Ravishankar, T.N.; Vaz, M.d.O.; Ramakrishnappa, T.; Teixeira, S.R.; Dupont, J. Ionic liquid–assisted hydrothermal synthesis of Nb/TiO2 nanocomposites for efficient photocatalytic hydrogen production and photodecolorization of Rhodamine B under UVvisible and visible light illuminations. Mater. Today Chem. 2019, 12, 373–385. [Google Scholar] [CrossRef]
  260. Furubayashi, Y.; Hitosugi, T.; Yamamoto, Y.; Inaba, K.; Kinoda, G.; Hirose, Y.; Shimada, T.; Hasegawa, T. A transparent metal: Nb-doped anatase TiO2. Appl. Phys. Lett. 2005, 86, 252101. [Google Scholar] [CrossRef]
  261. De Trizio, L.; Buonsanti, R.; Schimpf, A.; Llordés, A.; Gamelin, D.; Simonutti, R.; Milliron, D. Nb-Doped Colloidal TiO2 Nanocrystals with Tunable Infrared Absorption. Chem. Mater. 2013, 25, 3383–3390. [Google Scholar] [CrossRef]
  262. Su, H.; Huang, Y.-T.; Chang, Y.-H.; Zhai, P.; Hau, N.; Cheung, P.; Yeh, W.-T.; Wei, T.; Feng, S.-P. The Synthesis of Nb-doped TiO2 Nanoparticles for Improved-Performance Dye Sensitized Solar Cells. Electrochim. Acta 2015, 182, 230–237. [Google Scholar] [CrossRef]
  263. Khan, S.; Cho, H.; Han, S.S.; Lee, K.; Cho, S.-H.; Song, T.; Choi, H. Defect engineering toward strong photocatalysis of Nb-doped anatase TiO2: Computational predictions and experimental verifications. Appl. Catal. B Environ. 2017, 206, 520–530. [Google Scholar] [CrossRef]
  264. Lim, J.; Murugan, P.; Lakshminarasimhan, N.; Kim, J.; Lee, J.S.; Lee, S.-H.; Choi, W. Synergic photocatalytic effects of nitrogen and niobium co-doping in TiO2 for the redox conversion of aquatic pollutants under visible light. J. Catal. 2014, 310, 91–99. [Google Scholar] [CrossRef]
  265. Lim, J.; Monllor-Satoca, D.; Jang, J.; Lee, S.; Choi, W. Visible light photocatalysis of fullerol-complexed TiO2 enhanced by Nb doping. Appl. Catal. B Environ. 2014, 152–153, 233–240. [Google Scholar] [CrossRef]
  266. Ferraz, N.; Marcos, F.; Nogueira, A.; Martins, A.; Lanza, M.; Assaf, E.; Asencios, Y. Hexagonal-Nb2O5/Anatase-TiO2 mixtures and their applications in the removal of Methylene Blue Dye under various conditions. Mater. Chem. Phys. 2017, 198, 331–340. [Google Scholar] [CrossRef]
  267. Kou, Y.; Yang, J.; Li, B.; Fu, S. Solar photocatalytic activities of porous Nb-doped TiO2 microspheres by coupling with tungsten oxide. Mater. Res. Bull. 2015, 63, 105–111. [Google Scholar] [CrossRef]
  268. Silva, A.; Muche, D.; Dey, S.; Hotza, D.; Castro, R. Photocatalytic Nb2O5-doped TiO2 nanoparticles for glazed ceramic tiles. Ceram. Int. 2015, 42, 5113–5122. [Google Scholar] [CrossRef] [Green Version]
  269. Lee, D.; Park, J.-H.; Kim, Y.-H.; Lee, M.-H.; Cho, N.-I. Effect of Nb doping on morphology, crystal structure, optical band gap energy of TiO2 thin films. Curr. Appl. Phys. 2014, 14, 421–427. [Google Scholar] [CrossRef]
  270. Ullah, I.; Haider, A.; Khalid, N.; Ali, S.; Ahmed, S.; Khan, Y.; Ahmed, N.; Zubair, M. Tuning the band gap of TiO2 by tungsten doping for efficient UV and visible photodegradation of Congo red dye. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2018, 204, 150–157. [Google Scholar] [CrossRef] [PubMed]
  271. Yadav, M.; Yadav, A.; Fernandes, R.; Popat, Y.; Orlandi, M.; Dashora, A.; Kothari, D.C.; Miotello, A.; Ahuja, B.; Patel, N. Tungsten-doped TiO2/reduced Graphene Oxide nano-composite photocatalyst for degradation of phenol: A system to reduce surface and bulk electron-hole recombination. J. Environ. Manag. 2017, 203, 364–374. [Google Scholar] [CrossRef] [PubMed]
  272. Grabowska, E.; Nadolna, J.; Zaleska-Medynska, A. Mechanism of phenol photodegradation in the presence of pure and modified-TiO2: A review. Water Res. 2012, 46, 5453–5471. [Google Scholar] [CrossRef]
  273. Thomas, M.; Natarajan, T. TiO2-High Surface Area Materials Based Composite Photocatalytic Nanomaterials for Degradation of Pollutants: A Review. In Photocatalytic Nanomaterials for Environmental Applications; MRF: Millersville, PA, USA, 2018; ISBN 978-1-945291-58-6. [Google Scholar]
  274. Palmisano, L.; Sclafani, A.; Venezia, A.M.; Campostrini, R.; Carturan, G.; Martin, C.; Rives, V.; Solana, G. Influence of tungsten oxide on structural and surface properties of sol-gel prepared TiO2 employed for 4-nitrophenol photodegradation. J. Chem. Soc. Faraday Trans. 1996, 92, 819–829. [Google Scholar] [CrossRef]
  275. Keller, V.; Bernhardt, P.; Garin, F. Photocatalytic oxidation of butyl acetate in vapor phase on TiO2, Pt/TiO2 and WO3/TiO2 catalysts. J. Catal. 2003, 15, 129–138. [Google Scholar] [CrossRef]
  276. Y, T.; Song, K.; Lee, W.; Choi, G.; Y, R. Photocatalytic Behavior of WO3-Loaded TiO2 in an Oxidation Reaction. J. Catal. 2000, 191, 192–199. [Google Scholar] [CrossRef]
  277. Du, S.; Liao, Z.; Qin, Z.; Zuo, F.; Li, X. Polydopamine microparticles as redox mediators for catalytic reduction of methylene blue and rhodamine B. Catal. Commun. 2015, 72, 86–90. [Google Scholar] [CrossRef]
  278. Li, J.; Luo, D.; Yang, C.; He, S.; Chen, S.; Lin, J.; Zhu, L.; Li, X. Copper(II) imidazolate frameworks as highly efficient photocatalysts for reduction of CO2 into methanol under visible light irradiation. J. Solid State Chem. 2013, 203, 154–159. [Google Scholar] [CrossRef]
  279. Cai, H.; Chen, X.; Li, Q.; He, B.; Tang, Q. Enhanced photocatalytic activity from Gd, La codoped TiO2 nanotube array photocatalysts under visible-light irradiation. Appl. Surf. Sci. 2013, 284, 837–842. [Google Scholar] [CrossRef]
  280. Ki, S.J.; Park, Y.-K.; Kim, J.-S.; Lee, W.-J.; Lee, H.; Jung, S.-C. Facile Preparation of Tungsten Oxide Doped TiO2 Photocatalysts using Liquid Phase Plasma Process for Enhanced Degradation of Diethyl Phthalate. Chem. Eng. J. 2018, 377, 120087. [Google Scholar] [CrossRef]
  281. Azadi, S.; Karimi-Jashni, A.; Javadpour, S. Modeling and Optimization of Photocatalytic Treatment of Landfill Leachate using Tungsten-doped TiO2 Nano-photocatalysts: Application of Artificial Neural Network and Genetic Algorithm. Process Saf. Environ. Prot. 2018, 117, 267–277. [Google Scholar] [CrossRef]
  282. Kondamareddy, K.; D, N.; Lu, D.; Peng, T.; Macias, M.; Wang, C.; Yu, Z.; Cheng, N.; Fu, D.; Zhao, X.-Z. Ultra-trace (parts per million-ppm) W6+ dopant ions induced anatase to rutile transition (ART) of phase pure anatase TiO2 nanoparticles for highly efficient visible light-active photocatalytic degradation of organic pollutants. Appl. Surf. Sci. 2018, 456, 676–693. [Google Scholar] [CrossRef]
  283. Mayoufi, A.; Nsib, M.; Houas, A. Doping level effect on visible-light irradiation W-doped TiO2–anatase photocatalysts for Congo red photodegradation. Comptes Rendus Chim. 2014, 17, 818–823. [Google Scholar] [CrossRef]
  284. Wang, X.; Sun, M.; Murugananthan, M.; Zhang, Y.; Zhang, L. Electrochemically Self-Doped WO3/TiO2 Nanotubes for Photocatalytic Degradation of Volatile Organic Compounds. Appl. Catal. B Environ. 2019, 260, 118205. [Google Scholar] [CrossRef]
  285. Xu, H.; Liao, J.; Yuan, S.; Zhao, Y.; Zhang, M.; Wang, Z.; Shi, L. Tuning the morphology, stability and photocatalytic activity of TiO2 nanocrystal colloids by tungsten doping. Mater. Res. Bull. 2014, 51, 326–331. [Google Scholar] [CrossRef]
  286. Wang, Q.; Zhang, W.; Hu, X.; Xu, L.; Chen, G.; Li, X. Hollow spherical WO3/TiO2 heterojunction for enhancing photocatalytic performance in visible-light. J. Water Process Eng. 2021, 40, 101943. [Google Scholar] [CrossRef]
  287. Fouad, K.; Gar Alalm, M.; Bassyouni, M.; Saleh, M. A novel photocatalytic reactor for the extended reuse of W–TiO2 in the degradation of sulfamethazine. Chemosphere 2020, 257, 127270. [Google Scholar] [CrossRef] [PubMed]
  288. Abraham, C.; Devi, L.G. One-pot facile sol-gel synthesis of W, N, C and S doped TiO2 and its application in the photocatalytic degradation of thymol under the solar light irradiation: Reaction kinetics and degradation mechanism. J. Phys. Chem. Solids 2020, 141, 109350. [Google Scholar] [CrossRef]
  289. Hunge, Y.; Mahadik, M.; Moholkar, A.; Bhosale, C. Photoelectrocatalytic degradation of oxalic acid using WO3 and stratified WO3/TiO2 photocatalysts under sunlight illumination. Ultrason. Sonochem. 2016, 35, 233–242. [Google Scholar] [CrossRef]
  290. Wattanawikkam, C.; Pecharapa, W. Structural studies and photocatalytic properties of Mn and Zn co-doping on TiO2 prepared by single step sonochemical method. Radiat. Phys. Chem. 2020, 171, 108714. [Google Scholar] [CrossRef]
  291. El-Mehasseb, I.; Kandil, S.; Elgindy, K. Advanced visible-light applications utilizing modified Zn-doped TiO2 nanoparticles via non-metal in situ dual doping for wastewater detoxification. Optik 2020, 213, 164654. [Google Scholar] [CrossRef]
  292. Sanchez-Dominguez, M.; Morales-Mendoza, G.; Rodriguez-Vargas, M.; Ibarra-Malo, C.; Rodríguez, A.; Vela-Gonzalez, A.; Perez, A.; Gomez, R. Synthesis of Zn-doped TiO2 nanoparticles by the novel oil-in-water (O/W) microemulsion method and their use for the photocatalytic degradation of phenol. J. Environ. Chem. Eng. 2015, 3, 3037–3047. [Google Scholar] [CrossRef]
  293. Michaelson, H.B. The work function of the elements and its periodicity. J. AppI. Phys. 2021, 48, 4729–4733. [Google Scholar] [CrossRef] [Green Version]
  294. Linic, S.; Christopher, P.; Ingram, D. Plasmonic-metal nanostructures for efficient conversion of solar to chemical energy. Nat. Mater. 2011, 10, 911–921. [Google Scholar] [CrossRef]
  295. Kochuveedu, S.; Jang, Y.H.; Kim, D.H. A Study on the Mechanism for the Interaction of Light with Noble Metal-Metal Oxide Semiconductor Nanostructures for Various Photophysical Applications. Chem. Soc. Rev. 2013, 42, 8467–8493. [Google Scholar] [CrossRef]
  296. Ali, F.; Khan, S.; Kamal, T.; Alamry, K.; Asiri, A.M. Chitosan-titanium oxide fibers supported zero-valent nanoparticles: Highly efficient and easily retrievable catalyst for the removal of organic pollutants. Sci. Rep. 2018, 8, 6260. [Google Scholar] [CrossRef] [Green Version]
  297. Naya, S.-I.; Niwa, T.; Kume, T.; Tada, H. Visible-Light-Induced Electron Transport from Small to Large Nanoparticles in Bimodal Gold Nanoparticle-Loaded Titanium (IV) Oxide. Angew. Chem. Int. Ed. Engl. 2014, 53, 7305–7309. [Google Scholar] [CrossRef] [PubMed]
  298. Ahmad Rather, R.; Sharma, P.; Kumar, P.; Singh, S.; Pal, B. Plasmonic Stimulated Photocatalytic/Electrochemical Hydrogen Evolution from Water by (001) Faceted and bimetallic Loaded Titania Nanosheets under Sunlight Irradiation. J. Clean. Prod. 2017, 175, 394–401. [Google Scholar] [CrossRef]
  299. Ghosh, S.; Mallik, A.; Basu, R. Enhanced photocatalytic activity and photoresponse of poly(3,4-ethylenedioxythiophene) nanofibers decorated with gold nanoparticle under visible light. Sol. Energy 2017, 159, 548–560. [Google Scholar] [CrossRef]
  300. Wang, M.; Ye, M.; Iocozzia, J.; Lin, C.; Lin, Z. Plasmon-Mediated Solar Energy Conversion via Photocatalysis in Noble Metal/Semiconductor Composites. Adv. Sci. 2016, 3, 1600024. [Google Scholar] [CrossRef] [Green Version]
  301. Li, G.; Li, J.; Li, G.; Jiang, G. N and Ti3+ co-doped 3D anatase TiO2 superstructures composed of ultrathin nanosheets with enhanced visible light photocatalytic activity. J. Mater. Chem. A 2015, 3, 22073–22080. [Google Scholar] [CrossRef]
  302. Cheng, L.; Zhang, D.; Liao, Y.; Li, F.; Zhang, H.; Xiang, Q. Constructing functionalized plasmonic Au/TiO2 nanosheets with small Au nanoparticles for efficient photocatalytic hydrogen evolution. J. Colloid Interface Sci. 2019, 555, 94–103. [Google Scholar] [CrossRef]
  303. Naya, S.; Tada, H. Dependence of the plasmonic activity of Au/TiO2 for the decomposition of 2-naphthol on the crystal form of TiO2 and Au particle size. J. Catal. 2018, 364, 328–333. [Google Scholar] [CrossRef]
  304. Shoueir, K.; Kandil, S.; El-Hosainy, H. Tailoring the surface reactivity of plasmonic Au@TiO2 photocatalyst bio-based chitosan fiber towards cleaner of harmful water pollutants under visible-light irradiation. J. Clean. Prod. 2019, 230, 383–393. [Google Scholar] [CrossRef]
  305. Cojocaru, B.; Andrei, V.; Tudorache, M.; Lin, F.; Cadigan, C.; Richards, R.; Parvulescu, V. Enhanced photo-degradation of bisphenol pollutants onto gold-modified photocatalysts. Catal. Today 2016, 284, 153–159. [Google Scholar] [CrossRef]
  306. Hazra Chowdhury, I.; Roy, M.; Kundu, S.; Naskar, M. TiO2 hollow microspheres impregnated with biogenic gold nanoparticles for the efficient visible light-induced photodegradation of phenol. J. Phys. Chem. Solids 2019, 129, 329–339. [Google Scholar] [CrossRef]
  307. Celebi, N.; Aydın, M.Y.; Soysal, F.; Ciftci, Y.; Salimi, K. Ligand-free fabrication of Au/TiO2 nanostructures for plasmonic hot-electron-driven photocatalysis: Photoelectrochemical water splitting and organic-dye degredation. J. Alloys Compd. 2021, 860, 157908. [Google Scholar] [CrossRef]
  308. Wang, Y.; Yang, C.; Chen, A.; Pu, W.; Gong, J. Influence of yolk-shell Au@TiO2 structure induced photocatalytic activity towards gaseous pollutant degradation under visible light. Appl. Catal. B Environ. 2019, 251, 57–65. [Google Scholar] [CrossRef]
  309. Nasirian, M.; Mehrvar, M. Modification of TiO2 to Enhance Photocatalytic Degradation of Organics in Aqueous Solutions. J. Environ. Chem. Eng. 2016, 4, 4072–4082. [Google Scholar] [CrossRef]
  310. Al-Hajji, L.; Ismail, A.; Bumajdad, A.; al Saidi, M.; Ahmed, S.A.; Almutawa, F.; Alhazza, A. Construction of Au/TiO2 Heterojunction with high photocatalytic performances under UVA illumination. Ceram. Int. 2020, 46, 20155–20162. [Google Scholar] [CrossRef]
  311. Tiwari, A.; Shukla, A.; Lalliansanga; Tiwari, D.; Lee, S. Au-nanoparticle/Nanopillars TiO2 meso-porous thin films in the degradation of tetracycline using UV-A light. J. Ind. Eng. Chem. 2018, 69, 141–152. [Google Scholar] [CrossRef]
  312. Singh, J.; Sahu, K.; Satpati, B.; Shah, J.; Kotnala, R.K.; Mohapatra, S. Facile synthesis, structural and optical properties of Au-TiO2 plasmonic nanohybrids for photocatalytic applications. J. Phys. Chem. Solids 2019, 135, 109100. [Google Scholar] [CrossRef]
  313. Nalenthiran, P.; Murugesan, S.; Sathishkumar, P.; Sambandam, A. Photocatalytic degradation of ceftiofur sodium in the presence of gold nanoparticles loaded TiO2 under UV–visible light. Chem. Eng. J. 2014, 241, 401–409. [Google Scholar] [CrossRef]
  314. Hernandez, R.; Olvera-Rodríguez, I.; Guzmán, C.; Medel, A.; Escobar-Alarcón, L.; Brillas, E.; Sirés, I.; Escalante, K. Microwave-assisted sol-gel synthesis of an Au-TiO2 photoanode for the advanced oxidation of paracetamol as model pharmaceutical pollutant. Electrochem. Commun. 2018, 96, 42–46. [Google Scholar] [CrossRef]
  315. Khaywah, M.; Jradi, S.; Louarn, G.; Lacroute, Y.; Toufaily, J.; Hamieh, T.; Adam, P.-M. Ultrastable, Uniform, Reproducible, and Highly Sensitive Bimetallic Nanoparticles as Reliable Large Scale SERS Substrates. J. Phys. Chem. 2015, 119, 26091–26100. [Google Scholar] [CrossRef]
  316. Kallel, W.; Chaabene, S.; Bouattour, S. Novel (Ag,Y) doped TiO2 plasmonic photocatalyst with enhanced photocatalytic activity under visible light. Physicochem. Probl. Miner. Process. 2019, 55, 745–759. [Google Scholar] [CrossRef]
  317. Katchala, N.; Reddy, S.K.; Rao, T.; Anandan, S. Energy level matching for efficient charge transfer in Ag doped–Ag modified TiO2 for enhanced visible light photocatalytic activity. J. Alloys Compd. 2019, 794. [Google Scholar] [CrossRef]
  318. Ali, T.; Ahmed, A.; Alam, U.; Uddin, I.; Tripathi, P.; Muneer, M. Enhanced photocatalytic and antibacterial activities of Ag-doped TiO2 nanoparticles under visible light. Mater. Chem. Phys. 2018, 212, 325–335. [Google Scholar] [CrossRef]
  319. Peng, C.; Wang, W.; Zhang, W.; Liang, Y.; Zhuo, L. Surface plasmon-driven photoelectrochemical water splitting of TiO2 nanowires decorated with Ag nanoparticles under visible light illumination. Appl. Surf. Sci. 2017, 420, 286–295. [Google Scholar] [CrossRef]
  320. Bhavani, K.; Naresh, G.; Srinivas, B.; Venugopal, A. Plasmonic resonance nature of Ag-Cu/TiO2 photocatalyst under solar and artificial light: Synthesis, characterization and evaluation of H2O splitting activity. Appl. Catal. B Environ. 2016, 199, 282–291. [Google Scholar] [CrossRef]
  321. Ryu, S.-Y.; Chung, J.W.; Kwak, S.-Y. Dependence of photocatalytic and antimicrobial activity of electrospun polymeric nanofiber composites on the positioning of Ag-TiO2 nanoparticles. Compos. Sci. Technol. 2015, 117, 9–17. [Google Scholar] [CrossRef]
  322. Du, M.; Xiong, S.; Wu, T.; Zhao, D.; Zhang, Q.; Fan, Z.; Zeng, Y.; Ji, F.; He, Q.; Xu, X. Preparation of a Microspherical Silver-Reduced Graphene Oxide-Bismuth Vanadate Composite and Evaluation of Its Photocatalytic Activity. Materials 2016, 9, 160. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  323. Rahul, T.K.; Sandhyarani, N. Plasmonic and Photonic Effects on Hydrogen Evolution over Chemically Modified Titania Inverse Opals. ChemNanoMat 2018, 4, 642–648. [Google Scholar] [CrossRef]
  324. Low, J.; Qiu, S.; Xu, D.; Jiang, C.; Cheng, B. Direct evidence and enhancement of surface plasmon resonance effect on Ag-loaded TiO2 nanotube arrays for photocatalytic CO2 reduction. Appl. Surf. Sci. 2017, 434, 423–432. [Google Scholar] [CrossRef]
  325. Perumal, S.; MonikandaPrabu, K.; Sambandam, C.G.; Mohamed, A.P. Synthesis and characterization studies of solvothermally synthesized undoped and Ag-doped TiO2 nanoparticles using toluene as a solvent. J. Eng. Res. Appl. 2014, 4, 184–187. [Google Scholar]
  326. Zhang, L.; Ni, C.; Jiu, H.; Xie, C.; Yan, J.; Qi, G. One-pot synthesis of Ag-TiO2/reduced graphene oxide nanocomposite for high performance of adsorption and photocatalysis. Ceram. Int. 2017, 43, 5450–5456. [Google Scholar] [CrossRef]
  327. Chen, P. A novel synthesis of Ti3+ self-doped Ag2O/TiO2 (p–n) nanoheterojunctions for enhanced visible photocatalytic activity. Mater. Lett. 2016, 163, 130–133. [Google Scholar] [CrossRef]
  328. Wang, T.; Tang, T.; Gao, Y.; Chen, Q.; Zhang, Z.; Bian, H. Hydrothermal preparation of Ag-TiO2-reduced graphene oxide ternary microspheres structure composite for enhancing photocatalytic activity. Phys. E Low-Dimens. Syst. Nanostruct. 2018, 112, 128–136. [Google Scholar] [CrossRef]
  329. Devi, L.; Kavitha, R. A review on plasmonic metal-TiO2 composite for generation, trapping, storing and dynamic vectorial transfer of photogenerated electrons across the Schottky junction in a Photocatalytic system. Appl. Surf. Sci. 2015, 360, 601–622. [Google Scholar] [CrossRef]
  330. Khalid, N.R.; Ahmed, E.; Ahmad, M.; Niaz, N.A.; Ramzan, M.; Shakil, M.; Iqbal, T.; Majid, A. Microwave-assisted synthesis of Ag-TiO2/graphene composite for hydrogen production under visible light irradiation. Ceram. Int. 2016, 42, 18257–18263. [Google Scholar] [CrossRef]
  331. Rong, X.; Qiu, F.; Zhang, C.; Fu, L.; Wang, Y.; Yang, D. Preparation of Ag–AgBr/TiO2–graphene and its visible light photocatalytic activity enhancement for the degradation of polyacrylamide. J. Alloys Compd. 2015, 639, 153–161. [Google Scholar] [CrossRef]
  332. Quan, X.; Tan, H.; Zhao, Q.; Sang, X. Preparation of lanthanum-doped TiO2 photocatalysts by coprecipitation. J. Mater. Sci. 2007, 42, 6287–6296. [Google Scholar] [CrossRef]
  333. Ashraf, M.; Liu, Z.; Peng, W.-X.; Jermsittiparsert, K.; Hosseinzadeh, G.; Hosseinzadeh, R. Combination of sonochemical and freeze-drying methods for synthesis of graphene/Ag-doped TiO2 nanocomposite: A strategy to boost the photocatalytic performance via well distribution of nanoparticles between graphene sheets. Ceram. Int. 2019, 46, 7446–7452. [Google Scholar] [CrossRef]
  334. Demirci, S.; Dikici, T.; Yurddaskal, M.; Gültekin, S.; Toparli, M.; Çelik, E. Synthesis and characterization of Ag doped TiO2 heterojunction films and their photocatalytic performances. Appl. Surf. Sci. 2016, 390, 591–601. [Google Scholar] [CrossRef]
  335. Malfatti, L.; Carboni, D.; Pinna, A.; Lasio, B.; Marmiroli, B.; Innocenzi, P. In situ growth of Ag nanoparticles in graphene–TiO2 mesoporous films induced by hard X-ray. J. Sol-Gel Sci. Technol. 2016, 79, 295–302. [Google Scholar] [CrossRef]
  336. Zhou, G.; Meng, H.; Cao, Y.; Kou, X.; Duan, S.; Fan, L.; Xiao, M.; Zhou, F.; Li, Z.; Xing, Z. Surface Plasmon Resonance-Enhanced Solar-Driven Photocatalytic Performance from Ag Nanoparticles-Decorated Ti3+ Self-Doped Porous Black TiO2 Pillars. J. Ind. Eng. Chem. 2018, 64, 188–193. [Google Scholar] [CrossRef]
  337. Liu, R.; Wang, P.; Wang, X.; Yu, H.; Yu, J. UV- and Visible-Light Photocatalytic Activity of Simultaneously Deposited and Doped Ag/Ag(I)-TiO2 Photocatalyst. J. Phys. Chem. C 2012, 116, 17721–17728. [Google Scholar] [CrossRef]
  338. Hou, X.; Liu, A.-D.; Huang, M.-D.; Liao, B.; Wu, X.-L. First-Principles Band Calculations on Electronic Structures of Ag-Doped Rutile and Anatase TiO2. Chin. Phys. Lett. 2009, 26, 077106. [Google Scholar] [CrossRef]
  339. Xiong, Z.; Ma, J.; Ng, J.; Waite, T.; Zhao, X.S. Silver-Modified Mesoporous TiO2 Photocatalyst for Water Purification. Water Res. 2011, 45, 2095–2103. [Google Scholar] [CrossRef]
  340. Onkani, S.P.; Diagboya, P.; Mtunzi, F.; Klink, M.J.; Olu-Owolabi, B.; Pakade, V. Comparative study of the photocatalytic degradation of 2–chlorophenol under UV irradiation using pristine and Ag-doped species of TiO2, ZnO and ZnS photocatalysts. J. Environ. Manag. 2020, 260, 110145. [Google Scholar] [CrossRef] [PubMed]
  341. Komaraiah, D.; Radha, E.; Sivakumar, J.; Reddy, M.V.; Sayanna, R. Photoluminescence and photocatalytic activity of spin coated Ag+ doped anatase TiO2 thin films. Opt. Mater. 2020, 108, 110401. [Google Scholar] [CrossRef]
  342. Kiran, T.; Ahmed, H.M.; Begum, N.; Manjunatha, K.G. Sun light driven photocatalytic performance of Ag decorated TiO2 nanocomposite thin films by sol gel method. Mater. Today Proc. 2021, 46, 5948–5952. [Google Scholar] [CrossRef]
  343. Abbad, S.; Guergouri, K.; Gazaout, S.; Djebabra, S.; Zertal, A.; Barille, R.; Zaabat, M. Effect of silver doping on the photocatalytic activity of TiO2 nanopowders synthesized by the sol-gel route. J. Environ. Chem. Eng. 2020, 8, 103718. [Google Scholar] [CrossRef]
  344. Ribao, P.; Corredor, J.; Rivero, M.; Ortiz, I. Role of reactive oxygen species on the activity of noble metal-doped TiO2 photocatalysts. J. Hazard. Mater. 2018, 372, 45–51. [Google Scholar] [CrossRef] [PubMed]
  345. Ling, L.; Yawei, F.; Li, H.; Chen, Y.; Wen, J.; Zhu, J.; Bian, Z. Microwave induced surface enhanced pollutant adsorption and photocatalytic degradation on Ag/TiO2. Appl. Surf. Sci. 2019, 483, 772–778. [Google Scholar] [CrossRef]
  346. Sui, M.; Dong, Y.; Wang, Z.; Wang, F.; You, H. A biocathode-driven photocatalytic fuel cell using an Ag-doped TiO2/Ti mesh photoanode for electricity generation and pollutant degradation. J. Photochem. Photobiol. A Chem. 2017, 348, 238–245. [Google Scholar] [CrossRef]
  347. Pazoki, M.; Parsa, M.; Farhadpour, R. Removal of the hormones dexamethasone (DXM) by Ag doped on TiO2 photocatalysis. J. Environ. Chem. Eng. 2016, 4, 4426–4434. [Google Scholar] [CrossRef]
  348. Scarisoreanu, M.; Ilie, G.; Goncearenco, E.; Banici, A.; Morjan, I.; Dutu, E.; Tanasă, E.; Fort, I.; Stan, M.; Mihailescu, C.; et al. Ag, Au and Pt decorated TiO2 biocompatible nanospheres for UV & Vis photocatalytic water treatment. Appl. Surf. Sci. 2019, 509, 145217. [Google Scholar] [CrossRef]
  349. Suwarnkar, M.; Dhabbe, R.; Kadam, A.; Garadkar, K. Enhanced photocatalytic activity of Ag doped TiO2 nanoparticles synthesized by a microwave assisted method. Ceram. Int. 2014, 40, 5489–5496. [Google Scholar] [CrossRef]
  350. Zhang, P.; Li, X.; Wu, X.; Zhao, T.; Wen, L. Influence of In3+-doping and Ag0-depositing on the visible-light-induced photocatalytic activity of TiO2. J. Alloys Compd. 2016, 673, 405–410. [Google Scholar] [CrossRef]
  351. R, S.; Manoj, D.; Qin, J.; Highly Cited Researcher, M.; Gracia, F.; Lee, A.; Khan, M.M.; Pinilla, M.Á. Mechanothermal synthesis of Ag/TiO2 for photocatalytic Methyl Orange degradation and hydrogen production. Process Saf. Environ. Prot. 2018, 120, 339–347. [Google Scholar] [CrossRef]
  352. Vaiano, V.; Iervolino, G.; Sannino, D.; Murcia, J.; Hidalgo, M.; Ciambelli, P.; Navio, J.A. Photocatalytic removal of Patent Blue V dye on Au-TiO2 and Pt-TiO2 catalysts. Appl. Catal. B Environ. 2016, 188, 134–146. [Google Scholar] [CrossRef]
  353. Masakazu, A.; Masato, T. The Design and Development of Highly Reactive Titanium Oxide Photocatalysts Operating under Visible Light Irradiation. J. Catal. 2003, 216, 505–516. [Google Scholar]
  354. Alamelu, K.; Ali, J. TiO2-Pt composite photocatalyst for photodegradation and chemical reduction of recalcitrant organic pollutants. J. Environ. Chem. Eng. 2018, 6, 5720–5731. [Google Scholar] [CrossRef]
  355. Xu, T.; Zhao, H.; Zheng, H.; Zhang, P. Atomically Pt implanted nanoporous TiO2 film for photocatalytic degradation of trace organic pollutants in water. Chem. Eng. J. 2019, 385, 123832. [Google Scholar] [CrossRef]
  356. Kitsiou, V.; Zachariadis, G.; Lambropoulou, D.; Tsiplakides, D.; Poulios, I. Mineralization of the antineoplastic drug carboplatin by heterogeneous photocatalysis with simultaneous synthesis of platinum-modified TiO2 catalysts. J. Environ. Chem. Eng. 2018, 6, 2409–2416. [Google Scholar] [CrossRef]
  357. Alahmadi, N.; Amin, M.S.; Mohamed, R.M. Facile synthesis of mesoporous Pt-doped, titania-silica nanocomposites as highly photoactive under visible light. J. Mater. Res. Technol. 2020, 9, 14093–14102. [Google Scholar] [CrossRef]
  358. Khan, H.; Rigamonti, M.; Boffito, D. Enhanced Photocatalytic Activity of Pt-TiO2/WO3 Hybrid Material with Energy Storage Ability. Appl. Catal. B Environ. 2019, 252, 77–85. [Google Scholar] [CrossRef]
  359. Pol, R.; Guerrero, M.; García-Lecina, E.; Altube, A.; Rossinyol, E.; Garroni, S.; Baró, M.D.; Pons, J.; Sort, J.; Pellicer, E. Ni-, Pt- and (Ni/Pt)-doped TiO2 nanophotocatalysts: A smart approach for sustainable degradation of Rhodamine B dye. Appl. Catal. B Environ. 2016, 181, 270–278. [Google Scholar] [CrossRef] [Green Version]
  360. Nguyen-Phan, T.-D.; Luo, S.; Vovchok, D.; Llorca, J.; Sallis, S.; Kattel, S.; Xu, W.; Piper, L.; Polyansky, D.; Senanayake, S.; et al. Three-Dimensional Ruthenium-Doped TiO2 Sea Urchins for Enhanced Visible-Light-Responsive H2 Production. Phys. Chem. Chem. Phys. 2016, 18, 15972–15979. [Google Scholar] [CrossRef]
  361. Ismael, M. High effective ruthenium-doped TiO2 nanoparticles photocatalyst for visible-light-driven photocatalytic hydrogen production. New J. Chem. 2019, 43, 9596–9605. [Google Scholar] [CrossRef]
  362. Wang, Z.; Liu, B.; Xie, Z.; Li, Y.-M.; Shen, Z.-Y. Preparation and Photocatalytic Properties of RuO2/TiO2 Composite Nanotube Arrays. Ceram. Int. 2016, 42, 13664–13669. [Google Scholar] [CrossRef]
  363. González, A.; Solis-Cortazar, J.; Arellano, C.; Ramirez Morales, E.; Espinosa de los Monteros, A.; Silva Martínez, S. Synthesis of Ruthenium-Doped TiO2 Nanotube Arrays for the Photocatalytic Degradation of Terasil Blue Dye. J. Nanosci. Nanotechnol. 2019, 19, 5211–5219. [Google Scholar] [CrossRef]
  364. García-Ramírez, P.; Ramirez Morales, E.; Cortazar, J.; Sirés, I.; Silva Martínez, S. Influence of ruthenium doping on UV- and visible-light photoelectrocatalytic color removal from dye solutions using a TiO2 nanotube array photoanode. Chemosphere 2020, 267, 128925. [Google Scholar] [CrossRef]
  365. Fatimah, I.; Nurillahi, R.; Sahroni, I.; Muraza, O. TiO2-pillared saponite and photosensitization using a ruthenium complex for photocatalytic enhancement of the photodegradation of bromophenol blue. Appl. Clay Sci. 2019, 183, 105302. [Google Scholar] [CrossRef]
  366. Elsalamony, R.; Mahmoud, S. Preparation of nanostructured ruthenium doped titania for the photocatalytic degradation of 2-chlorophenol under visible light. Arab. J. Chem. 2017, 10, 194–205. [Google Scholar] [CrossRef] [Green Version]
  367. Długokęcka, M.; Łuczak, J.; Polkowska, Ż.; Zaleska-Medynska, A. The effect of microemulsion composition on the morphology of Pd nanoparticles deposited at the surface of TiO2 and photoactivity of Pd-TiO2. Appl. Surf. Sci. 2017, 405, 220–230. [Google Scholar] [CrossRef]
  368. Nguyen, C.H.; Fu, C.-C.; Juang, R.-S. Degradation of methylene blue and methyl orange by palladium-doped TiO2 photocatalysis for water reuse: Efficiency and degradation pathways. J. Clean. Prod. 2018, 202, 413–427. [Google Scholar] [CrossRef]
  369. Bai, X.; Lv, L.; Zhang, X.; Hua, Z. Synthesis and Photocatalytic Properties of Palladium-loaded Three Dimensional Flower-like Anatase TiO2 with Dominant {001} Facets. J. Colloid Interface Sci. 2015, 467, 1–9. [Google Scholar] [CrossRef] [PubMed]
  370. Dadsetan, S.; Baghshahi, S.; Farshidfar, F.; Hadavi, M. Photodeposition of Pd nanoparticles on TiO2 Utilizing a Channel Type Quartz Reactor. Ceram. Int. 2017, 43, 9322–9326. [Google Scholar] [CrossRef]
  371. Fodor, S.; Kovács, G.; Hernadi, K.; Danciu, V.; Baia, L.; Pap, Z. Shape tailored Pd nanoparticles’ effect on the photocatalytic activity of commercial TiO2. Catal. Today 2017, 284, 137–145. [Google Scholar] [CrossRef] [Green Version]
  372. Niu, X.; Li, S.; Chu, H.; Zhou, J. Preparation, characterization of Y3+-doped TiO2 nanoparticles and their photocatalytic activities for methyl orange degradation. J. Rare Earths 2011, 29, 225–229. [Google Scholar] [CrossRef]
  373. Ahmad, A.L.; Otitoju, T.A.; Ooi, B.S. Optimization of a high performance 3-aminopropyltriethoxysilane-silica impregnated polyethersulfone membrane using response surface methodology for ultrafiltration of synthetic oil-water emulsion. J. Taiwan Inst. Chem. Eng. 2018, 93, 461–476. [Google Scholar] [CrossRef]
  374. Mazierski, P.; Lisowski, W.; Grzyb, T.; Winiarski, M.; Klimczuk, T.; Mikolajczyk, A.; Flisikowski, J.; Hirsch, A.; Kołakowska, A.; Puzyn, T.; et al. Enhanced photocatalytic properties of lanthanide-TiO2 nanotubes: An experimental and theoretical study. Appl. Catal. B Environ. 2017, 205, 376–385. [Google Scholar] [CrossRef]
  375. Nadolna, J.; Grzyb, T.; Wei, Z.; Klein, M.; Kowalska, E.; Ohtani, B.; Zaleska-Medynska, A. Photocatalytic activity and luminescence properties of RE3+–TiO2 nanocrystals prepared by sol–gel and hydrothermal methods. Appl. Catal. B Environ. 2016, 181, 825–837. [Google Scholar] [CrossRef]
  376. Li, F.; Li, X.; Hou, M. Photocatalytic Degradation of 2-Mercaptobenzothiazole in Aqueous La3+–TiO2 Suspension for Odor Control. Appl. Catal. B Environ. 2004, 48, 185–194. [Google Scholar] [CrossRef]
  377. Zhang, Y.; Zhang, H.; Xu, Y.; Wang, Y. Significant effect of lanthanide doping on the texture and properties of nanocrystalline mesoporous TiO2. J. Solid State Chem. 2004, 177, 3490–3498. [Google Scholar] [CrossRef]
  378. Ganesan, M.; Viswanathan, B.; Viswanath, R.P.; Varadarajan, T.K. Photocatalytic behavior of CeO2-TiO2 system for the degradation of methylene blue. Indian J. Chem. Sect. A Inorg. Phys. Theor. Anal. Chem. 2009, 48, 480–488. [Google Scholar]
  379. Jo, W.; Kim, J.-T. Application of visible-light photocatalysis with nitrogen-doped or unmodified titanium dioxide for control of indoor-level volatile organic compounds. J. Hazard Mater. 2008, 164, 360–366. [Google Scholar] [CrossRef]
  380. Li, F.; Li, X.; Hou, M.; Cheah, K.W.; Choy, W.C.H. Enhanced photocatalytic activity of Ce3+–TiO2 for 2-mercaptobenzothiazole degradation in aqueous suspension for odour control. Appl. Catal. A Gen. 2005, 285, 181–189. [Google Scholar] [CrossRef] [Green Version]
  381. Cai, W.; Chen, F.; Shen, X.; Chen, L.; Zhang, J. Enhanced catalytic degradation of AO7 in the CeO2–H2O2 system with Fe3+ doping. Appl. Catal. B Environ. 2010, 101, 160–168. [Google Scholar] [CrossRef]
  382. Vieira, G.; José, H.; Peterson, M.; Baldissarelli, V.; Alvarez, P.; Moreira, R. CeO2/TiO2 nanostructures enhance adsorption and photocatalytic degradation of organic compounds in aqueous suspension. J. Photochem. Photobiol. A Chem. 2017, 353, 325–336. [Google Scholar] [CrossRef]
  383. Zhang, L.; Xie, J.; Li, G.; Zhang, H. Effect of Ce4+-doping on structural and photocatalytic properties of sol-gel prepared titanium dioxide thin-films. Bandaoti Guangdian/Semicond. Optoelectron. 2013, 34, 98–102. [Google Scholar]
  384. Tong, T.; Zhang, J.; Tian, B.; Chen, F.; He, D.; Anpo, M. Preparation of Ce-TiO2 catalysts by controlled hydrolysis of titanium alkoxide based on esterification reaction and study on its photocatalytic activity. J. Colloid Interface Sci. 2007, 315, 382–388. [Google Scholar] [CrossRef] [PubMed]
  385. Xiao, J.; Peng, T.; Li, R.; Peng, Z.; Yan, C. Preparation, Phase Transformation and Photocatalytic Activities of Cerium-Doped Mesoporous Titania Nanoparticles. J. Solid State Chem. 2006, 179, 1161–1170. [Google Scholar] [CrossRef]
  386. Guangqin, L.; Liu, C.; Liu, Y. Different effects of cerium ions doping on properties of anatase and rutile TiO2. Appl. Surf. Sci. 2006, 253, 2481–2486. [Google Scholar] [CrossRef]
  387. Xie, J.; Jiang, D.; Chen, M.; Li, D.; Zhu, J.; Yan, C. Preparation and characterization of monodisperse Ce-doped TiO2 microspheres with visible light photocatalytic activity. Colloids Surfaces A Physicochem. Eng. Asp. 2010, 372, 107–114. [Google Scholar] [CrossRef]
  388. Aman, N.; Satapathy, P.; Mishra, T.; Mahato, M.; Das, N. Synthesis and photocatalytic activity of mesoporous cerium doped TiO2 as visible light sensitive photocatalyst. Mater. Res. Bull. 2011, 47, 179–183. [Google Scholar] [CrossRef]
  389. Singh, K.; Harish, S.; Kristy, P.; Vemulapalli, S.; Archana, J.; M, N.; Shimomura, M.; Hayakawa, Y. Erbium doped TiO2 interconnected mesoporous spheres as an efficient visible light catalyst for photocatalytic applications. Appl. Surf. Sci. 2018, 449, 755–763. [Google Scholar] [CrossRef]
  390. Manasa, M.; Chandewar, P.; Mahalingam, H. Photocatalytic degradation of ciprofloxacin & norfloxacin and disinfection studies under solar light using boron & cerium doped TiO2 catalysts synthesized by green EDTA-citrate method. Catal. Today 2021, 375, 522–536. [Google Scholar] [CrossRef]
  391. Shayegan, Z.; Haghighat, F.; Lee, C.-S. Surface Fluorinated Ce-doped TiO2 Nanostructure Photocatalyst: A Trap and Remove Strategy to Enhance the VOC Removal from Indoor Air Environment. Chem. Eng. J. 2020, 401, 125932. [Google Scholar] [CrossRef]
  392. Mureseanu, M.; Chivu, V.; Osiac, M.; Madalina, C.; Bucur, C.; Parvulescu, V.; Cioatera, N. New photoactive mesoporous Ce-modified TiO2 for simultaneous wastewater treatment and electric power generation. Catal. Today 2021, 366, 164–176. [Google Scholar] [CrossRef]
  393. Rapsomanikis, A.; Apostolopoulou, A.; Stathatos, E.; Lianos, P. Cerium-modified TiO2 nanocrystalline films for Visible light photocatalytic activity. J. Photochem. Photobiol. A Chem. 2014, 280, 46–53. [Google Scholar] [CrossRef]
  394. Chen, F.; Ho, P.-L.; Ran, R.; Chen, W.; Si, Z.; Wu, X.; Weng, D.; Huang, Z.; Lee, C. Synergistic effect of CeO2 modified TiO2 photocatalyst on the enhancement of visible light photocatalytic performance. J. Alloys Compd. 2017, 714, 560–566. [Google Scholar] [CrossRef]
  395. Alvarez-Láinez, M.; Cano Franco, J.C. Effect of CeO2 content in morphology and optoelectronic properties of TiO2-CeO2 nanoparticles in visible light organic degradation. Mater. Sci. Semicond. Process. 2019, 90, 190–197. [Google Scholar] [CrossRef]
  396. Liang, J.; Wang, J.; Yu, K.; Kexian, S.; Wang, X.; Liu, W.; Hou, J.; Liang, C. Enhanced Photocatalytic Performance of Nd3+-Doped TiO2 Nanosphere under Visible Light. Chem. Phys. 2019, 528, 110538. [Google Scholar] [CrossRef]
  397. Nithya, N.; Bhoopathi, G.; Magesh, G.; Kumar, C. Neodymium doped TiO2 nanoparticles by sol-gel method for antibacterial and photocatalytic activity. Mater. Sci. Semicond. Process. 2018, 83, 70–82. [Google Scholar] [CrossRef]
  398. Nadolna, J.; Grzyb, T.; Sobczak, J.; Lisowski, W.; Gazda, M.; Ohtani, B.; Zaleska-Medynska, A. Visible light activity of rare earth metal doped (Er3+, Yb3+ or Er3+/Yb3+) titania photocatalysts. Appl. Catal. B Environ. 2015, 163, 40–49. [Google Scholar] [CrossRef]
  399. Zheng, Y.; Wang, W. Electrospun nanofibers of Er3+-doped TiO2 with photocatalytic activity beyond the absorption edge. J. Solid State Chem. 2014, 210, 206–212. [Google Scholar] [CrossRef]
  400. Bhethanabotla, V.C.; Russell, D.R.; Kuhn, J.N. Assessment of mechanisms for enhanced performance of Yb/Er/titania photocatalysts for organic degradation: Role of rare earth elements in the titania phase. Appl. Catal. B Environ. 2017, 202, 156–164. [Google Scholar] [CrossRef]
  401. Farbod, M.; Kajbafvala, M. Surface modification of TiO2 nanoparticles by magnetic ions: Synthesis and application in enhancement of photocatalytic performance. Appl. Catal. B Environ. 2017, 219, 344–352. [Google Scholar] [CrossRef]
  402. Parnicka, P.; Mazierski, P.; Lisowski, W.; Klimczuk, T.; Nadolna, J.; Zaleska-Medynska, A. A new simple approach to prepare rare-earth metals-modified TiO2 nanotube arrays photoactive under visible light: Surface properties and mechanism investigation. Results Phys. 2019, 12, 412–423. [Google Scholar] [CrossRef]
  403. Toloman, D.; Popa, A.; Stefan, M.; Pana, O.; Silipas, D.; Sergiu, M.; Barbu-Tudoran, L. Impact of Gd ions from the lattice of TiO2 nanoparticles on the formation of reactive oxygen species during the degradation of RhB under visible light irradiation. Mater. Sci. Semicond. Process. 2017, 71, 61–68. [Google Scholar] [CrossRef]
  404. Dal’Toé, A.; Lopes Colpani, G.; Padoin, N.; Fiori, M.; Soares, C. Lanthanum doped titania decorated with silver plasmonic nanoparticles with enhanced photocatalytic activity under UV-visible light. Appl. Surf. Sci. 2018, 441, 1057–1071. [Google Scholar] [CrossRef]
  405. Zhang, J.; Wu, W.; Yan, S.; Chu, G.; Zhao, S.; Wang, X.; Li, C. Enhanced photocatalytic activity for the degradation of rhodamine B by TiO2 modified with Gd2O3 calcined at high temperature. Appl. Surf. Sci. 2015, 344, 249–256. [Google Scholar] [CrossRef]
  406. Nair, R.R.; Arulraj, J.; Devi, K.S. Ceria doped titania nano particles: Synthesis and photocatalytic activity. Mater. Today Proc. 2016, 3, 1643–1649. [Google Scholar] [CrossRef]
  407. Jung, S.-C.; Bang, H.-J.; Lee, H.; Ha, H.-H.; Yu, Y.H.; Kim, S.-J.; Park, Y.-K. Assessing the photocatalytic activity of europium doped TiO2 using liquid phase plasma process on acetylsalicylic acid. Catal. Today 2020. In Press. [Google Scholar] [CrossRef]
  408. Hu, Z.; Xu, T.; Liu, P.; Oeser, M. Microstructures and optical performances of nitrogen-vanadium co-doped TiO2 with enhanced purification efficiency to vehicle exhaust. Environ. Res. 2021, 193, 110560. [Google Scholar] [CrossRef] [PubMed]
  409. Teh, C.; Mohamed, A. Roles of titanium dioxide and ion-doped titanium dioxide on photocatalytic degradation of organic pollutants (phenolic compounds and dyes) in aqueous solutions: A review. J. Alloy. Compd. 2011, 509, 1648–1660. [Google Scholar] [CrossRef]
  410. Skorb, E.; Antonouskaya, L.I.; Belyasova, N.; Shchukin, D.; Moehwald, H.; Sviridov, D. Antibacterial Activity of Thin-film Photocatalysts Based on Metal-modified TiO2 and TiO2:In2O3 Nanocomposite. Appl. Catal. B Environ. 2008, 84, 94–99. [Google Scholar] [CrossRef]
  411. Li, C.; Ming, T.; Wang, J.; Wang, J.; Yu, J.; Yu, S. Ultrasonic aerosol spray-assisted preparation of TiO2/In2O3 composite for visible-light-driven photocatalysis. J. Catal. 2014, 310, 84–90. [Google Scholar] [CrossRef]
  412. Tahir, M.; Saidina Amin, N.A. Photocatalytic CO2 reduction and kinetic study over In/TiO2 nanoparticles supported microchannel monolith photoreactor. Appl. Catal. A Gen. 2013, 467, 483–496. [Google Scholar] [CrossRef]
  413. Glez, V.; Moreno-Rodríguez, A.; May, M.; Tzompantzi, F.J.; Gomez, R. Slurry photodegradation of 2,4-dichlorophenoxyacetic acid: A comparative study of impregnated and sol–gel In2O3–TiO2 mixed oxide catalysts. J. Photochem. Photobiol. A Chem. 2008, 193, 266–270. [Google Scholar] [CrossRef]
  414. Yu, Y.; Wang, E.; Yuan, J.; Cao, Y. Enhanced photocatalytic activity of titania with unique surface indium and boron species. Appl. Surf. Sci. 2013, 273, 638–644. [Google Scholar] [CrossRef]
  415. Zhou, J.; Zhang, Y.; Zhao, X.S.; Ray, A. Photodegradation of Benzoic Acid over Metal-Doped TiO2. Ind. Eng. Chem. Res. 2006, 45, 3503–3511. [Google Scholar] [CrossRef]
  416. Okajima, T.; Yamamoto, T.; Kunisu, M.; Yoshioka, S.; Tanaka, I.; Umesaki, N. Dilute Ga Dopant in TiO2 by X-ray Absorption Near-Edge Structure. Jpn. J. Appl. Phys. 2006, 45, 7028. [Google Scholar] [CrossRef]
  417. Umare, S.; Charanpahari, A.; Sasikala, R. Enhanced visible light photocatalytic activity of Ga, N and S codoped TiO2 for degradation of azo dye. Mater. Chem. Phys. 2013, 140, 529–534. [Google Scholar] [CrossRef]
  418. Mohamed, H.; Qarni, F.; Al-Omair, N. Design of porous Ga doped TiO2 nanostructure for enhanced solar light photocatalytic applications. Mater. Res. Bull. 2021, 133, 111057. [Google Scholar] [CrossRef]
  419. Liu, X.; Khan, M.; Liu, W.; Xiang, W.; Guan, M.; Jiang, P.; Cao, W. Synthesis of nanocrystalline Ga–TiO2 powders by mild hydrothermal method and their visible light photoactivity. Ceram. Int. 2015, 41, 3075–3080. [Google Scholar] [CrossRef]
  420. Deng, Q.; Han, X.; Gao, Y. Remarkable optical red shift and extremely high optical absorption coefficient of V-Ga co-doped TiO2. J. Appl. Phys. 2012, 112. [Google Scholar] [CrossRef]
  421. Lee, D.-K.; Yoo, H.-I. Electrical conductivity and oxygen nonstoichiometry of acceptor (Ga)-doped titania. Phys. Chem. Chem. Phys. 2008, 10, 6890–6898. [Google Scholar] [CrossRef]
  422. Ahmad, A.; Senapati, S.; Khan, M.; Kumar, R.; Sastry, M. Extracellular Biosynthesis of Monodisperse Gold Nanoparticles by a Novel Extremophilic Actinomycete, Thermomonospora sp. Langmuir 2003, 19, 3550–3553. [Google Scholar] [CrossRef]
  423. Hopper, M.; Sauvage, F.; Chandiran, A.K.; Graetzel, M.; Poeppelmeier, K.; Mason, T. Electrical Properties of Nb-, Ga-, and Y-Substituted Nanocrystalline Anatase TiO2 Prepared by Hydrothermal Synthesis. J. Am. Ceram. Soc. 2012, 95, 3192–3196. [Google Scholar] [CrossRef]
  424. Song, S.; Wang, C.; Hong, F.; He, Z.; Cai, Q.; Chen, J. Gallium and iodine-co-doped titanium dioxide for photocatalytic degradation of 2-chlorophenol in aqueous solution: Role of gallium. Appl. Surf. Sci. 2011, 257, 3427–3432. [Google Scholar] [CrossRef]
  425. Chae, J.; Kim, D.; Kim, S.; Kang, M. Photovoltaic efficiency on dye-sensitized solar cells (DSSC) assembled using Ga-incorporated TiO2 materials. J. Ind. Eng. Chem. 2010, 16, 906–911. [Google Scholar] [CrossRef]
  426. Niyomkarn, S.; Puangpetch, T.; Chavadej, S. Mesoporous-assembled In2O3–TiO2 mixed oxide photocatalysts for efficient degradation of azo dye contaminant in aqueous solution. Mater. Sci. Semicond. Process. 2014, 25, 112–122. [Google Scholar] [CrossRef]
  427. An, G.; Mahadik, M.; Piao, G.; Chae, W.-S.; Park, H.; Cho, M.; Chung, H.-S.; Jang, J. Hierarchical TiO2 @In2O3 heteroarchitecture photoanodes: Mechanistic study on interfacial charge carrier dynamics through water splitting and organic decomposition. Appl. Surf. Sci. 2019, 480, 1–12. [Google Scholar] [CrossRef]
  428. Sangjan, S.; Wisasa, K.; Deddeaw, N. Enhanced photodegradation of reactive blue dye using Ga and Gd as catalyst in reduced graphene oxide-based TiO2 composites. Mater. Today Proc. 2019, 6, 19–23. [Google Scholar] [CrossRef]
  429. Ismail, A.; Abdelfattah, I.; Faisal, M.; Helal, A. Efficient Photodecomposition of Herbicide Imazapyr over Mesoporous Ga2O3-TiO2 Nanocomposites. J. Hazard Mater. 2017, 342, In. [Google Scholar] [CrossRef] [PubMed]
  430. Areerachakul, N.; Sakulkhaemaruethai, S.; Johir, M.A.H.; Kandasamy, J.; Vigneswaran, S. Photocatalytic degradation of organic pollutants from wastewater using aluminium doped titanium dioxide. J. Water Process Eng. 2019, 27, 177–184. [Google Scholar] [CrossRef]
  431. Sengele, A.; Robert, D.; Keller, N.; Colbeau-Justin, C.; Keller, V. Sn-doped and porogen-modified TiO2 photocatalyst for solar light elimination of sulfure diethyle as a model for chemical warfare agent. Appl. Catal. B Environ. 2018, 245, 279–289. [Google Scholar] [CrossRef]
  432. Li, J.; Shi, J.; Li, Y.; Ding, Z.; Huang, J. A biotemplate synthesized hierarchical Sn-doped TiO2 with superior photocatalytic capacity under simulated solar light. Ceram. Int. 2021, 47, 8218–8227. [Google Scholar] [CrossRef]
  433. Zhu, X.; Han, S.; Feng, W.; Kong, Q.; Dong, Z.; Wang, C.; Lei, J.; Yi, Q. The effect of heat treatment on the anatase–rutile phase transformation and photocatalytic activity of Sn-doped TiO2 nanomaterials. RSC Adv. 2018, 8, 14249–14257. [Google Scholar] [CrossRef] [Green Version]
  434. Wang, J.; Ren, H.; Chen, W.-F.; Koshy, P.; Sorrell, C. Fabrication and Photocatalytic Performance of Sn-Doped Titania Hollow Spheres Using Polystyrene as Template. Ceram. Int. 2017, 44, 4981–4989. [Google Scholar] [CrossRef]
  435. Mohammad, A.; Khan, M.; Cho, M.H.; Yoon, T. Fabrication of binary SnO2/TiO2 nanocomposites under a sonication-assisted approach: Tuning of band-gap and water depollution applications under visible light irradiation. Ceram. Int. 2021, 47, 15073–15081. [Google Scholar] [CrossRef]
  436. Albornoz, L.; Wohlmuth da Silva, S.; Bortolozzi, J.; Banús, E.; Brussino, P.; Ulla, M.; Bernardes, A. Degradation and mineralization of erythromycin by heterogeneous photocatalysis using SnO2-doped TiO2 structured catalysts: Activity and stability. Chemosphere 2021, 268, 128858. [Google Scholar] [CrossRef] [PubMed]
  437. Mugunthan, E.; Saidutta, M.B.; PonnanEttiyappan, J. Photocatalytic Degradation of Diclofenac using TiO2-SnO2 Mixed Oxide Catalysts. Environ. Technol. 2017, 40, 929–941. [Google Scholar] [CrossRef]
  438. Rangel-Vázquez, I.; Del Angel, G.; Bertin, V.; González, F.; Vázquez-Zavala, A.; Arrieta, A.; Padilla, J.M.; Barrera, A.; Ramírez, E. Synthesis and characterization of Sn doped TiO2 photocatalysts: Effect of Sn concentration on the textural properties and on the photocatalytic degradation of 2,4-dichlorophenoxyacetic acid. J. Alloys Compd. 2014, 643, S144–S149. [Google Scholar] [CrossRef]
Figure 1. Mechanism of photocatalytic activity of metal-ion modified TiO2 in the presence of visible light [38].
Figure 1. Mechanism of photocatalytic activity of metal-ion modified TiO2 in the presence of visible light [38].
Catalysts 11 01039 g001
Figure 2. Table of commonly used metallic ions to enhance the photocatalytic degradation of organic pollutants.
Figure 2. Table of commonly used metallic ions to enhance the photocatalytic degradation of organic pollutants.
Catalysts 11 01039 g002
Table 1. Traditional photodegradation methods and limitations.
Table 1. Traditional photodegradation methods and limitations.
ApproachLimitations
Biological treatmentChlorinated phenols are resistant to biodegradation and can accumulate in sediments. They transfer the contaminants from one medium to another and require disposal or further treatment
The biodegradation of organic pollutants such as 4-CP is slow and incomplete, and its byproducts are more toxic compared to other pollutants.
Biological processes usually require considerable processing time to decompose 4-CP.
By use of the biological treatment, it cannot be degraded due to a large number of aromatic structures in the dye molecules and the stability of modern dyes.
Azo bond is reduced to form a colorless but toxic and potentially carcinogenic aromatic amine.
Adsorption technology and activated carbon adsorption methodThe post-treatment of wastewater and regeneration of adsorbent materials requires an expensive operation.
Activated carbon adsorption requires the safe disposal of carbon.
During the adsorption process, the system cannot tolerate the suspended solids in the influent water as a result of blockage.
The operating cost is high due to the requirements of the carbon system.
The treatment may be problematic if the polluted carbon is not regenerated.
Chemical precipitationRequirements for a large amount of chemicals and a large amount of sludge produced.
Requires further treatment or disposal.
Due to the large amount of sludge needing to be treated, the method is not feasible economically.
Air strippingIt is susceptible to pollution.
Aesthetic limitations due to tower height.
Challenges involving mechanical reliability.
Membrane adsorptionThe purchasing cost of membranes and residues (very concentrated filtrate) is high and must be collected or may require further processing.
The physical method is not destructive, but only transfers pollutants to other media, causing secondary pollution.
Table 2. General properties and principal applications of the metals.
Table 2. General properties and principal applications of the metals.
Classes of MetalsFeatures/PropertiesPrincipal Applications
Transition metalsThey have incompletely filled d orbitals. The transition metals are more electronegative than the other metals and form stable compounds with neutral molecules (such as water or ammonia). The main advantages of these metals are malleability and ductility. Examples of transition metals used to modify TiO2 photocatalysts include vanadium (V), nickel (Ni), copper (Cu), manganese (Mn), zirconium (Zr), iron (Fe), chromium (Cr), molybdenum (Mo), cobalt (Co), niobium (Nb), tungsten (W), and zinc (Zn). The density of V, Ni, Cu, Mn, Zr, Fe, Cr, Mo, Co, Nb, W, and Zn are 6.0, 8.90, 8.96, 7.3, 6.52, 7.87, 7.15, 10.2, 8.86, 8.57, 19.3, and 7.134 g cm−3, respectively.Luminescence, electronic device, and water pipes applications.
Noble metalsThese metals are known as iron lovers due to their ability to dissolve in iron either as solid solutions or in the molten state. They have outstanding resistance to chemical attacks even at high temperatures. Furthermore, they are well known for their catalytic properties and associated capacity to facilitate or control the rates of chemical reactions. Examples of noble metals used to modify TiO2 photocatalysts include ruthenium (Ru), palladium (Pd), platinum (Pt), gold (Au), and silver (Ag). The density of Ru, Pd, Pt, Au, and Ag are 12.1, 12.0, 21.5, 19.3, and 10.5 g cm−3, respectively.Hydrogenation, total oxidation and, more recently, partial oxidation heterogeneous catalysts.
Rare-earth metalsRare earth metals possess good magnetic and luminescent properties. The common rare earth metals used to modify TiO2 photocatalysts include cerium (Ce), erbium (Er), holmium (Ho), gadolinium (Gd), terbium (Tb), neodymium (Nd), ytterbium (Yb), samarium (Sm), lanthanum (La), europium (Eu), and yttrium (Y). The density of Ce, Er, Ho, Gd, Tb, Nd, Yb, Sm, La, Eu, and Y is 6.77, 9.07, 8.80, 7.90, 8.23, 7.01, 6.90, 7.52, 6.15, 5.24 and 4.47 g cm−3, respectively.Cellphones, electrical and electronic components, lasers, glass, magnetic materials, fluorescent lights, defense, clean energy.
MetalsMetals are malleable, ductile, good conductors of heat and electricity. Metals have a luster with high tensile strength. Typical metals used to modify TiO2 photocatalysts include gallium (Ga), indium (In), aluminum (Al), and tin (Sn). The density of Ga, In, Al, and Sn is 5.91, 7.31, 2.70, and 7.287 g cm3, respectively.Automobiles, electronic devices.
Table 3. Summary of recent progress on V-TiO2 photocatalysts for organic pollutants degradation.
Table 3. Summary of recent progress on V-TiO2 photocatalysts for organic pollutants degradation.
Material (TiO2 Phase Transition)OptimumCatalyst DosageSynthetic MethodsPollutant (Volume)Light SourceUnmodifiedModifiedRef.
V-TNT nanosheet (mixed crystal of anatase and rutile phases)1.0% V-TNT0.1 gHydrothermal method20 mg/L RhB (100 mL)UV/Visible light70% within 70 min for TNTThe reaction rate was 3.26-fold and 9.27-fold as compared to unmodified TNT under visible light and UV–vis irradiation, respectively.[71]
V2O5/TiO2 nanoparticles (anatase phase)3 wt% V2O5/TiO20.05 gIncipient wet impregnation method0.2 mg/L Eosin Y (200 mL)Visible light8% within 180 min for V2O593% within 180 min, the stability only decreased by 20% for the first run[72]
V2O5/P25 nanoparticles (anatase phase)3 wt% V2O5/P250.05 gIncipient wet impregnation method0.2 mg/L Eosin Y (200 mL)Visible light8% within 180 min for V2O581% within 180 min[72]
V2O5/TiO2 nanoparticles (anatase phase)3 wt% V2O5/TiO20.05 gIncipient wet impregnation method0.2 mg/L Eosin Y (200 mL)Visible light100% within 180 min for TiO2 and P25, 33% within 180 min for V2O540% within 180 min[72]
V2O5/TiO2 coatings (anatase phase)2.0 wt% V2O5/TiO215 mm × 10 mmSol–gel method8 mg/L MO (10 mL)Sunlight35% within 8 h53% within 8 h[73]
V-TiO2 nanoparticles (rutile to stable anatase phase)1.0% V-TiO20.05 gSol–gel method10 mg/L RhB (100 mL)Xenon Lamp21.56% within 300 min53.74% within 300 min[74]
V-TiO2 nanopowders (anatase phase)50 wt% V2O5/TiO20.1 gSolid-state dispersion method25 mg/L 2,4-CP (50 mL)UV-B72% within 30 min for pure V2O5, 66% within 30 min for pure TiO285% within 30 min[75]
V-TiO2 coupons (rutile to stable anatase phase)6.0 wt% V-TiO215 mm × 10 mmPEOx10 mg/L MB (100 mL)Tungsten-halogen55% within 180 min for pure TiO285% within 180 min[76]
V-TiO2 nanoparticles (anatase to brookite and rutile
phases)
0.125 mol%-V-TiO20.1 g/LMWASG85 mg/L MB and SMX (100 mL)UV0.031 min−1 for MB within 60 min0.035 and 0.0262 min−1 for MB and SMX, respectively within 60 min[77]
MWASG: Microwave-assisted sol–gel method; PEOx: plasma electrolytic oxidation; RhB: rhodamine B; 2,4-CP: 2,4-dichlorophenol; MB: methylene blue; SMX: sulfamethazine.
Table 4. Summary of recent progress on Ni-TiO2 photocatalysts for organic pollutants degradation.
Table 4. Summary of recent progress on Ni-TiO2 photocatalysts for organic pollutants degradation.
Material (TiO2 Phase Transition)OptimumCatalyst DosageSynthetic MethodsPollutant (Volume)Light SourceUnmodifiedModifiedRef.
Ni-TiO2 nanotubes (mixed crystal of anatase and rutile phases)10% Ni-TNT0.1 gHydrothermal method5 mg/L MB (50 mL) Sunlight71.4% within 90 min for pure TiO2 ~65% within 90 min[83]
Ni-TiO2 nanoparticles (mixed crystal of anatase and rutile phases)0.9 wt% Ni- TiO21.5 g/LSol–gel method10 mg/L MB (100 mL)Hg vapor lamp62% within 120 min for pure TiO250% within 120 min[84]
Ni-TiO2 nanoparticles (anatase phases)0.01% Ni-TiO20.1 gHydrothermal method10 mg/L MB (100 mL)Tungsten-halogen lamp-75% within 300 min[85]
Ni-TiO2 inverse opal photonic microarray (IOPM) (anatase phases)3.0 wt% Ni-TiO2-Sol–gel method10 mg/L MB (100 mL)Sunlight60% within 90 min for TiO2 IOPM95% within 90 min (1.58 times larger compared to that over TiO2 IOPM)[86]
Ni-TiO2 nanoparticles (anatase phases)1.0 wt% Ni-TiO21 g/LMWASG10 mg/L BPA (200 mL)Phillip lamp60.1% (0.0098 min−1) within 120 min93% (0.0255 min−1) within 120 min[87]
Ni-TiO2 nanoparticles (anatase phases)0.50 wt% Ni-TiO22 g/LSol–gel method25 mg/L IBP (50 mL)Solar light76% within 6 h for pure TiO278% (0.0046 min−1) within 6 h[88]
Ni-TiO2 nanoparticles (anatase to rutile phase) 10 wt% Ni-TiO22 g/LCoprecipitation method100 mg/L ECR (200 mL)Visible light20.9% within 120 min for pure TiO237.4% within 120 min[89]
NiO/TiO2 nanopowders (mixed phase of rutile and anatase)0.5 wt% NiO/TiO20.05 gModified combustion-based method15 mg/L MB (50 mL)Daylight emission85% within 210 min90% within 210 min[90]
Ni-TiO2 nanopowders (anatase phase)0.05 mol% Ni1 g L−1Sol–gel method (non-aqueous)10 mg L−1 MB (500 mL)Xenon Lamp88.0% within 180 min for pure TiO298.9% within 180 min[91]
Ni-TiO2 nanoparticles (anatase phase)1.0% Ni-TiO20.025 gSol–gel method10−5 mol/L MO (50 mL)Visible light0.00075 and 0.002631 min−1 for visible and UV irradiation, respectively. 16.3 and 95.6% for visible and 95.6 for visible and UV light, irradiation within 120 min13.8% for visible light irradiation within 120 min. 0.00063 min−1 and 0.00390 min−1 for visible and UV light irradiation, respectively[92]
Ni-TiO2 nanoparticles (anatase phase)1.0% Ni- TiO20.05 g/0.5 gSol–gel method10 mg/L 4-CP, NPX (250 mL)UV68.9 and 84.9% for 4-CP and NPX, respectively within 6 h89.5 and 84% for 4-CP and NPX, respectively within 6 h[93]
Ni-TiO2 nanoparticles (anatase phase)1 wt% Ni-TiO20.1 gSol–gel method5 mg/L MO, MB (100 mL)Visible light44.18 and 26.80% degradation rates for MB and MO, respectively for unmodified TiO2 within 5 h.
45% of MB for Degussa P25 within 5 h.
71.18 and 39.57% for MB and MO, respectively within 5 h[94]
MWASG: Microwave-assisted sol–gel method; MB: methylene blue; MO: methyl orange; NPX: naproxen; ECR: Eriochrome cyanine red; IBP: ibuprofen; BPA: bisphenol A; MG: malachite green; 4-CP: 4-chlorophenol.
Table 5. Summary of recent progress on Cu-TiO2 photocatalysts for organic pollutants degradation.
Table 5. Summary of recent progress on Cu-TiO2 photocatalysts for organic pollutants degradation.
Materials (TiO2 Phase Transition)OptimumCatalyst DosageSynthetic MethodsPollutants (Volume)Light SourceUnmodifiedModifiedRef.
Cu-TiO2 nanoparticles (anatase phase)1.0% Cu-TiO20.5 g/0.05 gSol–gel method10 mg/L 4-CP (250 mL)UV79% (0.1210 L g−1 min−1) within 6 h.90% (0.1827 L/g−1 min−1) within 6 h[93]
Hybrid Cu-TiO2/polythiophene nanorods (rutile phase)-0.1 gSol–gel method5 mg/L RhB (150 mL)Visible light6.9% within 75 min70.5% within 75 min. Furthermore, there was no significant decrease in photocatalytic activity after three continuous cycles.[112]
Cu-TiO2 nanoparticles (anatase phase)1.0% Cu-TiO20.05 g/0.5 gSol–gel method10 mg/L NPX (250 mL)UV84.9% (0.0124 L g−1 min−1) within 6 h.87.4% (0.0259 L g−1 min−1) within 6 h.[93]
Hybrid Cu-TiO2/polythiophene nanorods (rutile phase)-0.1 gSol–gel method5 mg/L OG (150 mL)Visible light6.9% within 75 min98% within 75 min. Furthermore, there was no significant decrease in photocatalytic activity after three continuous cycles.[112]
Cu-TiO2 nanoparticles (anatase phase)0.2 wt% Cu-TiO2-Sol–gel method50 mg/L MO (50 mL)UV27% (0.45 min−1) within 30 min73% (2.83 min−1) within 30 min[113]
Cu-TiO2 nanocrystals (mixed anatase and brookite phases)0.1% Cu-TiO21 g/LSol–gel method10 mg/L PNP (10 mL)Visible irradiation (halogen lamp)12% for P25 within 24 h65% (5 times more efficient than P25) within 8 h[114]
Cu-TiO2 nanoparticles (stable anatase phase)2% Cu-TiO20.002 gSol–gel method10 mg/L MB (50 mL)150 W Xenon lamp0.0046 min−1 within 2 h0.0082 min−1 within 2 h[115]
Cu-TiO2 nanoparticles (stable anatase phase)2% Cu-TiO20.002 gSol–gel method10 mg/L PNP (50 mL)150 W Xenon lamp18% (0.0016 min−1) within 2 h29% (0.0027 min−1) within 2 h[115]
Cu-TiO2 nanoparticles (anatase to rutile phase)1.2% Cu-TiO2 0.08 gSol–gel method35 mg/L E131 VF (100 mL)UV0.083 min−1 within 30 min0.0044 min−1 within 100 min[116]
Cu-TiO2/GO nanoparticles (anatase phase)1 wt% Cu -TiO2/GO0.05 gImpregnation method20 mg/L TC (50 mL)300 W mercury lamp23.1% within 90 min for pure TiO298% within 90 min (reaction rate constant was about 1.4 times than that of TiO2/GO), the removal ratio of Cu-TiO2/GO exceeded 98% after five cycles.[117]
Cu-TiO2 nanopowders (anatase to rutile phase)Cu: TiO2 both in powder (42.35%)0.3 gSol–gel method10 mg/L MB (50 mL)Visible light2.6% (0.00008 min−1) within 6 h42.4% (0.00140 min−1) within 6 h[98]
Cu-TiO2 film (anatase to rutile phase)Cu: TiO2 both in film (25.10%)0.3 gSol–gel method10 mg/L MB (50 mL)Visible light2.6% (0.00008 min−1) within 6 h25.1% (0.00082 min−1) within 6 h[98]
Cu-TiO2 nanopowders (Anatase phase)2.0% Cu-TiO21 g/LSol–gel method20 mg/L DFC (100 mL)Visible light25% within 7 h33.26% within 7 h[118]
Cu-TiO2 nanopowders (Anatase to rutile phase)3% Cu-TiO20.125 mol L−1Sol–gel method20 mg/L MB (50 mL)Xenon Lamp0.08124 min−1 within 6 h0.00575 min−1 within 6 h[119]
Cu-TiO2 nanoparticles (anatase phase)-0.1 gMWASG30 mg/L MO (100 mL)Visible light1.4 × 10−3 min−1 within 6 h7.0 × 10−3 min−1 within 6 h[120]
Cu-TiO2 nanoparticles (anatase phase)-0.1 gMWASG30 mg/L MB (100 mL)Visible light7.0 × 10−4 min−1 within 6 h5.6 × 10−3 min−1 within 6 h[120]
Cu-TiO2 nanoparticles (anatase phase)-0.1 gMWASG30 mg/L MO (100 mL)UV5 × 10−3 min−1 within 6 h1.2 × 10−2 min−1 within 6 h[120]
Cu-TiO2 nanoparticles (anatase phase)-0.1 gMWASG30 mg/L MB (100 mL)UV2.5 × 10−3 min−1 within 6 h8.6 × 10−3 min−1 within 6 h[120]
Cu-TiO2 nanoparticles (anatase phase)2 wt% Cu-TiO20.38 g/LSol–gel method10 mg/L Phenol (100 mL)UV82% within 60 min for P2598% within 60 min[121]
Cu-TiO2 nanoparticles (anatase phase)2 wt% Cu-TiO20.38 g/LSol–gel method10 mg/L Phenol (100 mL)Visible light35 and 75% after 60 min and 180 min, respectively for P2522 and 37% after 60 min and 180 min, respectively[121]
Cu-TiO2 nanopowders (anatase phase)0.21 mol% Cu-TiO23 g·dm−3Sol–gel method20 mg/L 2-CP (50 mL)UV-98.92% within 6 h[122]
Cu/TiO2/bentonite composite nanoparticles (anatase phase)-0.02 gThermal decomposition and reduction method10 mg/L Deltamethrin insecticide (50 mL)Sunlight87.01% within 120 min for TiO2/bentonite97% within 120 min[123]
Cu-TiO2 nanoparticles (anatase phase)3% Cu-TiO20.1 gSol–gel method15 mg/L MO (100 mL)Xenon Lamp0.0011 min−1 within 60 min61% (0.0166, min−1) within 60 min[124]
Cu-TiO2 films (anatase phase)4% Cu-TiO20.1 gSol–gel method25 mg/L MB (100 mL)UV92% ((0.015 min−1) within 180 min16% (0.001 min−1) within 180 min[125]
Cu/TiO2 nanoparticles (anatase to rutile phase)10% Cu-TiO20.5 gCoprecipitation method100 mg/L ECR (200 mL)Visible light20.9% within 120 min60.6% within 120 min[89]
Cu-TiO2 nanoparticles (anatase phase)1% Cu-TiO22 g/LSol–gel method100 mg/L MO, MB (100 mL)Visible light45% for MB within 5 h for Degussa P2581.22 and 44.05% for MB and MO, respectively within 5 h.[94]
Cu-TiO2 nanoparticles (anatase phase)5% Cu-TiO20.5 g/LSol–gel method60 mg/L Orange II
(35 mL)
UV95% for P25 within 180 min82% within 180 min[126]
Cu-TiO2 nanoparticles (anatase phase)5% Cu-TiO20.5 g/LImpregnation method60 mg/L Orange II (35 mL)UV99% for P25 within 180 min90% within 180 min[126]
Cu2+/TiO2 nanoparticles (mixed anatase and brookite phases)0.5 mol% Cu2+/TiO21 g/LSol–gel method10 mg/L PNP
(100 mL)
UV/Visible light20 and 50% and under visible and UV/Visible light, respectively for Degussa P25 within 24 h42 and 55% under visible and UV/visible light, respectively within 24 h[127]
Cu-TiO2 nanoparticles (mixed anatase and rutile phases)-0.1 gSol–gel method20 mg/L MB (50 mL)Visible light89.69% within 60 min27.5% within 60 min[128]
MB: methylene blue; PNP: p-nitrophenol; MO: methyl orange; 2-CP: 2-chlorophenol; ECR: Eriochrome cyanine red; DFC: diclofenac; TC: tetracycline; OG: orange G; E131 VF: food colorant; MWASG: microwave-assisted sol–gel method; NPX: naproxen; RhB: rhodamine B; 4-CP: 4-chlorophenol.
Table 6. Summary of recent progress on Mn-TiO2 photocatalysts for organic pollutants degradation.
Table 6. Summary of recent progress on Mn-TiO2 photocatalysts for organic pollutants degradation.
Materials (TiO2 Phase Transition)OptimumCatalyst DosageSynthetic MethodsPollutants (Volume)Light SourceUnmodifiedModifiedRef.
Mn-TiO2 nanoparticles (anatase phase)10% Mn-TiO20.025 gSol–gel method10−5 mol/L MO (50 mL)UV/Visible light0.00075 min−1 and 0.002631 min−1 for visible light and UV light, respectively within 120 min.
95.6 and 16.3% under UV light and visible light, respectively
0.00070 min−1 and 0.00074 min−1 for visible light and UV light, respectively within 120 min.
Modifying with Mn presents a negative effect on the photocatalytic
degradation of MO.
12.8% under UV light
[92]
Mn2+/TiO2 nanoparticles (mixed crystal of anatase and rutile phases)1.90 wt% Mn2+/TiO20.05 gHydrothermal method50 mg/L MG (50 mL)Visible light48% within 180 min for pure TiO281% within 180 min[150]
Mn-TiO2 nanopowders (anatase phase)0.6% Mn-TiO20.05 g/LSol–gel method10 mg/L DCF (100 mL)UV51% within 240 min0.0033 min−1 within 240 min[151]
Mn-TiO2 nanoparticles (mixture of anatase and rutile phases)0.1% Mn-TiO2 0.1 gModified sol–gel methodBenzene, Xylene, toluene (100 mL)UV/Visible light60, 60 and 55% for benzene, xylene and toluene, respectively under UV light.60, 60, and 40% for benzene, xylene and toluene, respectively under UV light.
34 and 22% for toluene and xylene, respectively under visible irradiation within 60 min
[152]
Mn-TiO2 nanopowders4 wt% Mn-TiO20.1 g/LHydrothermal route10 mg/L Phenol (50 mL)UV30% within 7 h90% within 7 h[153]
Mn-TiO2
nanoparticles (anatase phase)
1 wt% Mn-TiO21 gSol–gel method30 mg/L Toluene (100 mL)VUV70% without photocatalyst89.8% within 120 min[154]
Mn-TiO2 nanoparticles (anatase phase)0.3 wt% Mn-TiO20.1 gSol–gel method5 mg/L MO, MB (100 mL)Visible light26.80 and 44.18% degradation rates for MO and MB, respectively for unmodified TiO2 within 5 h.
45% of MB for Degussa P25 within 5 h.
15.48 and 16.41% for MO and MB, respectively within 5 h[94]
Mn-TiO2
films (anatase phase)
5 wt% Mn-TiO20.5 g/LSol–gel method6 mg/L Orange II
(50 mL)
UV95% for P25 within 180 min97% within 180 min[126]
Mn-TiO2 nanoparticles (anatase-brookite phase)5 wt% Mn-TiO20.5 g/LImpregnation method6 mg/L Orange II (100 mL)UV99% for P25 within 180 min97% within 180 min[126]
Mn-TiO2 nanoparticles (mixed anatase and brookite phases)0.5 mol% Mn-TiO21 g/LSol–gel method10 mg/L PNP (100 mL)UV/Visible light50 and 20% under UV/Visible and visible light irradiation for Degussa P2 within 8 h.19 and 5% under UV/Visible and visible light irradiation, respectively within 8 h.[127]
Mn-TiO2 nanoparticles (anatase phase)5 wt% Mn-TiO2 0.1 gSol–gel method5 mg/L MB, MO (100 mL)UV/Visible light65.3 and 72.4% under UV irradiation for MO and MB, respectively within 60 min.
4.1 and 6.2% for MO and MB, respectively under visible irradiation within 240 min
85.2.1 and 93.6% under UV irradiation for MO and MB, respectively within 60 min.
65.2 and 73.2% for MO and MB, respectively under visible irradiation within 240 min
[38]
MB: methylene blue; PNP: p-nitrophenol; MO: methyl orange; DCF: diclofenac.
Table 7. Summary of recent progress on Zr-TiO2 photocatalysts for organic pollutants degradation.
Table 7. Summary of recent progress on Zr-TiO2 photocatalysts for organic pollutants degradation.
Materials (TiO2 Phase Transition)OptimumCatalyst DosageSynthetic MethodsPollutant (Volume)Light SourceUnmodifiedModifiedRef.
Zr-TiO2 nanoparticles (mixed anatase and rutile phases)0.01% Zr-TiO2 0.1 gSol–gel method20 mg/L MB
(50 mL)
Visible light89.69% within 60 min81.9% within 60 min[128]
Zr-TiO2 films (stable anatase phase)0.05% Zr/TiO20.005 gEISA via dip-coating0.001 mM herbicide chloridazon, henol, 4-CP (50 mL)Xenon lamp42% within 4 h for phenol, 58% within 4 h for 4-CPThe Zr-TiO2 film showed the superior photocatalytic activity for all pollutants but showed lowest rate using phenol[167]
Zr-TiO2 nanoparticles (anatase phase)0.2% Zr-TiO20.4 gSol–gel method1.0±0.5 mg/m3 formaldehydeUV10% within 48 h for P2595.14% within 48 h.
Furthermore, 94.38% removal efficiency was achieved even after seven cycles
[168]
Zr-TiO2 nanoparticles (anatase phase)Zr:TiO2 = 0.08:0.920.005 gEISA method0.001 mM formaldehydeXenon lamp45% within 48 h for TiO292% after 48 h[166]
Zr-TiO2 nanoparticles (anatase phase)5% Zr-TiO20.04 g/LA combined sol–gel and CVD method10 ppm IBPUV-80% after 40 min[33]
Zr-TiO2 nanoparticles (mixed anatase and brookite phase)2.0% Zr-TiO21 g/LSol–gel method10 mg/L PNP (100 mL)UVA/Visible light20 and 84% for pure TiO2 and commercial Evonik P25, respectively UV-A light.
25 and 12% for pure TiO2 and Evonik P25, respectively under visible irradiation within 17 h
96% under UV-A light, 38% under visible irradiation within 17 h[169]
Zr-TiO2 nanoparticles (anatase phase)0.5% Zr-TiO20.25 gSol–gel method10 mg/L antipyrine (phenazone) (10 mL)Visible light-90% within 360 min.
Furthermore, 8% lower degradation rate was achieved, even after six hours of irradiation
[170]
Zr-TiO2 hollow microspheres (anatase phase)12.6 wt% Zr-TiO20.02 gFacile solvothermal method20 mg/L RhB (50 mL)UV14.6% within 60 min96.3% within 60 min[171]
Zr-TiO2 nanoparticles (mixed anatase and brookite phase)2.0% Zr-TiO21 g/LSol–gel method10 mg/L PNP (100 mL)UV/Visible light20 and 84% for pure TiO2 and commercial Evonik P25, respectively within 24 h under UV–Visible light.
25 and 12% for pure TiO2 and commercial Evonik P25, respectively within 24 h under visible light.
96% within 24 h under UV-Visible light.
38% within 24 h under visible light
[169]
Zr-TiO2/reduced Graphene Oxide nanoparticles (mixed anatase and brookite phase)3% ZrO2–TiO20.025 gSol–gel method20 mg/L EB (100 mL)Visible light-75.75% within 90 min[172]
ZrO2/TiO2 nanofibers (anatase phase)40 wt% ZrO2/TiO2 (Zr:Ti = 1:3)0.02 gSol–gel method10 mg/L MB (50 mL)Mercury lamp43.3% within 180 min.82.7% within 180 min[173]
ZrO2/TiO2 nanoparticles (anatase phase)6.9% ZrO2–TiO20.02 gFacile surfactant self-assembly5 mg/L RhB (100 mL)UV52.5% within 3 h.86.9% within 3 h. Rate constant was 3.0 times higher than unmodified TiO2 within 3 h[165]
ZrO2/TiO2 nanoparticles (anatase phase)5% ZrO2/TiO20.1 gSol–gel method5 × 10−5 mol/L MB, MO (100 mL)UV/Visible light65.3 and 72.4% for MO and MB, respectively within 60 min under UV light.
4.1 and 6.2% for MO and MB, respectively within 240 min under visible light.
70.1 and 77.3% for MO and MB, respectively within 60 min under UV light.
22.6 and 38.2% for MO and MB, respectively within 240 min under visible light.
[38]
Zr-TiO2 nanoparticles (anatase phase)6 mol% Zr-TiO21 g/LSol–gel method5 mg/L Ponceau BS (180 mL)UV24.45% (0.0107 min−1) within 30 min99.3% (0.1810 min−1) within 30 min[155]
PNP: p-nitrophenol; MO: methyl orange; RhB: rhodamine B; MB: methylene blue; EB: eosin blue; 4-CP: 4-chlorophenol; IBP: ibuprofen.
Table 8. Summary of recent progress on Fe-TiO2 photocatalysts for organic pollutants degradation.
Table 8. Summary of recent progress on Fe-TiO2 photocatalysts for organic pollutants degradation.
Materials (TiO2 Phase Transition)OptimumCatalyst DosageSynthetic MethodsPollutants (Volume)Light SourceUnmodifiedModifiedRef.
Fe-TiO2 nanopowders (mixture of anatase and rutile phases)2wt% Fe-TiO2 0.01 gConventional solid-state reaction20 mg/L MB (50 mL)Xenon arc lamp90.80% within 210 min94.93% within 210 min[182]
Fe-TiO2 nanoparticles (anatase phase)1 wt% Fe-TiO2 0.1 gSol–gel method5 mg/L MO, MB (100 mL)Visible light44.18 and 26.80% degradation rates for MB and MO, respectively for unmodified TiO2 within 5 h; 45% for MB using Degussa P25 within 5 h1.41 and 7.94% for MB and MO, respectively within 5 h[94]
Fe-TiO2 nanoparticles (anatase phase)0.1 mol% Fe-TiO2 0.03 g/LPrecipitation methodMO, 4-CP (100 mL)Xenon arc lamp30 and 45% of 4-CP and MO, respectively within 4 h65 and 95% for 4-CP and MO, respectively within 4 h[183]
Fe-TiO2 nanoparticles (anatase phase)0.0017 mol% Fe-TiO23 g/LSol–gel method10 mg/L AO7 (100 mL)Visible light10% within 60 min80% within 60 min[184]
Fe-TiO2 nanoparticles (anatase phase)1.0% Fe-TiO20.05 g/0.5 gSol–gel method10 mg/L NPX, 4-CP (100 mL)UV68.9% (0.1069 L g−1 min−1) and 84.9% (0.0124 L g−1 min−1) for 4-CP and NPX, respectively for unmodified TiO2 within 6 h97.7% (0.1111 L g−1 min−1) and 37% (0.0053 L g−1 min−1) for NPX and 4-CP, respectively within 6 h[93]
Fe-TiO2 nanoparticles (anatase phase)0.5% Fe-TiO20.5 g/LSol–gel method10 mg/L Phenol (200 mL)Visible light33% within 90 min57% (about 73% increase) within 90 min. that the activity of pristine TiO2 was higher than that of the modified catalysts with high content[185]
Fe-TiO2
nanoparticles (anatase phase)
0.1 wt% Fe-TiO20.5 gSol–gel method20 mg/L AO7 (500 mL)Visible light25% within 300 min54% within 300 min[186]
Fe-TiO2
nanoparticles (mixture of anatase and rutile phases)
0.05 mol% Fe-TiO20.5 g/LFacile ultrasonic assisted hydrothermal method10 mg/L PNP (50 mL)Visible light42% within 5 h for pure TiO292% within 5 h[187]
Fe-TiO2 nanoparticles (mixture of anatase and rutile phases)0.075 mol% Fe-TiO2 NPs0.5 g/LFacile ultrasonic assisted hydrothermal method10 mg/L MB (50 mL)Visible light50% within 180 min for pure TiO293% within 180 min[187]
Fe3+/TiO2 nanoparticles (anatase phase)3% Fe3+/TiO225 cm × 25 cmSol–gel spin coating technique5 mg/L MO (50 mL)Visible light34% within 360 min.95.25% within 360 min.[188]
Fe3+/TiO2 nanoparticles (anatase phase)3% Fe3+/TiO225 cm × 25 cmSol–gel spin coating technique5 mg/L MB (50 mL)Visible light 34% within 360 min.Fe-TiO2 (3% Fe3+) film exhibits high efficiency of about 97.8% after 10 cycling runs of MB degradation.[188]
Fe-TiO2 nanoparticles (anatase phase)0.5%Fe-TiO20.20 g/LWet-impregnation method5 ppm MO (200 mL)UV/Visible light 85 and 11% under UV light and visible light irradiation, respectively within 2 h.82.8 and 74.4% under UV light and visible light irradiation, respectively. within 2 h[189]
Fe-TiO2 thin films (anatase phase)0.5% Fe-TiO2-Sol–gel method2.51 × 10−4 M NB (-)UV70.52% within 240 min88.45% within 240 min[190]
Fe-TiO2 nanoparticles (mixture of anatase and rutile phases)2.5 wt% Fe-TiO20.1 gSol–gel method10 mg/L MB (50 mL)UV ~17% within 240 min. 54% within 90 min40% within 240 min. 78% within 90 min.
Furthermore, the photocatalytic activity only decreased 7.7% after three runs.
[191]
Fe-TiO2 nanoparticles (anatase phase)0.5 mol% Fe-TiO20.5 g/LUltrasonic dispersion method10 mg/L RhB (50 mL)UV 60% within 150 min91.11% within 150 min[192]
Fe-TiO2 microsized powder (anatase phase)-0.2 gsol–gel and hydrothermal treatment50 mg/L BO2 (3 L)Solar25% within 180 min.≥90% within 180 min.[193]
Fe-TiO2 nanoparticles (anatase phase)6 wt% Fe-TiO2-Sol–gel method10 mg/L MB (100 mL)Visible light36% without irradiation for Fe-TiO299.5% within 3 h[194]
Fe-TiO2 nanoparticles (anatase phase)0.4 wt% Fe-TiO20.2 g/LSol–gel method10 mg/L Acid Blue 80 (50 mL)UV27.86% (1.6 × 10−3 min−1) within 120 min31.13% (2.0 × 10−3 min−1) within 120 min[195]
Fe-TiO2 nanoparticles (anatase phase)0.4 wt% Fe-TiO20.2 g/LUltrasound assisted approach10 mg/L Acid Blue 80 (50 mL)UV27.86% (1.6 × 10−3 min−1) within 120 min38.01% (2.4 × 10−3 min−1) within 120 min[195]
Fe-TiO2 nanoparticles (anatase phase)1.5 wt% Fe-TiO20.4 g/LHydrothermal20 mg/L Diazinon (100 mL)UV~20% within 100 min75% within 100 min[196]
Fe-TiO2 nanoparticles (anatase phase)10% Fe-TiO20.005 gSol–gel method20 mg/L MB (250 mL)UV76.09% within 150 min96.66% within 150 min[197]
Fe-TiO2 nanoparticles (anatase phase)2% Fe-TiO20.5 g/LSol–gel method50 mg/L AO7 (2.5 L)UV/Visible/Solar light9.2% within 6 hUV (100%), visible (100%) and solar light (90%) within 6 h.
Furthermore, the Fe-TiO2 photocatalysts were stable and can maintain performance up to 6 recycle use.
[198]
Fe-TiO2 nanotubes (anatase phase)Hydrothermal temperature of 150 °C for 3 h-Hydrothermal5 mg/L CR (35 mL)Visible light26.32% within 180 min92.5% within 180 min[199]
Fe-TiO2 thin films (anatase phase)0.02% Fe/TiO2-Sol–gel method20 mg/L MB (100 mL)Visible light~84 within 200 minNo enhancement in photocatalytic activity of Fe-TiO2 thin films was achieved. This was be attributed to sodium diffusion from the substrate used.[200]
Fe3+/TiO2 nanoparticles (anatase phase)7% Fe3+/TiO2-Simple spin coating technique3 mg/L MB (50 mL)Xenon arc lamp80% within 4 h96.7% within 4 h. Furthermore, the degradation rate of MB was ~83.8% after 10 cyclic runs.[201]
Fe-TiO2 nanoparticles (anatase phase)0.5 wt% Fe-TiO2-Sol–gel method20 mg/L MO (200 mL)Visible light24% within 60 min98% within 60 min[202]
Fe-TiO2
nanoparticles (anatase phase)
3% Fe-TiO20.2 gSol–gel method20 mg/L RhB (100 mL)Solar34% within 120 min64% within 120 min[203]
Fe-TiO2 nanoparticles (anatase-brookite phase)5 wt% Fe-TiO20.5 g/LImpregnation method6 mg/L Orange II (100 mL)UV99% for P25 within min69% within 180 min[126]
Fe-TiO2 nanoparticles (anatase-brookite phase)5 wt% Fe-TiO20.5 g/LSol–gel method6 mg/L Orange II (100 mL)UV95% for P25 within min95% within 180 min[126]
Fe-TiO2 nanopowders (anatase phase)0.5% Fe-TiO20.1 g/LSol–gel method2.45 × 104 M NB (-)Mercury lamp32.14% within 240 min for pure TiO284.91% within 240 min[204]
Fe-TiO2 nanopowders (anatase phase)1 wt% Fe-TiO20.1 gSol–gel method10 mg/L MO, 4-CP (100 mL)Visible light31 and 28% for MO and 4-CP, respectively within 180 min54 and 49% for MO and 4-CP, respectively within 180 min[205]
Fe3+/TiO2 nanoparticles (mixed anatase and brookite phases)0.5 mol% Fe3+/TiO21 g/LSol–gel method10 mg/L PNP (100 mL)UV/Visible light 50 and 20% under UV/Visible light and visible light, respectively using Degussa P25 within 8 h56 and 44% under UV/Visible light and visible light, respectively within 8 h[127]
Fe-TiO2 nanoparticles (mixture of anatase and rutile phases)2 wt% Fe-TiO20.25 g/LSurface impregnation method0.0755 mg/L Carbendazim and 0.0096 mg/L for propiconazole(-)Sunlight -98.5 and 92% of carbendazim and propiconazole, respectively within 60 min[206]
Fe-TiO2 nanotube (anatase phase)1% Fe-TNAs0.001 g/Lsolvothermal method20 mg/L MB (100 mL)Solar -98.79% within 120 min[207]
Fe-TiO2 nanotube (anatase phase)0.1 mol% Fe-TiO2 1 gSol–gel method10 mg/L AR-27 (1000 mL) Visible light78% within 2 h.99% within 2 h. The removal efficiency was kept at 89% after catalyst used for 4 times.[208]
Fe-TiO2 thin films (anatase phase)25 wt% Fe-TiO2-Sol–gel method4 mg/L MB (200 mL)Visible light0.018/h within 8 h0.038/h within 8 h[209]
Fe-TiO2 nanoparticles (anatase phase)1 wt% Fe-TiO20.03 gSol–gel method10 mg/L Orange II (500 mL)UV/Visible light 30% within 60 min24% within 60 min[210]
Fe3+/TiO2 nanoparticles (anatase phase)0.1% Fe3+/TiO2 0.2 gSol–gel method10 mg/L MB (100 mL)Xenon arc lamp90% within 6 h0.716 min−1 within 6 h[211]
Fe3+/TiO2 thin films (anatase phase)25 wt% Fe3+/TiO2 -Sol–gel method10 mg/L MB (100 mL)Visible light13.4% within 8 h24.9% within 8 h[212]
Fe-TiO2 nanoparticles (anatase phase)0.5 wt% Fe-TiO20.25 g/LSol–gel method(0.37–8.45) × 10−4 M NB (-)Mercury lamp48% within 240 min99% within 240 min[213]
Fe-TiO2 nanotube (anatase phase)5 wt% Fe/TNAs0.5 mol·L−1Electrochemical anodization and subsequent dip-coating10 mg/L BPA (200 mL)Xenon lamp-18.3% within 240 min[214]
Fe-TiO2 nanoparticles (anatase phase)0.15% Fe-TiO2 0.3 g/LSol–gel method20 mg/L RB5 (250 mL)Xenon lamp0.0180 min−1 within 60 min0.0875 min−1 within 60 min[215]
Fe-TiO2 nanoparticles (anatase phase)0.5 wt.% Fe-TiO2 0.25 g/LSol–gel method(0.37–8.45) × 10−4 M NB (-)Medium-pressure
mercury lamp
48% within 240 min99% within 240 min[213]
Fe-TiO2 nanoparticles (anatase phase)1.0% Fe-TiO20.05 g/0.5 gSol–gel method10 mg/L NPX,4-CP (100 mL)UV79% (0.1210 L g−1 min−1) within 6 h97% (0.1111 L g−1 min−1) within 6 h[93]
Fe-TiO2 nanoparticles (anatase phase)0.05 wt% Fe-TiO20.001 g/cm2Sol–gel method25 mg/L DB15 (100 mL)UV3.3% within 1 h31% within 1 h[216]
Fe-TiO2 nanoparticles (anatase phase)1.75 wt% Fe-TiO20.03 g/LNanosol/dip-coating method. 30 mg/L MB (50 mL)UV-76% within 2 h[217]
A07: acid orange 7; RB5: reactive black 5; 4-CP: 4-chlorophenol; NPX: naproxen; MB: methylene blue; MO: methyl orange; BO2: basic orange 2; RR 198: reactive red 198; PNP: p-nitrophenol; AR-27: acid red 27; NB: nitrobenzene; BPA: bisphenol A; RhB: rhodamine B; NB: nitrobenzene; CR: Congo red; PNP: 4-nitrophenol.
Table 9. Summary of recent progress on Cr-TiO2 photocatalysts for organic pollutants degradation.
Table 9. Summary of recent progress on Cr-TiO2 photocatalysts for organic pollutants degradation.
Materials (TiO2 Phase Transition)OptimumCatalyst DosageSynthetic MethodsPollutantsLight SourceUnmodifiedModifiedRef.
Cr-TiO2 nanoparticles (anatase phases)0.5 wt% Cr-TiO21.5 g/LSol–gel method10 mg/L MB (100 mL)Hg vapor lamp62% within 120 min for pure TiO258% within 120 min[84]
Cr-TiO2 nanoparticles (anatase phase)1 mol% Cr0.1 gSol–gel method15 mg/L MB (50 mL)UV33.43% for pure anatase within 24 h50.01% within 24 h[222]
Cr-TiO2 nanoparticles (anatase phase)4 mol% Cr0.1 gSol–gel method100 ppm CR (50 mL)UV0.73% for pure anatase within 24 h17.78% within 24 h[222]
Cr-TiO2 nanoparticles (anatase phase)10% Cr-TiO20.025 gSol–gel method10−5 mol/L MO (50 mL)UV/Visible light 95.6 and 16.35% under UV and visible light irradiation, respectively within 120 min0.00276 min−1 (3.7 larger than that of the unmodified TiO2 (0.00075 min−1) within 120 min[92]
Cr-TiO2 nanoparticles (anatase phase)1 wt% Cr-TiO20.1 gSol–gel method5 mg/L MO, MB (100 mL)Visible light44.18 and 26.80% for MB and MO, respectively for unmodified TiO2.
45% for MB within 5 h for Degussa P25.
11.47 and 15.48% for MB and MO, respectively within 5 h.[94]
Cr-TiO2 nanoparticles (anatase phase)0.5% Cr-TiO20.008 gSol–gel method10 mg/L 4-CP (100 mL)Visible light76.5% within 390 min90.7% within 390 min[223]
Cr-TiO2 nanoparticles (anatase phase)-0.4 gFSP, coprecipitation, and sol–gel synthesis techniques50 μM 4-CP (50 mL)Visible light2% within 4 h61% within 4 h.[224]
Cr-TiO2 nanoparticles
(mixed anatase and brookite phases)
0.5 mol% Cr-TiO21 g/LSol–gel method10 mg/L PNP (100 mL)UV/Visible light 50 and 20% under UV/Visible and visible light, respectively for Degussa P25 within 8 h.38 and 27% under UV/Visible and visible light, respectively within 8 h.[127]
4-CP: 4-chlorophenol; FSP: flame spray pyrolysis; MO: methyl orange; PNP: p-nitrophenol; MB: Methyl blue; CR: Congo red.
Table 10. Summary of recent progress on Mo-TiO2 photocatalysts for organic pollutants degradation
Table 10. Summary of recent progress on Mo-TiO2 photocatalysts for organic pollutants degradation
Materials (TiO2 Phase Transition)OptimumCatalyst DosageSynthetic MethodsPollutants (Volume)Light SourceUnmodifiedModifiedRef.
Mo-TiO2 nanocrystals (anatase phase)5% Mo-TiO20.03 gFacile hydrothermal method25 ppm Benzene (-)UV13.6% within 4 h79.1% (0.0065 min−1, which was 4.81 times higher than the unmodified TiO2) within 4 h. Furthermore, the degradation ratio of benzene at the sixth run was 76.4%.[234]
Mo-TiO2 layer (rutile to stable anatase phase)6.0% Mo-Ti O215 mm × 10 mmPEOx10 mg/L MB (100 mL)660-nm
Tungsten-halogen
55% within 180 min for pure TiO286% within 180 min[76]
Mo-TiO2 nanoparticles (anatase phase)1 wt% Mo-TiO20.2 g/LEISA method0.156 mmol L−1 4-CP (1.5 L)UV32% within 100 min95% (three times faster than Degussa P25) within 100 min[233]
Mo-TiO2 nanoparticles (anatase to brookite and rutile
phases)
0.125 mol% Mo-TiO20.1 g/LMWASG85 mg/L MB, SMX (100 mL)UV0.031 min−1 for MB within 64 min0.04 and 0.0318 min−1 for MB and SMX, respectively within 64 min[77]
MoO3/P25 nanoparticles (anatase phase)0.25% MoO3/P250.1 gImpregnation method15 mg/L MB (100 mL)High pressure sodium lamp8% within 150 min38% within 150 min[235]
MB: Methylene blue; SMX: sulfamethoxazole; PEOx: plasma electrolytic oxidation; EISA: evaporation induced self-assembly; MWASG: microwave-assisted sol–gel method; 4-CP: 4-chlorophenol.
Table 11. Summary of recent progress on Co-TiO2 photocatalysts for organic pollutants degradation.
Table 11. Summary of recent progress on Co-TiO2 photocatalysts for organic pollutants degradation.
Materials (TiO2 Phase Transition)OptimumCatalyst DosageSynthetic MethodsPollutants (Volume)Light SourceUnmodifiedModifiedRef.
Co-TiO2
nanosheets (anatase phase)
2.81% Co-TiO2100 mgHydrothermal route30 mg/L TC (100 mL)Visible light51.56% within min50.19% within 140 min.[247]
Co-TiO2 nanoparticles (anatase phase)1 wt% Co-TiO20.5 g/LSol–gel and precipitation methods10 mg/L MO (250 mL)White compact fluorescent lamp16.3% within 360 min.34.7% within 360 min[248]
Cobalt phthalocyanine complex sensitized TiO2 nanoparticles (anatase phase)-0.3 g/LHydrothermal methods40 mg/L 4-CP (200 mL)Visible light25% within 90 min.50% within 90 min.[249]
Co-TiO2 film (anatase phase)8% Co-TiO2-MBE methods20 mg/L MB, Azo (100 mL)Visible light38 and ~37% for MB and AZO dyes, respectively within 70 min.91 and 88% for MB and Azo dye, respectively within 70 min[239]
Co-TiO2 nanosheets (Co-TNT) (anatase phase) 0.152% Co- TNT/rGO0.152 gHydrothermal methods30 mg/L TC (100 mL)Visible light48.57% within 180 min.~60% within 180 min. Reusing the optimal photocatalyst after five successive cycles showed ~7% decline in its activity for degrading of TC.[250]
Co-TiO2 nanoparticles (anatase phase)0.06% Co/TiO20.1 gSol–gel using cobalt resinate as both template and cobalt source.10 mg/L MB, RhB (100 mL)Visible lightUnmodified TiO2 and unmodified Degussa P25 titania (P25) was 55 and 4.7%, respectively within 4 h.82 and 55% for MB and RhB, respectively within 4 h[251]
Co-TiO2 nanoparticles (anatase phase)0.060% Co/TiO20.1 gSol–gel using resin acids as both template and cobalt source.10 mg/L MB, RhB (100 mL)Visible lightUnmodified TiO2 and unmodified Degussa P25 titania (P25) were 5.2 and 4.7%, respectively within 4 h.86 and 94% for MB and RhB, respectively within 4 h[251]
Co-TiO2 nanoparticles (anatase phase)1 wt% Co-TiO2 NPs0.5 gSol gel method40 mg/L 2, 4-DCP (100 mL)Visible light43.65% within 120 min68.03% within 120 min[252]
Co-TiO2 flim (anatase phase)1 wt% Co-TiO20.5 gPhase inversion technique40 mg/L 2, 4-DCP (100 mL)UV/Visible light -61.6 and 63.74% in the presence of UV and visible light, respectively within 120 min. Furthermore, the membrane flux was increased by 53%.[252]
Co-TiO2 nanoparticles (anatase phase)2 wt% Co-TiO2 NPs0.38 g/LPhotochemical deposition-assisted sol–gel technique. 10 mg/L Phenol (120 mL)UV/Visible light~95 and ~58% in the presence of UV and visible light, respectively for neat TiO2 within 180 min.
Moreover, the neat P25 shows ~85 and ~80% in the presence of UV and visible light, respectively.
~98 and ~67% in the presence of UV and visible light, respectively within 180 min[121]
Co-TiO2 nanopowders (anatase phase)1 wt% Co0.1 g/LSol–gel method25 mg/L NB (-)Medium-pressure mercury lamp-~81% within 180 min[253]
Co-TiO2 nanopowders (anatase phase)2 wt% Co- TiO20.01 gConventional solid-state reaction20 mg/L MB (100 mL)Xenon arc lamp90.80% within 210 min.90.92% within 210 min[182]
Co-TiO2 nanoparticles (anatase phase)1% Co-TiO2 0.1 gSol–gel method5 mg/L MO, MB (100 mL)Visible light44.18 and 26.80% degradation rates for MB and MO, respectively for unmodified TiO2 within 5 h.
45% for MB for Degussa P25
66.17 and 38.14% for MB and MO, respectively within 5 h.[94]
Co-TiO2 nanoparticles (anatase phase)0.05% Co-TiO22 g/LSol–gel method75 μM Crystal violet (100 mL)UV60% within 120 min~100% within 120 min[254]
CoO-RGO-TiO2 nanoparticles (anatase phase)0.5% TiO2–RGO–CoO0.05 gSol–gel method10 mg/L 2-CP (100 mL)Visible light35.7% within 12 h.58.9% within 12 h[255]
Co-TiO2 nanoparticles (anatase phase) c1% Co-TiO20.01 gSol–gel method10 mg/L RhB (10 mL)Visible light~86% within 40 min99% within 40 min[237]
Co-TiO2 nanopowders (anatase phase)1% Co-TiO20.1 g/LSol–gel method2.5 × 10−4 M NB (-)UV57.13% within 2 h81.03% within 2 h[256]{FormattingCitation}
Co-TiO2 nanopowders (anatase phase)2% Co-TiO20.1 g/LSol–gel method2.5 × 10−4 M NB (-)UV57.13% within 2 h67.17% within 2 h[256]{FormattingCitation}
Co-TiO2/(GO) inverse opal photonic crystal (IO PC) (anatase phase)-0.1 gSimple casting and calcination process using polystyrene opal as the template0.013 M 4-CP (100 mL) LT50 lamp60% within 120 min.75% within 120 min[257]
Co-TiO2 nanopowders (anatase phase)0.05 mol% Co-TiO21 g L−1Sol–gel method (non-aqueous)10 mg L−1 MB (500 mL)UV88.0% within 120 min97.7% within 120 min[91]
Co-TiO2 nanoparticles (anatase phase)1.75% Co-TiO20.03 g/LNanosol/dip-coating method.30 mg/L MB (50 mL)UV-69% within 2 h[217]
Co-TiO2 nanoparticles (anatase phase)1% Co-TiO20.025 gSol–gel method10−5 mol/L MO (50 mL)Visible light0.00075 and 0.002631 min−1 for visible and UV irradiation, respectively. 16.3 and 95.6% for visible and 95.6 for visible and UV light, irradiation within 120 min0.00149 and 0.00280 min−1 (2 times larger than that of the unmodified TiO2) for visible and UV irradiation, respectively.[92]
Co-TiO2 nanoparticles (anatase phase)5% Co-TiO20.1 gSol–gel method5×10−5 mol/L MB, MO (100 mL)UV/Visible light 65.3 and 72.4% for MO and MB, respectively within 60 min under UV irradiation.
4.1 and 6.2% for MO and MB, respectively within 240 min under visible irradiation.
78.2 and 86.4% for MO and MB, respectively within 60 min under UV irradiation.
53.2 and 64.3% for MO and MB, respectively within 240 min under visible irradiation.
[38]
Co-TiO2 nanoparticles (mixed anatase and brookite phases)0.5 mol% Co-TiO21 g/LSol–gel method10 mg/L PNP (100 mL)UV/Visible light 50 and 20% under UV/Visible and visible light, respectively for Degussa P25 within 8 h.32 and 15% under UV/Visible and visible light, respectively within 8 h.[127]
2,4-DCP: 2,4-dichlorophenol; Co-TiO2: cobalt modified TiO2; PNP: p-nitrophenol; MO: methyl orange; RhB: rhodamine B; MB: methyl blue; NB: nitrobenzene; 4-CP: 4-chlorophenol; 2-CP: 2-chlorophenol; TC: tetracycline.
Table 12. Summary of recent progress on Nb-TiO2 photocatalysts for organic pollutants degradation.
Table 12. Summary of recent progress on Nb-TiO2 photocatalysts for organic pollutants degradation.
Materials (TiO2 Phase Transition)OptimumCatalyst DosageSynthetic MethodsPollutants (Volume)Light SourceUnmodifiedModifiedRef.
Nb-TiO2
Nanoparticles (Anatase phase)
0.5 mol%Nb-TiO21 g/LSol–gel method100 uM 4-CP (-)Visible light8% within 120 minThe degradation of Nb-TiO2 was only slightly enhanced from the bare TiO2 case.39% within120 min[264]
Nb-TiO2 nanoparticles (anatase phase)0.5 at% Nb-TiO20.5 g/LSol–gel method100 uM 4-CP (30 mL)Visible light~90% within 4 h~100% within 4 h[265]
Nb2O5/TiO2 hexagonal (anatase phase)60 wt% or 0.6 of Nb2O5 1 g/LSol–gel method50 mg/L MB (100 mL)Visible light~14% within 150 in~23% within 150 in[266]
Nb-TiO2 microspheres microspheres (anatase phase)-0.1 gUltrasonic spray pyrolysis method combined with impregnation method10 mg/L MB (100 mL)Visible light irradiation5% within 60 minNb-TiO2 microspheres show higher activity as compared to unmodified TiO2, 26% within 60 min[267]
Nb2O5/TiO2 nanoparticles (anatase phase)0.5 mol% Nb2O5/TiO250 mgCoprecipitation10 μmol/L MB (300 mL)UVP25 showed the highest photocatalytic activity within 1 h.Degradation rate for 0.5 mol% Nb2O5-TiO2 was slightly higher than the unmodified TiO2 within 1 h[268]
Nb-TiO2 nanoparticles (anatase phase)2 mol% Nb-TiO2-Hydrothermal method10 ppm MB (-)Solar light~40% within 100 min~40% within 100 min[262]
Nb–TiO2 films (mixture of anatase and rutile phases)0.58 mol% Nb–TiO2-Spin coating technology10 ppm MB (-)UV~10% (1.07 × 10−3 min−1) within 120 min~60% (4.97 × 10−3 min−1) within 120 min[269]
XC: carbon xerogel; 4-CP: 4-chlorophenol; MB: methyl blue.
Table 13. Summary of recent progress on W-TiO2 photocatalysts for organic pollutants degradation.
Table 13. Summary of recent progress on W-TiO2 photocatalysts for organic pollutants degradation.
Materials (TiO2 Phase Transition)OptimumCatalyst DosageSynthetic MethodsPollutantLight SourceUnmodifiedModifiedRef.
W-TiO2 nanoparticles (anatase phase)5% W-TiO20.5 gLiquid phase plasma10 mg/L DEP (600 mL)UV/Blue light1.22 × 10−2 min−1 and 4.72 × 10−2 min−1 for UV light and blue light, respectively within 180 min0.92 × 10−2 min−1 and 29.37 × 10−2 min−1 for UV light and blue light, respectively within 180 min[280]
W-TiO2 nanoparticles (anatase phase)2.5 wt% W-TiO22 g/LSol–gel method600 mg/L COD (150 mL)Fluorescent light-50% within 34 h[281]
W-TiO2/reduced graphene oxide nanoparticles (anatase phase)1% W-TiO20.01 gSol–gel method10 mg/L PNP (50 mL)UV30% within 180 min62% within 180 min[271]
W-TiO2 nanoparticles (anatase phase to stable rutile phase)TiO2/W50 ppm0.1 gHydrothermal method20 mg/L RhB, MB, MO (100 mL)Visible light~81% of MO within 15 min. 0.0031 min−1, 0.01612 min−1 and 0.1441 min−1 for RhB, MB and MO, respectivelyDegradation rate is 10 times the unmodified TiO2. 0.03181 min−1, 0.04148 min−1, 0.2730 min−1 for RhB, MB and MO, respectively.
~99.4% of MO within 15 min. Furthermore, 93.1 and 87.1% of MB and RhB were degraded, respectively, within 60 min, while 99.61% of MO within 20 min.
[282]
W-TiO2 nanoparticles (anatase phase)0.5 mol% W-TiO20.5 gSol–gel method50 mg/L CR (250 mL)Visible light1.25 min−1 for Degussa P25 within 180 min2.62 min−1 within 180 min[283]
WO3/TiO2 nanotube (anatase phase)-1 cm × 2 cmElectrochemical approach100 ppm Toluene (-)Xenon lamp-65% within 60 min[284]
Colloidal W- TiO2 nanocrystals (anatase phase)2% W–TiO21 g/LHydrothermal method10 mg/L Phenol (-)UV48.9% (0.1092 h−1) within 6 h80.0% (0.2694 h−1) within 6 h[285]
W-TiO2 layer (rutile to stable anatase phase)6 wt% W-TiO215 mm×10 mmPEOx method10 mg/L MB (100 mL)660-nm
Tungsten-halogen
55% within 180 min for pure TiO295% within 180 min[76]
W-TiO2 nanoparticles (anatase phase) 1 wt% W-TiO20.2 g/LEISA method0.156 mol L−1 4-CP (70 mL)UV32% within 100 min88% within 100 min[233]
WO3/TiO2 nanoparticles (anatase phase) 10% WO3/TiO20.05 gcombination of hydrothermal and calcination method10 mg/L MB (50 mL)Visible light40.7 and 55.3% using pure TiO2 and WO3, respectively for MB within 2 h.
41.6 and 54.8% using pure TiO2 and WO3, respectively for MET within 2 h.
87.8 and 67.1% for MB and MET, respectively within 2 h.
After three cycles, 10% WO3/TiO2 samples could still remove 63.2% of MET and 79.8% of MB.
[286]
W-TiO2 nanoparticles (anatase phase) -0.5 g/LSol–gel method30 mg/L SMZ (200 mL)Metal-halide lamps80% within 30 minThe extended reuse of the photocatalysts for five consecutive runs obtained an SMZ degradation of 97.7, 97.6, 96.2, 95.1, and 90.34% in the same order.[287]
W-TiO2 nanoparticles (anatase phase)1.5 wt% W-TiO2 -Sol–gel method10 mg/L Thymol (100 mL)Solar light9.65% (0.085 × 10−2 min−1) within 120 min48.76% (0.582 × 10−2 min−1) within 120 min[288]
W-TiO2 nanosheets (anatase phase)-64 m2Spray pyrolysis method1 mM oxalic acid 2.5 × 10−4 M NB (-)Sunlight42% within 320 min~83% within 180 min[289]
W-TiO2 coupons (rutile to stable anatase phase)6 wt% W-TiO215 mm × 10 mmPEOx method10 mg/L MB (100 mL)660-nm
Tungsten-halogen
55% within 180 min for pure TiO295% within 180 min[76]
PNP: p-nitrophenol; SMZ: sulfamethazine; MB: methyl blue; PEOx: plasma electrolytic oxidation; 4-CP: 4-chlorophenol; RhB: rhodamine B; CR: Congo red; MO: methyl orange; COD: chemical oxygen demand; DEP: diethyl phthalate.
Table 14. Summary of recent progress on Zn-TiO2 photocatalysts for organic pollutants degradation.
Table 14. Summary of recent progress on Zn-TiO2 photocatalysts for organic pollutants degradation.
Materials (TiO2 Phase Transition)OptimumCatalyst DosageSynthetic MethodsPollutants (Volume)Light SourceUnmodifiedModifiedRef.
Zn-TiO2 nanoparticles (anatase phase)1 wt% Zn-TiO20.1 gSingle step of sonochemical method10 mg/L RhB (100 mL)UV18% within 90 min35% within 90 min[290]
Zn-TiO2 nanoparticles (mixed anatase and rutile phases)0.01% Zn-TiO20.1 gSol–gel method20 mg/L MB (50 mL)Visible light89.69% within 60 min99.64% within 60 min[128]
Zn-TiO2 nanoparticles
(mixed anatase and brookite phases)
0.5 mol% Zn-TiO21 g/LSol–gel method10 mg/L PNP (100 mL)UV/Visible light 50 and 20% under UV/Visible and visible light, respectively for Degussa P25 within 8 h.75 and 32% for UV/Visible and visible light, respectively within 8 h[127]
Zn-TiO2 nanoparticles (anatase phase)5 wt% Zn-TiO20.007 gSol–gel method5 mg/L MB (50 mL)Visible light50% within 150 min28.75 and 50% within 35 min and 47 min, respectively[291]
Zn-TiO2 nanoparticles (anatase phase)0.2 wt% Zn-TiO2-Sol–gel method50 mg/L MO (50 mL)Visible light27% (0.45 min−1) and 3% (0.40 min−1) under UV irradiation and visible irradiation, respectively within 30 min62% (1.54 min−1) and 40% (0.92 min−1) under UV irradiation and visible irradiation, respectively within 30 min[113]
Zn-TiO2 nanoparticles (anatase phase)5% Zn-TiO20.1 gOil-in-water microemulsions, (O/W) microemulsion method30 mg/L Phenol (200 mL)UV84% within 5 h93% within 5 h[292]
MB: methylene blue; MO: methyl orange; PNP: p-nitrophenol; RhB: rhodamine B.
Table 15. Summary of recent progress on Au-TiO2 photocatalysts for organic pollutants degradation.
Table 15. Summary of recent progress on Au-TiO2 photocatalysts for organic pollutants degradation.
Materials (TiO2 Phase Transition)OptimumCatalyst DosageSynthetic MethodsPollutantLight SourceUnmodifiedModifiedRef.
Au-TiO2 nanoparticles (anatase phase)1% Au-TiO20.005 gPhotodeposition method10 mg/L MB, CBN, MNZ (10 mL)UV73.9 and 88.7% for CBN and MNZ, respectively35.7 and 46.8% for CBN and MNZ, respectively[304]
Au-P25 nanoparticles (mixed anatase rutile phase)1% Au-P250.005 gPrecipitation–deposition method10 mg/L BPA (10 mL)UV35% within 24 h76% within 24 h[305]
Au-TiO2 nanoparticles (mixed anatase rutile phase)1% Au-TiO20.005 gSol–gel method10 mg/L BPA (10 mL) UV35% within 24 h~57% within 24h[305]
Au-TiO2 nanoparticles (mixed anatase rutile phase)1% Au-TiO2 NPs0.005 gSol–gel method10 mg/L BPA (10 mL)UV35% within 24 h~56% within 24 h[305]
Au-meso-TiO2 nanoparticles (mixed anatase rutile phase) 1% Au-meso-TiO2 0.005 gSol–gel method10 mg/L BPA (10 mL)UV35% within 24 h37% within 24 h[305]
TiO2 hollow microspheres impregnated with biogenic Au nanoparticles (stable anatase phase)5% Au-TiO2 NPs0.005 gSol–gel method2.5 × 10−4 mol/L Phenol (10 mL)Visible light21% within 1 h95% within 1 h[306]
Au-TiO2 nanoparticles (anatase phase)0.154 wt% Au-TiO20.004 gSimple synthesis route3.125 × 10−5 mol/L MB (-)Visible LED light25% within 150 min97% within 150 min[307]
Au-TiO2 yolk-shell (anatase phase)0.14 at% Au-TiO20.005 gSeed-growth method300 ppm gaseous toluene (-)Visible light20% within 3 h57.3% within 3 h[308]
Au-TiO2 nanoparticles (anatase phase)--Sol–gel, wet chemical synthesis, hydrothermal, and plasma oxidative pyrolysis20 mg/L CR, MO (100 mL)UV-40.1 and 19.7 for MO and CR, respectively within 210 min.[309]
Ag-TiO2 nanoparticles (anatase phase)0.005 wt% Ag-TiO21.005 gSol–gel, wet chemical synthesis, hydrothermal, and plasma oxidative pyrolysis20 mg/L CR, MO (100 mL)UV-52.3 and 34.4% for MO and CR, respectively within 3.5 h[309]
Au-TiO2 nanoparticles (anatase phase)0.1% Au-TiO2 0.5 gHydrothermal method10 mg/L Resorcinol (500 mL)UVA72.36% (0.0044 min−1) within 5 h95.34% (0.0102 min−1) within 5 h (1.5 times of magnitude higher than pure TiO2 NPs while the rate constant was 2.5 times greater than that for pure TiO2). Furthermore, it shows 95% after five repeating experiments[310]
Au-TiO2 films (anatase phase)0.02 wt% Au/TiO2 meso-porous thin films-Sol–gel method5 mg/L TC (100 mL)UVA27% within2 hAfter 6 cycles, the percentage removal of TC is decreased from 58.51% to 57.89% only (i.e., a 0.62% decrease)[311]
Au-TiO2 plasmonic nanohybrids (anatase phase)-0.005 gWet chemical method10 mg/L MB, MO (100 mL)Sunlight25% within 20 min94% for MB, 85% for MO, and 87% for the mixture of MO and MB within 20 min (3.3 times greater than that of the unmodified TiO2)[312]
Au-TiO2 nanoparticles (Mixed anatase rutile phase)2 wt% Au-TiO21.3 g/LDeposition–precipitation method50 mg/L CFS (100 mL)UV/Visible light<5% within 2 h under visible light irradiation95 and 65% within 2 h under UV–visible and visible light irradiation, respectively[313]
Au-TiO2 photoanode on carbon cloth (anatase phase)0.1 wt% Au-TiO22 cm2Sol–gel method78.5 mg/L paracetamol (100 mL)PEC/Sunlight-62 and 66% under PEC process and sunlight radiation in solar PEC, respectively within 180 min[314]
Au-TiO2 film (anatase phase)3 at% Au-TiO2 film10 cm × 10 cmChemical spray pyrolysis technique1 mmol/L
benzoic acid (-)
UV37% within 400 min49% within 400 min[29]
CBN: carbofuran; MNZ: metronidazole; MO: methyl orange; CFS: ceftiofur sodium; MB: methylene blue; CR: Congo red; TC: tetracycline; BPA: bisphenol A.
Table 16. Summary of recent progress on Ag-TiO2 photocatalysts for organic pollutants degradation.
Table 16. Summary of recent progress on Ag-TiO2 photocatalysts for organic pollutants degradation.
Materials (TiO2 Phase Transition)OptimumCatalyst DosageSynthetic MethodsPollutant (Volume)Light SourceUnmodifiedModifiedRef.
Ag-TiO2 nanoparticles (anatase phase)5% Ag-TiO20.0113 mgSol–gel method8 mg/L 2-CP (300 mL)UV16% within 150 min30.4% within 150 min[340]
Ag-TiO2 thin films (anatase phase)5%Ag-TiO22.5 cm × 2.5 cmSol–gel method5 mg/L MO, MB (25 mL)Visible light40.10% for MB within 240 min98.86% for MB within 240 min, 96.34% for MO within 180 min. After ten consecutive cycles, approximately 98.85% MB dye were removed.[341]
Ag-TiO2
Films (anatase phase)
Ag-TiO2 films-Sol–gel method10 mg/L RB (-)Solar light/Visible light/UV45% within 70 min98, 78, and 50% under solar light, visible and UV within 70 min. After four consecutive cycles, approximately 96.5% RB dye was removed.[342]
Ag-TiO2
nanopowders (anatase phase)
10% Ag-TiO2 0.05 gSol–gel method0.03 mg/L MB (50 mL)UV96% within 2 h, 97% within 35 min97% within 35 min, 96% within 2 h[343]
Ag-TiO2 nanoparticles (anatase phase)10 wt% Ag-TiO20.3 g/LPolyol method1000 mg/L DCA (100 mL)UV65.4% within 480 min79.3% within 480 min[344]
Ag-TiO2 nanoparticles (anatase phase)10 wt% Ag-TiO20.05 gSol–gel method0.01 mmol/L MB (-)Visible light~30% within 50 hAlmost 100% within 50 h[317]
Ag-TiO2 nanoparticles (anatase phase)1.4 wt% Ag-TiO20.05 gPhotoreduction method10 mg/L 4-CP (50 mL)UV-After 8 times of reaction, the degradation effect of 4-CP decreased to 65%.[345]
Ag-TiO2 nanoparticles (anatase phase)4.0 mol% Ag-TiO20.18 gSol–gel method10 mg/L MB (180 mL)Visible light irradiation30% within 60 min96% within 60 min. After five consecutive cycles, approximately 89% MB dye was removed.[318]
Ag TiO2 nanopowders (anatase phase)Ag 3% TiO2 0.125 mol/LSol–gel method2.10−5 mol./L MB (-)UV0.08124 min−1 within 120 min0.11319 min−1 within 120 min[119]
Ag/TiO2 photoanode (anatase phase)--Photoreduction method30 mg/L RhB (250 mL)UV97.3% (0.0301 min−1) within 120 min99.5% (0.0451 min−1) within 120 min[346]
Ag/TiO2 nanoparticles (mixture of anatase and rutile phases)1.06 wt% Ag-TiO21.5 g/LPhotodeposition method 5 mg/L DXM (600 mL)UV/VIS2% within 240 min77.6 and 63.8% for UV and visible light irradiation, respectively[347]
Ag/TiO2 nanoparticles (mixture of anatase and rutile phases)0.03 wt% Ag-TiO21.0 g/LLP followed by wet impregnation and reduction methods125 µM MO (130 mL)UV6.54 × 10−3 min−1 and 0.19 × 10−3 min−1 for UV and visible light irradiation within 120 min, respectively28.74 × 10−3 min−1 and 16.78 × 10−3 min−1 for UV and visible light irradiation within 120 min, respectively[348]
Ag/TiO2 nanoparticles (anatase phase)0.25 wt% Ag-TiO21 g/dm3MWASG20 mg/L MO (100 mL)UV57% within 70 min99.5% within 70 min[349]
Ag-TiO2 nanoparticles (mixture of anatase and rutile phases)0.15 wt% Ag-TiO20.5 gSol–gel method8 mg/L MB (100 mL)Visible light40 and 42% for unmodified TiO2 and P25, respectively within 7 h~68% within 7 h[350]
Ag2+/TiO2 nanoparticles (anatase-brookite TiO2 nanoparticles with a spherical shape)2 mol% Ag2+/TiO21 g/LSol–gel method10 mg/L PNP (100 mL)UV/Visible light50 and 20% for Degussa P25 under UV/Visible light and visible light, respectively within 8 h60 and 31% under UV/Visible light and visible light, respectively within 8 h[127]
Ag-TiO2 nanopowders (anatase phase)3% Ag-TiO20.125 molL−1Sol–gel method20 mg/L MB (100 mL)UV0.08124 min−1 within 6 h0.11319 min−1 within 6 h[119]
Ag-TiO2 nanoparticles (anatase phase)10% Ag-TiO20.5 gSol–gel, mechanothermal decomposition method5 × 10−5 M MO (500 mL)UV/Solar69% within 60 min under UV, 8% within 80 min under solar irradiation98.9% within 60 min under UV, 99.3% within 80 min under solar irradiation[351]
Ag-TiO2
Nanoparticles (anatase phase)
1.75% Ag-TiO20.03 g/LNanosol/dip-coating method30 mg/L MB (50 mL)UV-74% within 2 h[217]
MWASG: microwave-assisted sol–gel method; RB: rose Bengal; DXM: dexamethasone; 4-CP: p-chlorophenol; DCA: dichloroacetic acid; 2CP: 2-chlorophenol; PNP: p-nitrophenol; LP: laser pyrolysis; MO: methyl orange; RhB: rhodamine B; MB: methylene blue; 4-CP: 4-chlorophenol.
Table 17. Summary of recent progress on Pt-TiO2 photocatalysts for organic pollutants degradation.
Table 17. Summary of recent progress on Pt-TiO2 photocatalysts for organic pollutants degradation.
Materials (TiO2 Phase Transition)OptimumCatalyst DosageSynthetic MethodsPollutant (Volume)Light SourceUnmodifiedModifiedRef.
Pt-TiO2 nanoparticles (mixed rutile and anatase phase)5 wt% Pt-TiO21 g/Lhydrothermal method10 mg/L RhB (100mL)Visible lightAmount of RhB
degraded in 90 min
reaction time 0.326 ± 0.011 × 10−9moles/min/mg.
99.5% within 90 min (4.5-fold better than pristine TiO2)[354]
Pt-TiO2 nanocrystals (mixed anatase and brookite phases)0.1wt% Pt-TiO21g/LSol–gel method10 mg/L PNP (10mL)Halogen lamp20 and 12% for TiO2 and P25, respectively within 24 h under visible light. 47 and 84% for TiO2 and P25, respectively within 24 h under UV/Visible light.~46% within 24 h under visible light.
~78% within 24 h under UV/Visible light.
[114]
Pt-TiO2 film (mixed anatase and brookite phases)0.1% Pt-TiO2size of 100 mm × 120 mmImmersion and reduction method100 ppb/500 ppb Ethenzamide (-)V-UV50 and 42% for 500 ppb and 100 ppb, respectively.94.52 and 100% for 500 ppb and 100 ppb, respectively. The rate constant (0.180 min−1) was 1.86 times as compared to unmodified nanoporous TiO2 film (0.097 min−1)[355]
Pt-TiO2 nanoparticles (mixed rutile and anatase phase)10 wt% Pt-TiO2 0.3 g/LPolyol method1000 mg/L DCA (100 mL)UV34.6% within 420 minAlmost 100% within 420 min[344]
Pt-P25 nanoparticles (mixed rutile and anatase phase)/0.5 g L−1Sol–gel method20 mg/L MG (250 mL)UV70% within 30 min for P25100% within 30 min[356]
Pt-MTiO2 nanoparticles (mixed rutile and anatase phase)1.5 wt% Pt-MTiO21.5 g/Lfacile synthesis procedureCIP (-)Visible light~5% within 120 min100% within 120 min. After five cycles, approximately 100% was still maintained.[357]
Pt/P25 and Pt/TiO2 nanoparticles (mix ed rutile and anatase phase)0.14 wt% Pt/TiO2 and Pt/P251.0 g/Llaser pyrolysis (LP) followed by wet impregnation and reduction methods;125 µM MO (-)UV/Visible light6.54 × 10−3 and 9.6 × 10−3 min−1 for pure P25 and TiO2 (LP), respectively in the presence of UV light.
0.19 × 10−3 and 1.8 × 10−3 min−1 for pure P25 and TiO2 (LP), respectively in the presence of VIS lamp
In the presence of UV lamp, the rate constant of Pt/P25 and Pt/TiO2 was 9.78 × 10−3 min−1 and 10.16 × 10−3 min−1.
In the presence of VIS lamp, the rate constant of Pt/P25 and Pt/TiO2 were 5.47 × 10−3 min−1 and 4.1 × 10−3 min−1.
[348]
Pt-TiO2 nanoparticles (mixed anatase and brookite phases)0.8 wt% Pt-TiO20.2 g/LSol–gel method6 mg/L MB(50 mL)UVC-57% within 60 min[358]
Pt-TiO2 nanoparticles (mixed anatase and brookite phases)0.5 mol% Pt-TiO21 g/LSol–gel method10 mg/L PNP (100 mL)UV/Visible light 50 and 20% for UV/Visible light and visible light, respectively for Degussa P25 within 8 h80 and 45% for UV/Visible light and visible light, respectively within 8 h[127]
Pt-TiO2 nanoparticles (mixed anatase and brookite phases)3 wt% Pt-HPT0.3 g/LHydrothermal method10 mg/L RhB (100 mL)UV/Visible light0.0013 min−1 for core(metal)-shell (TiO2) within 180 min63% (0.0053 min−1) within 180 min[359]
Pt-TiO2 nanoparticles (mixed anatase and brookite phases)10 wt% Pt-TiO2 0.3 g/LPolyol method1000 mg/L DCA (100 mL)UV34.6% within 420 min100% within 420 min[344]
CIP: ciprofloxacin; LP: laser pyrolysis; HPT: highly porous TiO2; DCA: dichloroacetic acid; MG: malachite green; MO: methyl orange; MB: p-nitrophenol; methylene blue; PNP: RhB: rhodamine B.
Table 18. Summary of recent progress on Ru-TiO2 photocatalysts for organic pollutants degradation.
Table 18. Summary of recent progress on Ru-TiO2 photocatalysts for organic pollutants degradation.
Materials (TiO2 Phase Transition)OptimumCatalyst DosageSynthetic MethodsPollutants (Volume)Light SourceUnmodifiedModifiedRef.
Ru-TiO2 nanotube arrays (anatase phase)0.15 wt%Ru-TiO2 -Anodic oxidation0.11 M TB
(20 mL)
UV/Visible light 84 and 26% under UV and visible light, respectively within 120 min82 and 28% under UV and visible light, respectively within 120 min.[364]
[Ru(4,4′-dicarboxy-2,2′-bipyridine)3]Cl2 (RuC)-TiO2 (anatase phase)-0.2 gHydrothermal method20 mg/L Bromophenol blue (500 mL)UV0.0018 min−1 within 120 min0.0038 min−1 within 120 min[365]
Ru-TiO2 nanoparticles (mixed anatase and brookite phases)0.2 wt% Ru-TiO20.002/LIncipient wet impregnation method100 mg/L 2-CP (500 mL)UV/Visible light 45 and 40% for UV and visible light, respectively within 60 min for pure TiO2.61 and 53% for UV and visible light, respectively within 60 min.[366]
RuO2/TiO2 composite nanotube arrays (anatase phase)0.0030 mol/L Ru-TiO2-Anodic oxidation method combined with dipping8 mg/L MB (12 mL)Fluorescent lamp37.8% within 2 h69% within 2 h[362]
Ru-TiO2 nanotube arrays (mixed anatase and brookite phases)0.16 wt% Ru-TiO20.01 mol L−1Electrochemical anodization4 ppm TB (-)UV0.0150 min−1; 81.4% color removal within 120 min1.33 higher activity than the unmodified TiO2[363]
2-CP: 2-chlorophenol; PNP: p-nitrophenol; MB: Methyl blue; TB: Terasil blue.
Table 19. Summary of recent progress on Pd-TiO2 photocatalysts for organic pollutants degradation.
Table 19. Summary of recent progress on Pd-TiO2 photocatalysts for organic pollutants degradation.
Materials (TiO2 Phase Transition)OptimumCatalyst DosageSynthetic MethodsPollutants (Volume)Light SourceUnmodifiedModifiedRef.
Pd-TiO2 nanoparticles (anatase phase)0.5 wt% Pd-TiO20.01 gSol–gel methods20 mg/L MB, MO (1 L)UV96.7% (0.024 min−1) and 82.5% (0.012 min−1) for MB and MO, respectively within 120 min.92.6% (0.018 min−1) and 99.4% (0.044 min−1) for MO and MB, respectively within 120 min. The catalytic performance remained nearly unchanged and the degradation of MB was maintained at 95.9% after 10 continuous cycles.[368]
Pd-TiO2 flower-like structures (anatase phase) 2 wt% Pd-TiO20.01 gUV light-induced method10 mg/L BPA (500 mL)UV/Visible light 100% within 45 min.100% within 45 min, 50% within 10 min under UV light; 100% within 240 min under visible light.
The rate constants were 2.92, and 3.88 times higher than the P25 TiO2. Furthermore, they were 1.65 and 1.91 times higher than those of unmodified TiO2 under visible and UV lights, respectively.
[369]
Pd-TiO2 nanoparticles (mixed anatase and brookite phases)1.3 wt% Pd-TiO20.01 g/LPhotodeposition20 mg/L MB (100 m)Xenon Lamp0.008 min−1 within 45 min.0.023 min−1 within 45 min.[370]
Pd-TiO2 nanoparticles (anatase phase)1 wt% Pd-TiO21 g/LHydrothermal5 mM Oxalic acid (100 mL)Fluorescent lamps50, 54, and 87% for bare Aldrich rutile, bare P25 and P25 (cubical), respectively within 2 h,96, 70, 72% for AA-Pd(spherical), AA-Pd(cubical), P25-Pd(spherical), respectively within 120 min. ~100 for AR-Pd(cubical) within 100 min[371]
Pd-TiO2 nanoparticles (anatase phase)1 wt% Pd-TiO21 g/LHydrothermal0.5 mM PhenolFluorescent lamps87, 63, 41% for P25, AA and bare Aldrich rutile, respectively within 2 h72, 66, 98, 90, 75, and 71% for P25-Pd (spherical), P25-Pd(cubical), AA-Pd(spherical), AA-Pd(cubical), AR-Pd(spherical)), and AR-Pd(cubical), respectively within 2 h.[371]
AA: Aldrich anatase; AR: Aldrich rutile; P25: Evonik AEROXIDE P25; BPA: bisphenol A; MB: methyl blue; MO: methyl orange.
Table 20. Summary of recent progress on Y-TiO2 photocatalysts for organic pollutants degradation.
Table 20. Summary of recent progress on Y-TiO2 photocatalysts for organic pollutants degradation.
Materials (TiO2 Phase Transition)OptimumCatalyst DosageSynthetic MethodsPollutants (Volume)Light SourceUnmodifiedModifiedRef.
Y-TiO2 nanoparticles (Anatase phase)-4 cm2Electrochemical method0.21 mM Toluene
(200 μL)
Visible light9% within 60 min21% within 60 min[374]
Y3+/TiO2
nanoparticles (Anatase phase)
0.25 mol% Y3+/TiO20.125 gHydrothermal methods0.21 mM Phenol (100 mL)UV/Visible light2.40 h−1 within 60 min3.85 h−1 within 60 min[375]
Y3+/TiO2 nanoparticles (Anatase phase)0.25 mol% Y3+/TiO20.125 gSol–gel methods0.21 mM Phenol (-)UV/Visible light2.40 h−1 within 60 min1.43 h−1 within 60 min[375]
Table 21. Summary of recent progress on lanthanides modified TiO2 photocatalysts for organic pollutants degradation.
Table 21. Summary of recent progress on lanthanides modified TiO2 photocatalysts for organic pollutants degradation.
Material (TiO2 Phase Transition)OptimumCatalyst DosageSynthetic MethodsPollutants (Volume)Light sourceUnmodifiedModifiedRef.
Er-TiO2 mesoporous spheres (anatase phase)1 mM of Er-TiO21 gSolvothermal method10 ppm
RhB (50 mL)
Visible light84.27% (0.0428 min−1) within 60 min98.78% (0.1544 min−1) within 60 min[389]
Ce-TiO2 nanoparticles (mixture of anatase and rutile phases)1% Ce-TiO20.5 g/LFacile EDTA-citrate method 10 mg/L CIP, NOR (200 mL)Sunlight69.29 and 75% for CIP and NOR, respectively within 180 min90–93% for both CIP and NOR within180 min[390]
Ce-TiO2 nanoparticles (anatase phase)0.5% Ce-TiO20.5 gFacile hydrothermal methodMEK (200 mL)Visible light-76% for 0.5% Ce-TiO2[391]
Ce-TiO2 nanoparticles (anatase phase)0.5% Ce-TiO20.5 gFacile hydrothermal methodMEK (200 mL)UV-62% for 0.5% Ce-TiO2[391]
Ce-TiO2 nanoparticles (anatase phase)0.41% Ce-TiO210 mg/cm2Sol–gel method25 mg/L DB15Visible light3.3% within 1 h33% within 1 h[216]
Ce-TiO2 mesoporous (mixture of anatase and rutile phases)2.5% Ce-TiO20.001 gEISA approach20 mg/L Phenol (100 mL)UV-95–99% within 4 h[392]
Ce-TiO2 nanocrystalline films (anatase phase)2.5% Ce-TiO2-EISA approach20 mg/L Phenol (100 mL)Solar light99% within 4 h for the mesoporous TiO255-68% within 4 h[392]
Ce-TiO2 nanocrystalline films (anatase phase)10% Ce-TiO20.05 gSol–gel method2.5 × 10−5 M BB-41 (80 mL)UV16.8 × 10−3 min−1 within 180 min15.9 × 10−3 min−1 within 180 min[393]
Ce-TiO2 nanocrystalline films (anatase phase)12% Ce-TiO20.05 gSol–gel method2.5 × 10−5 M BB-41 (80 mL)Visible light11.2 × 10−3 min−1 within 180 min16.1 × 10−3 min−1 within 180 min[393]
CeO2/TiO2 core-shell nanoparticles (anatase phase)-0.04 gHydrothermal route assisted with the Stöber method1 × 10−5
M RhB (40 mL)
Xe arc-lamp0.004 min−1 for CeO2 nanocubes and 0.003 min−1 for unmodified TiO2 within 120 min0.012 min−1 within 120 min[394]
CeO2/TiO2 nanoparticles (mixture of anatase and rutile phases) 0.25% CeO2/TiO20.1 gSol–gel method40 ppm MB (100 mL)Visible light96.43% within 150 min97.86% within 150 min[395]
Ho-TiO2 nanotubes (anatase phase)-4 cm2Electrochemical method0.21 mM
Toluene (200 μL)
Visible light9% within 60 min30% within 60 min[374]
Er-TiO2 nanotubes (anatase phase)-4 cm2Electrochemical method0.21 mM
Toluene (200 μL)
Visible light9% within 60 min22% within 60 min[374]
Gd-TiO2 nanotubes (anatase phase)-/4 cm2Electrochemical method0.21 mM
Toluene (200 μL)
Visible light9% within 60 min28% within 60 min[374]
Tb-TiO2 nanotubes (anatase phase)-4 cm2Electrochemical method0.21 mM
Toluene (200 μL)
Visible light9% within 60 min28% within 60 min[374]
Nd3+/TiO2 nanosphere (anatase phase)1.0 mol% Nd3+/TiO2 0.05 gtemplate-free method (recombined coprecipitation with hydrothermal method20 mg/L MB (50 mL)Visible light91.83% (0.46 h−1) within 120 min99.14% (2.3 h−1) within 120 min[396]
Nd3+/TiO2 nanosphere (anatase phase)1.0 mol% Nd3+/TiO2 0.05 gtemplate-free method (recombined coprecipitation with hydrothermal method20 mg/L MB (50 mL)Sunlight60.09% of dyes degraded without catalystalmost completely degradation within 80 min. [396]
Nd-TiO2 nanoparticles (anatase phase)2.0 mM Nd-TiO20.02 gSol–gel method40 mg/L CR (100 mL)UV64% within 30 min86% within 30 min, the photocatalytic efficiency after utilizing five times was more than 92%.[397]
Nd-TiO2 nanoparticles (anatase phase)2.0 mM Nd- TiO20.02 gSol–gel method40 mg/L MB (100 mL)UV74% within 45 min92% within 45 min, the photocatalytic efficiency after utilizing five times was more than 92%.[397]
Yb3+/TiO2 nanoparticles (anatase phase)1% Yb3+/TiO20.05 gSol–gel method0.21 mM phenol (5 mL)Visible light45% within 180 min89% within 180 min[398]
Er3+/TiO2 nanoparticles (anatase phase)0.5% Er3+/TiO20.05 gSol–gel method0.21 mM Phenol (5 mL)Visible light-75% within 180 min[398]
Er3+/TiO2 nanoparticles (anatase phase)2 mol% Er3+/TiO20.025 gElectrospinning10 mg/L MO (25 mL)Solar light 78% within 60 min43% within 60 min[399]
Er3+/TiO2 nanoparticles (anatase phase)0.5 mo% Er3+/TiO20.05 gElectrospinning10−5 mol/L RhB (25 mL)Solar light 71.3% within 10 h91.3% within 10 h[399]
Er3+/TiO2 nanoparticles (anatase phase)0.5 mol% Er3+/TiO20.05 gElectrospinning20 mg/L Phenol (25 mL)Solar light 23.9% within 72 h46.1% within 72 h[399]
Yb-TiO2 nanoparticles (anatase phase)10% Yb-TiO21 g/LHydrothermal process10 ppm
Phenol (100 mL)
Solar light ~Initial Rate with 31.9 mol/L/h~Initial Rate with 54 mol/L/h[400]
Er-TiO2 nanoparticles (anatase phase)2% Er-TiO21 g/LHydrothermal process10 ppm
Phenol (100 mL)
Solar light ~Initial Rate with 31.9 mol/L/h~Initial Rate with 144 mol/L/h[400]
Gd-TiO2 nanoparticles (anatase phase)1.8 at% Gd- TiO20.1 gSol–gel method
sintering at 550 °C
20 mg/L CR (100 mL)UV42% within 1 h76% within 1 h[401]
Gd-TiO2 nanoparticles (anatase phase)1.8 at% Gd- TiO20.1 gSol–gel method
sintering at 700 °C
20 mg/L CR (100 mL)UV35% within 1 h84% within 1 h[401]
Sm-TiO2 nanoparticles (anatase phase)1.8 at% Sm-TiO20.1 gSol–gel method
sintering at 700 °C
20 mg/L CR (100 mL)UV35% within 1 h74% within 1 h[401]
Sm-TiO2 nanoparticles (anatase phase)1.8 at% Sm-TiO20.1 gSol–gel method
sintering at 550 °C
20 mg/L CR (100 mL)UV42% within 1 h70% within 1 h[401]
Er-TiO2 nanotube arrays (anatase phase)10 wt% Er-TiO2-Electrochemical anodization0.5 mM Toluene (8 mL)Visible light-2.85 × 10−3 μmol·min−1 within 60 min[402]
Ho-TiO2 nanotube arrays (anatase phase)10 wt% Ho-TiO2-Electrochemical anodization0.5 mM Toluene (8 mL)Visible light-2.87 × 10−3 μmol·min−1 within 60 min[402]
Gd-TiO2 nanoparticles (anatase phase)0.3% Gd-TiO20.01 gSol–gel method10 mg/L RhB (10 mL)Visible light75% within 240 min93% within 240 min[403]
La-TiO2 nanoparticles (mixture of anatase and rutile phases)0.05 La-TiO21 g/LUltrasound-assisted wet impregnation method10 mg/L MB (1 L)UV0.1372 ± 0.0038 min−1 and 0.1332 ± 0.0051 min−1 for pristine TiO2 and P25-TiO2, respectively within 30 min0.1528 ± 0.0017 min−1 within 30 min[404]
Gd-TiO2 nanoparticles (anatase phase)3% Gd3+/TiO20.05 gImpregnation method20 mg/L RhB (60 mL)Mercury lamp75% within 50 min81–96% within 50 min[405]
Ce3+/TiO2 nanoparticles (mixture of anatase and rutile phases)1 mol% Ce3+/TiO20.2 g/LCombustion synthesis method20 mg/L MB (100 mL)Visible light 41% within 120 min72% within 120 min[406]
La-TiO2 nanoparticles (anatase phase)1.0% La–TiO20.05 gSol–gel method10 mg/L RhB (100 mL)562-nm Xenon Lamp21.56% within 300 min11.09% within 300 min[74]
Ce-TiO2 nanoparticles (anatase phase)1.0% Ce–TiO20.05 gSol–gel method10 mg/L RhB (100 mL)562-nm Xenon Lamp21.56% within 300 min83.43% within 300 min[74]
Eu-TiO2 nanopowders (anatase phase)3% Eu-TiO20.125 molL−1Sol–gel method20 mg/L MB (250 mL)Xenon Lamp0.08124 min−1 within 6 h0.03719 min−1 within 6 h[119]
Gd-TiO2 nanopowders (anatase phase)1 wt% Gd-TiO20.1 gSol–gel method10 mg/L MO (100 mL)Visible light31% within 180 min78% within 180 min[205]
Gd-TiO2 nanopowders (anatase phase)1 wt% Gd-TiO20.1 gSol–gel method10 mg/L 4-CP (100 mL)Visible light28% within 180 min69% within 180 min[205]
Eu-TiO2 nanoparticles (anatase phase)10% Eu-TiO20.3 gLPP process50 mg/L ASA (250 mL)Blue light0.685 × 10−3 min−1 within 24 h1.475 × 10−3 min−1 within 24 h[407]
Eu-TiO2 nanoparticles (anatase phase)10% Eu-TiO20.3 gLPP process50 ppm ASA (250 mL)UV9.37 × 10−3 min−1 within 24 h10.65 × 10−3 min−1 within 24 h[407]
at. %: atomic %; NOR: norfloxacin; MEK: methyl ethyl ketone; EE2: 17-α-ethinylestradiol; DB15; direct blue 15; BB-41: basic blue 41; RhB: rhodamine B; LPP: liquid phase plasma; MB: methylene blue; ASA: acetylsalicylic acid; RB: reactive blue; Cr: Congo red; MO: methyl orange; CIP: ciprofloxacin; 4-CP: 4-chlorophenol; EISA: evaporation-induced self-assembly.
Table 22. Summary of recent progress on other metals modified TiO2 photocatalysts for organic pollutants degradation.
Table 22. Summary of recent progress on other metals modified TiO2 photocatalysts for organic pollutants degradation.
Materials (TiO2 Phase Transition)OptimumCatalyst DosageSynthetic MethodsPollutants (Volume)Light SourceUnmodifiedModifiedRef.
Ga3+-TiO2 nanoparticles (anatase phase)0.5 wt% Ga3+-TiO20.1 gUltrasonic irradiation20 mg/L MB (100 mL)Solar light70.6% (0.0096 min−1) in 150 min86.4% (0.021 min−1) in 150 min[418]
In2O3-TiO2 nanopowders (anatase phase)5 mol% In2O3/TiO20.5 gSol–gel method200 mg/L CR (-)UV0.15 h−1 in 4 h0.86 h−1 in 4 h[426]
In2O3-TiO2 nanorods (anatase phase)0.4 wt% In2O3/TiO21 × 1.5 cm2Hydrothermal method 10 uM MO (-)Sunlight86% in 6 h94% in 6 h[427]
In2O3-TiO2 nanorods (anatase phase)0.4 wt% In2O3/TiO21 × 1.5 cm2Hydrothermal method 2.5 uM BPA (-)Sunlight65% in 6 h68% in 6 h[427]
In2O3-TiO2 nanoparticles (mixture of anatase and rutile phases)0.15 mol% In2O3/TiO20.5 gSol–gel method8 mg/L MB (100 mL)high-voltage Mercury lamp40 and 42% in 7 h for pure TiO2 and P25, respectively.~80% in 7 h[350]
TiO2-In2O3 porous structure and spherical morphology (anatase phase)1 mol% In2O3/TiO20.02 gUAS assisted method10 ppm
MO (20 mL)
Visible light26% in 5 h for pure In2O3~98% within 5 h[411]
TiO2-In2O3 porous structure and spherical morphology (anatase phase)1 mol% In2O3/TiO20.02 gUAS assisted method10 ppm RhB (20 mL)Visible light18% in 3 h for pure In2O395% within 3 h[411]
Ga3+-TiO2 nanoparticles (anatase phase)0.5 wt% Ga3+-TiO2-Ultrasonic irradiation20 ppm Phenol (100 mL)Solar0.0052 min−1 in 180 min0.021 min−1 and ~4 times higher activity than unmodified TiO2[418]
Ga-TiO2/rGO nanoparticles (mixed anatase and rutile phases)1.0 wt% Ga-TiO2/rGO0.01 gSimple synthesis process10 mg/L RB dye (-)Visible light0.0015 min−1 in 180 min0.0029 min−1 in 180 min[428]
Ga2O3-TiO2 nanoparticles (anatase phase)0.1% Ga2O3-TiO20.05 gSol–gel method0.08 mmol/L Imazapyr (100 mL)UV19% in 180 min for mesoporous Ga2O3.98% within 180 min (10 times higher than the mesoporous Ga2O3).
The photodegradation efficiency of imazapyr continues to maintain over 95% after five cycles.
[429]
Al-TiO2 nanoparticles (anatase to rutile phase)0.25% Al-TiO22 g/LImpregnation method10−6 M
MB (100 mL)
LED lamp75% within 60 min for P-25-[430]
Al-TiO2 nanoparticles (anatase to rutile)0.25% Al-TiO20.5 g/LImpregnation method10−6 M
MB (100 mL)
LED lamp-80% in 60 min[430]
Al-TiO2 nanoparticles (anatase to rutile)0.25% Al-TiO21 g/LImpregnation methodMB (100 mL)LED lamp-85% in 60 min[430]
Sn-TiO2 nanoparticles (anatase to rutile)1% Sn-TiO20.2 gSol–gel method600 ppm DES (-)Solar light38% in 90 min,48% in 180 min~100% in 90 min, 75% in 150 min[431]
Sn-TiO2 nanoparticles (anatase to rutile phase)5 mol% Sn-TiO20.005 gSol–gel method10 mg/L MB (25 mL)Solar lamp98% within 60 min for TiO2-NTs, 94% within 60 min for TiO2–P2598% (reaction rate of 0.1215 min−1 which is 1.8 and 2.7 times as much as those of TiO2-NTs and TiO2P, respectively) within 30 min[432]
Sn-TiO2 nanoparticles (anatase to rutile phase)1 mol% Sn-TiO20.3 gSol−gel method10 mg/L RhB (300 mL)UV46.2% within 180 min99.5% (0.02732 min−1) within 180 min[433]
Sn-TiO2 Hollow Spheres1 mol% Sn-TiO20.025 gSol−gel method templated by polystyrene spheres0.00385 g/L MO (50 mL)UV~55% within 240 min~72% within 240 min[434]
SnO2-TiO2 nanoparticles (anatase to rutile phase)1% Sn-TiO21 g/LSonication–impregnation method10 mg/L TC (-)Visible light-95% within 15 min, 81–95% within 20 min[435]
Sn-TiO2 nanoparticles (anatase phase)0.20 ± 0.03% Sn-TiO2-Washcoating methodERY (1 L)UV-A~18% within 240 min67% within 240 min, 50% within 2280 min[436]
Sn-TiO2 nanoparticles5% Sn-TiO20.8 g/LHydrothermal method10 mg/L DCF (-)UV~62% within 300 min89% within 300 min 53% after 4 cycles (1200 min)[437]
Sn-TiO2 nanoparticles1 mol% Sn-TiO20.2 gSol−gel method40 ppm 2,4-DCA (-)UV77% within 180 min93% (16.8 × 10−3 min−1) within 180 min[438]
Al3+/TiO2 nanoparticles
(mixed anatase and brookite phases)
0.5 mol% Al3+/TiO21 g/LSol−gel method10−4 M PNP (100 mL)UV/Visible light50 and 20% under UV/Visible and visible light, respectively for Degussa P25 within 8 h,64 and 24% under UV/Visible and visible light, respectively within 8 h[127]
PNP: p-nitrophenol; CR: Congo red; UAS: Ultrasonic aerosol spray; MB: methylene blue; RhB: rhodamine B; DCA: dichloroacetic acid; DCF: diclofenac; ERY: erythromycin; TC: tetracycline; MO: methyl orange; BPA: bisphenol A; RB: reactive blue.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Jiang, D.; Otitoju, T.A.; Ouyang, Y.; Shoparwe, N.F.; Wang, S.; Zhang, A.; Li, S. A Review on Metal Ions Modified TiO2 for Photocatalytic Degradation of Organic Pollutants. Catalysts 2021, 11, 1039. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11091039

AMA Style

Jiang D, Otitoju TA, Ouyang Y, Shoparwe NF, Wang S, Zhang A, Li S. A Review on Metal Ions Modified TiO2 for Photocatalytic Degradation of Organic Pollutants. Catalysts. 2021; 11(9):1039. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11091039

Chicago/Turabian Style

Jiang, Dafu, Tunmise Ayode Otitoju, Yuanyuan Ouyang, Noor Fazliani Shoparwe, Song Wang, Ailing Zhang, and Sanxi Li. 2021. "A Review on Metal Ions Modified TiO2 for Photocatalytic Degradation of Organic Pollutants" Catalysts 11, no. 9: 1039. https://0-doi-org.brum.beds.ac.uk/10.3390/catal11091039

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