Open access peer-reviewed chapter

TiO2-Low Band Gap Semiconductor Heterostructures for Water Treatment Using Sunlight-Driven Photocatalysis

Written By

Raquel Del Angel, Juan C. Durán-Álvarez and Rodolfo Zanella

Reviewed: 15 March 2018 Published: 05 November 2018

DOI: 10.5772/intechopen.76501

From the Edited Volume

Titanium Dioxide - Material for a Sustainable Environment

Edited by Dongfang Yang

Chapter metrics overview

2,187 Chapter Downloads

View Full Metrics

Abstract

Heterogeneous photocatalysis is a promising advanced oxidation process for water purification, given its potential to fully oxidize organic pollutants and to inactivate microorganisms. Due to its versatility and high performance in a broad range of conditions, titanium dioxide (TiO2)-based photocatalysis has been systematically used at laboratory scale to treat water of different quality. Even though TiO2 is an exceptional photocatalyst, its broad band gap value (3.2 eV) makes necessary the use of UV light to achieve the photoactivation. This results in the underutilization of the material in sunlight-driven photocatalysis schemes. In order to overcome this handicap, the synthesis of heterostructures using low band gap semiconductors coupled with TiO2 has brought exceptional materials for visible light-driven photocatalysis. In this chapter, the fundamentals of the synthesis and photoactivation of TiO2-low band gap semiconductor heterostructures are explored. The mechanisms leading to the increase of the photocatalytic activity of such heterostructures are described. A summary of the available data on the photocatalytic performance of TiO2-based heterostructures is presented, in terms of degradation of organic pollutants in water using visible light and sunlight. A comparison of the depuration performance of powdered and thin film heterostructures is given at the end of the chapter.

Keywords

  • composites
  • heterojunctions
  • metallic oxides
  • organic pollutants
  • semiconductors
  • water treatment

1. Introduction

Water and wastewater treatment brings new challenges due to the occurrence of new and refractory contaminants produced by anthropogenic activities [1, 2]. In the past few years, water depuration aimed to remove particles, the bulk of organic matter and inactivate microorganisms. However, nowadays, the degradation not only of organic pollutants at trace levels but its precursors and by-products is especially pursued in drinking water treatment systems. For instance, the biological and chemical degradation of fluoroquinolone antibiotics results in the generation of some by-products displaying antibiotic residual activity. Biological and abiotic degradation of personal care products, such as triclosan and triclocarban, leads to the emergence of polychlorinated biphenyls and dioxins. Some bacterial residues in surface water bodies, such as microcystin and geosmin, may impact in the organoleptic properties of drinking water as well as express toxicity, while some iodinated pharmaceuticals may be precursors of trihalomethanes. Under this scenario, the most advanced water treatment systems should aim to completely mineralize the organic pollutants, in order to take the risks of water contamination to the minimum. The complete mineralization of organic pollutants can be warranted by few processes, such as photocatalysis [3, 4, 5, 6, 7]. This process is based on the generation of highly reactive OH radicals, which are able to fully oxidize organic molecules. Heterogeneous photocatalysis process has shown to be efficient in the degradation of organic pollutants in water, avoiding the transport and use of potentially hazardous materials, such as acids or H2O2, while the catalyst can be recovered and reused in several cycles [8, 9].

TiO2 is a widely used semiconductor in heterogeneous photocatalysis due to its high activity and stability [10, 11, 12]. This harmless material is currently used not only for water treatment but in food preparation and disinfection of surgical equipment. Anatase is the most photoactive phase of TiO2, followed by rutile; however, the wide band gap value of both phases (3.0–3.2 eV) results in the activation of these materials under UV-A light irradiation [13]. In order to take the photocatalysis process toward sustainability, it is necessary to develop materials with high photo-response under sunlight irradiation. Given that sunlight comprises only 4–5% of UV-A light, the need to find nanostructures capable to absorb visible light—which composes 50% of the sunlight spectrum [14]—becomes imperative. A growing number of modifications of TiO2 at nanometric scale have been performed in order to achieve the complete photoactivation of this semiconductor under visible light irradiation. Doping with nonmetal atoms has given partially positive results, since the redshift of the absorption edge of TiO2 increases in turn the recombination rate of the hole-electron pair. The development of heterostructures based on TiO2 coupled to low band gap semiconductors could be an efficient approach to improve the photocatalytic conversion of contaminants in water [12, 15, 16]. This chapter explores the fundamentals of the synthesis and photoactivation of the TiO2-low band gap semiconductor heterostructures, and presents some of the reported data on the photocatalytic activity of these materials, in order to bring light on their potential use for water purification at a higher scale.

Advertisement

2. Fundamentals on the formation and photoexcitation of TiO2-based heterostructures

A heterostructure or junction is defined as the interfacial union of two or more components. In photocatalysis, heterounions are formed to improve the efficiency of a semiconductor either by redshifting the light absorption or through the decrease of the recombination of the hole-electron pairs [12, 17, 18]. Heterostructures are commonly built up by combining a specific semiconductor with one or more materials, such as metals, semiconductors or organic molecules for example, azo dyes or polymers.

Semiconductor/metal junctions are heterostructures based on the deposition of metallic nanoparticles on crystalline semiconductors. Such union is able to hinder the electron/hole recombination through the sequestration of the photoelectrons from the conduction band of the semiconductor to the surface of the metallic nanoparticles [12] (Figure 1). The differences in the Fermi level reached in the semiconductor/metal junction triggers the sequestration of photoelectrons; although other factors, such as the higher work function and electronegativity of the metal, favor the transference of the charge carriers [16, 19, 20]. The defects created in the semiconductor/conductor junction are proposed as the main route of the electron transfer from one material to the other. The Schottky barrier created in the junction impedes the return of the charge carriers to the conduction band of the semiconductor. Typically, low loadings of metallic nanoparticles tend to be the optimal for achieving the highest potential of the electron trap process. However, when this optimal loading is surpassed, the electron traps are converted into charge carriers recombination sites, leading to the dramatic drop of the photocatalytic performance [21, 22, 23]. Even when the electron trap effect is unable to significantly redshift the light absorption of the heterostructure, some photocatalytic activity can be achieved by the surface plasmon resonance (SPR) effect, which is expressed by the oscillation of the electrons in the conduction band of the metallic nanoparticles when electromagnetic excitation at a determined wavelength is provided [24, 25, 26]. Some extent of photocatalytic activity has been attributed to SPR effect, although this cannot be considered as determinant in the photodegradation of pollutants in water when compared with that achieved under UV light irradiation.

Figure 1.

Photoactivation and charge carriers transfer in a semiconductor/conductor heterounion.

The synthesis of semiconductor/semiconductor nanocomposites provides an efficient way to redshift the photoactivation of TiO2-based materials. When these materials are exposed to visible light, the low band gap semiconductor is activated, producing the hole–electron pair. Then, the photoelectrons migrate from the surface of the low band gap semiconductor to the conduction band of TiO2, as an effect of the difference in the redox potential between both semiconductors (Figure 2). The process leads to the drop of the hole-electron pair recombination rate, resulting in the oxidation process in the low band gap semiconductor, and the reduction process on the TiO2 surface [17, 27].

Figure 2.

Photoactivation and charge carriers transfer in a TiO2-low band gap semiconductor heterounion when photoactivation with visible light occurs.

When the composite is photoactivated under UV light irradiation, the hole-electron pairs are produced in the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) of each semiconductor. Charge carriers are then transported and accumulated in the HOMO and LUMO of one of the semiconductors, as a function of gradient in the potential of the bands (Figure 3). For this kind of schemes, p-n heterostructures have shown the best results in the separation of the photo-formed charge carriers. However, in some cases, decreasing of the redox potential of charge carriers can occur, depending of the HOMO and LUMO positions of the semiconductors in the heterostructure, leading to the decrement in the photocatalytic performance [15, 17, 18].

Figure 3.

Photoactivation and charge carriers transfer in a TiO2-low band gap semiconductor heterounion when photoactivation with UV light occurs.

In some advanced approaches, metallic nanoparticles are settled in the heterounion of the two semiconductors. Noble metal nanoparticles can act as electron mediators, transporting the charge carriers from one semiconductor to another, increasing with this the electron trap effect. Heterostructures based on p-n semiconductors have shown a notable increase in the charge separation when noble metal nanoparticles are added in the semiconductor heterounion. These materials, known as all solid-state Z schemes, provide highly reductive photoelectrons and highly oxidative photo-holes by the process shown in Figure 4 [28].

Figure 4.

Photoactivation and charge carriers transfer in a TiO2 based all solid Z scheme when photoactivation when UV light activation occurs.

Photoelectrons are transported from semiconductor II to semiconductor I via the metallic nanoparticles. At the same time, photoelectrons are formed in semiconductor I and photoholes in the HOMO of semiconductor I are recombined with photoelectrons coming from the LUMO of semiconductor II. This kind of schemes impedes the recombination of the charge carriers with the highest oxidative and reductive potential, increasing not only the photocatalytic performance of the components, but bringing chances to photodegrade more recalcitrant pollutants because of the increment of the oxidative potential of the charge carriers.

When TiO2 nanoparticles are deposited on graphene sheets, the semiconductor/polymer junction displays higher photocatalytic activity due to the transfer of the charge carriers from the semiconductor to the polymer, as shown in Figure 5b. Even when the lifetime of the charge carriers is increased, visible light-driven activity is not improved; however, graphene materials can be used as an excellent electron mediator in all solid Z schemes. On the other hand, organic molecules, which are able generate the triplet state under visible light irradiation, can act as sensitizers when deposited on the TiO2 surface (Figure 5a). In this case, electrons are injected from the sensitized molecule to the LUMO of the semiconductor, triggering the photocatalytic process [29].

Figure 5.

Photoactivation and charge carriers transfer in a TiO2-sensitized materials under visible light activation (a) and in TiO2-graphene heterostructures when UV light occurs (b).

2.1. Classification of the semiconductors/semiconductor heterostructures

Semiconductor/semiconductor junctions can be classified depending on either the type of semiconductors that are being coupled or by the band structure they present. Considering the type of semiconductors, the semiconductor/semiconductor heterostructures can be classified either as anisotype—a p-n junction—, or isotype heterojunctions, in which both coupled semiconductors are n-type (n-n junction) or p-type (p–p junction). Some examples of n-type semiconductors include TiO2, WO3, ZnO, Fe2O3, CeO2, AgI, BiVO4, CdS, CdSe, Bi2WO6, and ZnSe, while for p-type semiconductors Bi2O3, V2O5, BiOI, BiOBr, BiOCl, CuO, Cu2O can be mentioned [18]. When semiconductor/semiconductor heterostructures are classified depending on the band position of the components, three main groups can be mentioned. Figure 6 shows the three possible band alignments in semiconductor/semiconductor heterostructures.

Figure 6.

Classification of semiconductor/semiconductor heterostructures based on the band alignment of the components.

In a type I heterostructure, also known as straddling alignment, the band gap value of the semiconductor B is smaller than that of semiconductor A (Figure 6). In this case, the potential of valence band of the semiconductor B is located at a higher position than that of semiconductor A, while conduction band of semiconductor A displays a lower potential than that of the conduction band of semiconductor B. In this scheme, electrons and holes are transferred from A to B, resulting in the accumulation of charge carriers in B [18]; this facilitates in turn the recombination of the charge carriers and decreases the photocatalytic activity. Obregon et al. [30] reported the formation of a type I heterojunction using monoclinic BiVO4 and TiO2. The 1 wt. % m-BiVO4/TiO2 nanocomposite was prepared by a simple impregnation method and was tested in the degradation of phenol. This system has been tested by other authors using different approaches and modification [31, 32, 33]. Other semiconductors that have been coupled to TiO2 to form a type I heterojunction [18] are WO3 [34, 35], Fe2O3 [36], MoS2 [37] and BiOI [36].

Type II heterostructure, or staggered lineup, represents the most efficient configuration for the charge carriers transfer. In this case, photo-generated electrons will be transferred from semiconductor B to semiconductor A, while the holes are moving from semiconductor A to semiconductor B. A more efficient charge separation is achieved, having both photoelectrons and photo-holes distributed in the two semiconductors. A large number of type II heterojunctions have been synthesized with TiO2, by the union with CdS [38], CdSe [39, 40], Bi2S3 [41], WS2 [42], V2O5 [43], Bi2O3 [44], CuO [45], Cu2O [45, 46], and CeO2 [47].

Type III heterojunctions are similar in structure to the type II composites, but a wider difference in the position of bands between semiconductor A and B is observed. This type of heterojunctions is also known as broke-gap junction, which is highly recommended to the construction of all solid Z schemes. For instance, Heng et al. [48] prepared a type III heterojunction by incorporating H3PW12O40, TiO2 and In2S3. The authors propose that the H3PW12O40 assists in the transfer of electronics from TiO2 to In2S3.

Advertisement

3. Synthesis methods to obtain TiO2-based heterostructures

A wide variety of physical and chemical methods have been developed for the synthesis of TiO2-based heterostructures, using a vast diversity of structures and morphologies. Some examples of such methods are sol-gel, solvothermal, impregnation, sputtering, dip-coating, co-precipitation, mechanical synthesis and chemical vapor deposition, among others. In many cases, a combination of two or more synthetic methods and reaction steps are needed in order to create a specific photocatalyst with the desired characteristics. The use of photocatalyst powders for water purification has been of great interest. This is because of the simplicity in the synthesis, the high exposition of the particle surface area to the target pollutants, as well as the high dispersion of the catalyst. In this section, some of the most used methods for the synthesis of TiO2-based heterostructures are explored.

Sol-gel: The sol-gel method consists of the acidic or basic hydrolysis of an organometallic precursor, followed by a slow polymerization. The obtained material is dried, allowing the decomposition and elimination of all the organic components present in the gel.

WO3/TiO2 nanocomposites have been synthesized by the sol-gel method using Ti(OBu4) and ammonium tungstate as main precursors [35, 49]. The photocatalytic activity of the material was tested via the degradation of malathion using natural sunlight. The complete degradation of the pollutant was achieved after a 2 h irradiation and a mineralization rate of 63% upon 5 h. Yang et al. [47] reported the synthesis of different CeO2/TiO2 heterostructures by the sol-gel synthetic route, using Ti(OBu)4 and Ce(NO3)3∙6H2O as precursors. The synthesized materials were tested for the degradation of methyl orange, noting the enhanced activity of the CeO2/TiO2 composite compared with unmodified TiO2 and CeO2.

Hydrothermal/solvothermal method: The solvothermal method allows the synthesis of crystalline materials by heating the precursors in solution inside a sealed reactor (autoclave). Water (hydrothermal) and several organic compounds such as ethylene glycol and glycerol (solvothermal) may be used as solvent during the reaction. The solvothermal method is one of the most common preparation techniques for heterostructures, since the variation of pressure and temperature parameters allows the formation of a wide diversity of crystal morphologies. The products obtained are usually well dispersed in form and size. Also, some additives and templates may be added into de reaction mixture to favor a desired morphology or crystallite size. Xu et al. [50] synthesized a rutile/anatase TiO2 heterostructure using titanium tetrachloride (TiCl4), urea and cetyl trimethyl ammoniumbromide (CTAB) as a template. Reaction was carried out at 160°C for 12 h, resulting in rutile/anatase nanoflowers with high surface area —up to 106.29 m2/g—. CdS/TiO2 composites were synthesized by Wu et al. [27] via the microemulsion-mediated solvothermal method, allowing the formation of anatase nanoparticles with highly dispersed CdS nanocrystals on the surface. The modification of TiO2 with CdS nanoparticles increased the absorption of visible light irradiation at 550 nm. Zhu et al. [51] prepared Bi2O3/TiO2 flower-like spheres, which displayed high photocatalytic activity due to an enhanced visible light absorption. TiO2@MoO3 core@shell structures were synthesized by Li et al. [52] using a one-step hydrothermal method, while Liu et al. [53] used this method to achieve the formation of a series of Cu2O@TiO2 core@shell structures by coating different Cu2O polyhedral nanoparticles on the TiO2 surface. Cu2O/TiO2 hollow spheres (HS) were synthesized by both solvothermal and sol–gel methods [54]. In a first step, TiO2 HS were synthesized by a sol–gel method, using carbon nanospheres as a template. In the second step, the HS were mixed with a glucose solution, containing CuSO4∙5H2O; then, the mixture was poured into a teflon-lined stainless-steel autoclave. In this reaction, glucose took the role of reducing agent, which helped to reduce copper from Cu2+ to Cu1+. These heterostructures were tested for the photocatalytic degradation of Rhodamine B under visible and sunlight irradiation, showing promising results.

Impregnation: This method consists of the saturation of one specific support—in this case TiO2—with a solution containing the desired precursor, usually a metal salt; this allows the metal ions to fill the support pores. Then, the material is dried and exposed to a thermal treatment.

Perales-Martínez et al. [55] reported the formation of the InVO4/TiO2 catalyst. In a first step, both InVO4 and TiO2 were prepared using the solvothermal method. Then, the composite was formed by suspending both oxides in methanol in order to achieve the impregnation assisted by ultrasonication. Lastly, the solvent was evaporated using a rotary evaporator. In another work, Maeda et al. [56] reported the formation of cobalt oxide nanoparticles supported on the surface of rutile TiO2 (Co3O4/TiO2 heterostructure). In this case, TiO2 was impregnated with the Co(NO3)2·6H2O solution, followed by a thermal treatment in air atmosphere. Fe2O3/TiO2 photocatalyst was prepared using a Fe(NO3)3∙9H2O ethanol solution, where TiO2 powders were stirred and sonicated [57]. Peng et al. [58] reported that after calcination at 300°C for 6 h, the Fe2O3 was deposited on the surface of TiO2 nanorods. In other report, the preparation of RuO2/TiO2 heterostructures was achieved by Uddin et al. [59] using ruthenium(III) pentan-2,4-dionate as RuO2 precursor.

UV light irradiation: This method consists of the reduction and precipitation of one or more soluble precursors over the surface of TiO2, which acts as a support. The presence of UV light irradiation allows the photo-formation of electrons in the support, which are responsible of reducing the chemical species in the solution.

MoS2/TiO2 and WS2/TiO2 have been synthesized by the photo-reduction of either (NH4)2MoS4 or (NH4)2WS4, directly on the surface TiO2 particles [42]. These materials showed good efficiency in the degradation of methylene blue and 4-chlorophenol.

Electrosynthesis: The electrosynthesis method consists of the use of electrochemical cells to produce the desired material. Yang et al. [39] achieved the electrodeposition of CdSe nanoparticles on the surface and the inner space of TiO2 nanotubes (TNTs). For this, three electrodes were used, namely a Pt wire (counter electrode), a saturated calomel electrode (reference electrode) and the TNTs (working electrode). The three electrodes were submerged in an electrolyte solution containing CdCl2 and SeO2. CdSe was deposited at −0.7 V, −6 V vs. reference electrode at room temperature. Well-dispersed CdSe nanoparticles deposited on TiO2 were obtained.

Mechanical mixing: This is one of the simplest synthesis methods, which involves the direct mixing of the heterostructure precursors. Manual mixing usually results in long reaction times and low homogeneity of the products. In certain cases, binding agents may be added to the mixture in order to increase the stability of the heterostructure.

Shifu et al. [60] prepared the WO3/TiO2 heterostructure using the mechanical mixing method. The precise amounts of TiO2 and WO3 powders were mixed in an agate ball milling tank. The two oxides were mixed for 12 h at 300 rpm. A loading of 3% wt. of WO3 presented the best efficiency for the degradation of monocrotophos. The coupled photocatalyst showed a redshift in its light absorption compared to pure TiO2.

Precipitation and co-precipitation: Both precipitation and co-precipitation methods consist of the formation of an insoluble material, starting from one or several solutions containing the soluble precursors. Usually, an increase in the pH value of the solution helps in the formation of insoluble hydroxides, allowing the precipitation. Yu et al. [61] prepared an Ag/AgCl/TiO2 heterostructure by the impregnation of TNTs with a 1 M HCl aqueous solution, followed by a second impregnation with a 0.1 M AgNO3 solution. As a result, AgCl nanoparticles were precipitated on the TNTs. Lastly, UV irradiation was applied to achieve a partial reduction of the formed AgCl nanoparticles into metallic Ag.

Thin films: One of the main burdens of using powder photocatalysts is the need of including a separation step for the effluent in order to reuse the photocatalyst in further cycles. This step can become difficult and very expensive, making the photocatalytic process less viable in a plant scale approach. A feasible solution is the immobilization of the material in a suitable support, such as glass, quartz or polymer. Some synthetic routes for obtaining photocatalysts as thin films are described below.

Dip-coating: This is one of the most used methods for the synthesis of thin films, which consists of submerging, at a constant rate, the substrate in a solution containing the precursor of the semiconductor. After a certain dwell time, the substrate is pulled out of the solution. Lastly, the solvent is dried, and a thermal treatment can be applied to eliminate organic residuals and induce crystallization of the semiconductor in the film.

Spin Coating: This process consists of putting a small amount of a solution containing the precursor of the thin film material on the surface of the substrate. Then, the substrate is rotated at high speed, eliminating the excess solution and leaving a uniform film once the solvent is dried.

Sputtering: In this route, ionized atoms (e.g., Ar+) are accelerated into the surface of a cathodic material. This collision causes that some atoms are ejected from the surface of the electrode. Subsequently, the ejected atoms are condensed on the surface of the substrate (anode), forming the thin film.

Chemical vapor deposition: This method uses volatile precursors at high temperature. The gaseous species react forming intermediates which are diffused and adsorbed on the surface of the substrate. Further reactions can take place on the surface on the substrate.

Advertisement

4. Photocatalytic activity of TiO2-based heterostructures under visible and simulated sunlight

The coupling of TiO2 with low band gap semiconductors leads to the activation of the photocatalyst material under visible light irradiation, as established earlier, resulting in turn in the generation of materials with high efficiency and stability. An important number of studies have reported the photocatalytic performance of these heterostructures, showing high conversion rates of organic and inorganic pollutants in water. Some of these results are shown in Table 1. As observed in the table, conversion of azo dyes molecules is the most used way to assess the photocatalytic activity of the synthesized materials. This is due to the easy analytical determination of such molecules in water—most of them for UV–vis spectroscopy—in comparison with uncolored molecules—such as phenols—organochlorinated compounds and pharmaceutical substances. However, as was recently pointed out, using azo dyes molecules in the evaluation of the photocatalytic performance of semiconductors may result in an artifact because of the sensitization of the semiconductors by the adsorbed organic molecules [62]. It is worth noting how the degradation rate constant is mostly determined using the pseudo first-order approach, forgetting the multiple phase conditions. In very few studies, other models—such as the Langmuir-Hinshelwood approximation—have been used [63]. Degradation yields is the most reported parameter in this kind of experiments. Very few studies follow the content of the total organic carbon throughout the process, ignoring with this the mineralization yield of the pollutants. This may lead to a miscalculation of the risk that treated water pose on the exposed organisms, since some of the photodegradation by-products may be more toxic or recalcitrant than the parent compound. Examples of this are benzoquinone, which degradation requires more energy than phenol and triclosan, which degrades into a low toxicity dioxin.

Heterostructure components Architecture Pollutant removed Reaction conditions Performance Reference
FeWO4/TiO2
Fe:Ti ratio = 95:5
Composite Salicylic acid (6.9 ppm) 300 W Xe lamp (λ ≥ 420 nm).
10 mg of composite were suspended in 50 mL.
120 min of irradiation time.
k* = 0.053 h−1 [69]
FeWO4/TiO2/CdS
Fe:Ti:Cd ratio = 5:94:1
k* = 0.099 h−1
60 wt. % Cu2S/TiO2 Orange II (15 ppm) 150 W tungsten-halogen-lamp with UV cutoff filter at 475 nm. pH ~ 6.4. k* ~ 6.1 × 10−3 min−1 [70]
Bi2O3/TiO2
Bi:Ti molar ratio = 1:100
p-chlorophenol (12.88 ppm) 150 W xenon lamp (λ ≥ 420 nm).
0.001 g mL−1 of catalyst.
6 h of irradiation time.
Degradation yield of 49% [71]
TiO2/Fe3O4 (30:70) Cr(VI)
(20 ppm of K2Cr2O7)
Mercury vapor lamp of 125 W with a broadband from 250 to 450 nm (maxima at 254, 312 and 365 nm).
30 mg of catalyst in 100 mL of solution.
Complete reduction to Cr(III) in 30 min.
k = 0.91 ppm g−1 min−1
[72]
Carbon-doped anatase/brookite TiO2 (80:20) Composite Methylene blue (3.2 ppm) Solar simulator with filter for blue light.
0.06 g of powder in 100 mL solution.
60 min irradiation.
k = 0.008 min−1 [73]
Rutile/anatase TiO2 Nanoflowers Methylene blue (15 ppm) 350 W xenon lamp.
10 mg of catalyst.
120 min of irradiation time.
72% degradation of degradation rate. [50]
Adding 0.5 mL of a 30% m/v H2O2 solution. ~88% of degradation rate.
BiVO4@TiO2 Core@shell Methylene blue (5 ppm) Osram Dulux S67 blue light Bulbs, with maximum emission at 450 nm.
50 mg of the catalyst were dispersed in 150 mL.
120 min irradiation.
~85% of Degradation rate. [74]
5 wt. % WO3/TiO2 Composite Methyl orange (20 ppm) 1000 W halogen lamp (λ > 420 nm).
Catalyst loading of 1.1 g L−1.
5 h of time irradiation.
85% degradation [34]
2,4-dichlorophenol (20 ppm) 73% degradation
12 wt. % ZnxCd1-xS/TiO2
Cd:Zn ratio = 3:1
Composite Rhodamine B (4.8 ppm) 500 W halogen lamp (420 nm < λ < 800 nm).
0.04 g of the catalyst was added to 80 mL of solution.
120 min of irradiation time.
~95% degradation [75]
Cu2O/TiO2 Cu2O nanospheres decorated with TiO2 nanoislands. Methyl Orange (30 ppm) Light intensity of ~23 mW cm−2 (λ > 400 nm).
0.03 g of catalyst was dispersed in 150 mL of solution.
40 min of irradiation time.
~97% degradation
r+~12.825 mg g−1 min−1
[76]
Cu2O@TiO2 Core@shell Polyhedra (cubes, cuboctahedra and octahedra) Methylene Blue (3.2 ppm) 300 W xenon lamp with a glass filter (λ > 400 nm). Light intensity of 23 mW cm−2.
Concentration of 0.2 mg photocatalyst mL−1.
4 h of irradiation time.
Up to 80% degradation for octahedra core-shell structure. [53]
4-Nitrophenol (10 ppm) 0.02 g of photocatalyst was dispersed in 50 mL of solution. Up to 50% degradation for octahedra core-shell structure.
Bi2MoO6/TiO2
Bi:Ti ratio = 1:2.6
Nanofibers Rhodamine B (10 ppm) 150 W xenon lamp.
0.1 g of catalyst was suspended in 100 mL of the compound solution.
5 h of irradiation.
92% degradation [77]
0.5 wt.% InVO4/TiO2 Composite Phenol (30 ppm) 300 W lamp with an intensity irradiation of 38.4 W m−2.
0.05 g of catalyst dispersed in 50 mL of solution.
80 min of irradiation time
100% of degradation rate.
k# = 8.1×10−4 s−1
[55]
Polydopamine@TiO2 Core@shell nanoparticles and brush-like structures. Geosmin (1000 ppb) 350 W xenon lamp equipped with a 420 nm cutoff filter.
120 min of irradiation
90% degradation [78]
Fluorene (1000 ppb) 98% degradation
Bi2O3/TiO2
Bi:Ti ratio = 10%
Composite Ofloxacin (25 ppm) Solar irradiation with average light intensity of 70.3 K lux.
Catalyst concentration of 0.5 g/L.
120 min of irradiation time
92.4% degradation [79]
60 wt. % Sb2S3/TiO2 Nanorods p-hydroxyazobenzene (10 ppm). 500 W halogen lamp with two cutoff filters (below 420 nm and above 850 nm).
40 mg of catalyst were dispersed in 80 mL.
5 h of irradiation time.
89% of degradation rate [80]
Methyl orange 96% of degradation rate
Ag/AlO2/TiO2
90:10 Al:Ti mass ratio
Composite Formaldehyde (10 ppm) Sunlight irradiation with an intensity of 90 ± 15 mW/cm2.
0.1 g of catalyst was dispersed in 100 mL of solution.
90 min of irradiation time.
97.8% of degradation rate [81]
Cu2O/TiO2 Nanotubes Rhodamine B (5 ppm) 500 W tungsten halogen lamp with an optical filter (λ > 420 nm).
120 min of irradiation time.
31% of degradation rate
k* = 0.00312 min−1
[82]
AgBr/Ag3PO4/TiO2
Ag:Ti molar ratio = 1:5
Spheres Methyl orange (8.2 ppm). Simulated sunlight lamp with intensity of 4 W m−2.
0.001 g of catalyst was dispersed in 6 mL of solution.
50 min of irradiation time.
~90% of degradation rate
k* = 0.1329 min−1
[83]
Microcystin-LR (50 ppm) 0.01 g of catalyst was dispersed in 30 mL of solution. 100% of degradation after 5 min.
k* = 0.6371 min−1
7 wt. % Fe2O3@TiO2 Core@shell Rhodamine B (10 ppm) 350 W xenon lamp.
5 mg of catalyst were dispersed in 50 mL of aqueous solution.
5 h of irradiation time.
60% of degradation rate
k* = 0.1605 h−1
[84]
Cu2O/TiO2 Octadecahedron/Quantum Dot Methyl orange (30 ppm) 500 W xenon lamp with glass filter (λ > 400 nm) and intensity of 100 mW cm−2.
25 mg of catalyst were dispersed in 100 mL of solution.
60 min of irradiation.
97% of degradation rate
k* = 0.055 min−1
[85]
BiOBr/TiO2
molar ratio Ti:Bi = 2:1
Nanorods Rhodamine B (15 ppm) 300 W Xenon lamp with a cut-off glass filter (λ > 420 nm). Illumination intensity of 7 × 103 mW cm−2.
70 mg of photocatalyst were dispersed in 70 mL of solution.
99.9% of degradation after 10 min of irradiation.
k = 0.49 min−1
88.5% mineralization.
[86]
Cr (VI) (50 ppm K2Cr2O7 solution) 95.4% of conversion after 120 min irradiation.
k = 0.025 min−1
Cu2O/TiO2 Nanoparticles/nanotubes p-nitrophenol (20 ppm) Xe lamp (λ > 420 nm). with an intensity of 100 mW cm−2.
210 min of irradiation.
35–40% of degradation rate under visible light irradiation.
>90% of degradation rate under sunlight irradiation.
[87]
Bi2S3/TiO2 Fungus-like mesoporous Bi2S3/ TiO2 nanotube 2,4-dichlorophenoxyacetic acid (2,4-D) (10 ppm) Xenon lamp with a radiation intensity of 85 mW cm−2.
180 min of irradiation time.
71% degradation
~75% mineralization (after 360 min of irradiation)
[41]
2,4-D and Cr(VI) (10 ppm) 81% degradation for 2,4-D
Complete Cr(VI) reduction.
Cu2O/TiO2 Hollow spheres Rhodamine B (5 ppm) 300 W xenon lamp.
0.1 g of catalyst was dispersed in 200 mL of the compound solution.
120 min irradiation time
~ 88% of degradation rate.
k* = 0.0165 min−1
[54]
300 W Xenon lamp with a 420 nm cut off filter.
300 min time irradiation.
~ 92% of degradation rate.
k* = 0.0081 min−1
1 wt. % Ag2CO3/TiO2 Composite Methyl orange (20 ppm) 300 W iodine tungsten lamp.
0.05 g of catalyst was suspended in 80 mL of the compound solution.
5 h of irradiation time.
~75% degradation
k* = 0.24 min−1
[88]

Table 1.

Degradation performance using TiO2-based heterostructures under different experimental conditions.

Degradation rate constants approximated to the pseudo first-order kinetics.


Reaction rate (degradation rate).


Kinetic constants of photodegradation estimated by the Langmuir-Hinshelwood model according to a first-order reaction.


In most cases reported in the literature, halogen lamps are the light sources in photocatalysis schemes, while the loading of the catalyst is maintained below 1 g/L, in order to avoid the screening of light due to the high turbidity in the suspension. Concentration of the organic pollutants is normally set at levels of mg/L, which are one to three magnitude orders higher than those observed in the environment, even in wastewater [1, 64, 65]. The use of such concentration levels enables the determination of the kinetic constants in the photocatalytic degradation of the pollutants in water. However, when experimental conditions require the use of complex liquid matrices, such as superficial water, groundwater or wastewater, the degradation rate is reduced, bringing the opportunity for lowering the initial concentration of the target compound toward environmentally relevant levels, such as μg/L or even ng/L, with no impact in the study of the kinetic constants.

Even when for the results shown in Table 1, direct comparisons are difficult to be established, it seems clear that tailored heterostructures formed by TiO2 and low band gap semiconductors are efficient to achieve high photocatalytic degradation yields under visible light irradiation when azo dyes are used as target pollutants. For most of the reported heterostructures, degradation yields above 80% were obtained; except for the Cu2O/TiO2 material, which performance was as low as 31%. On the other hand, very low photocatalytic efficiency is observed for refractory industrial pollutants, such as nitrophenols and chlorophenols. These compounds displayed degradation yields lower than 50% when Cu2O-TiO2 and Bi2O3/TiO2 heterostructures were used in photocatalytic assays under visible light irradiation. Some environmentally relevant pollutants, which are commonly found in surface water sources, are efficiently removed by the visible light-driven photocatalysis process. A high concentration of microcystin, a toxin produced by cyanobacteria, is fully degraded in 5 min under visible light irradiation, when AgBr/Ag3PO4/TiO2 nanospheres are used as photocatalyst, while 90% of degradation of environmentally relevant concentrations of geosmin was achieved upon 120 min using core@shell polydopamine@TiO2 composites. Also, high loads of pharmaceutically active substances, bisfenol A and the widely used herbicide 2,4-D are efficiently removed from water under visible light by the TiO2 based heterostructures. From these results, the use of these materials in advanced oxidation processes for ternary drinking water treatment sounds like a plausible option, keeping in mind that the high efficiency showed in these lab-scale studies can be affected by the complexity of the liquid matrix. Regarding heavy metals, the complete photocatalytic reduction of hexavalent chromium has been reported using Fe2O3/TiO2 and Bi2S3/TiO2 heterostructures in a very short time lapse, while BiOBr-based composites showed a slightly lower activity. For the studies reported in Table 1, the occurrence of synergistic effects was observed when the photocatalytic performance of the heterostructures and their single components was compared. For some cases, the increment in the degradation rate and degradation yields was found in the order of 1.5–5 fold, demonstrating that the efficiency of the heterostructure was significantly higher than the sum of the performance of the single components.

When the photocatalytic performance of thin films is assessed, a clear decrease in the degradation rate of organic pollutants is observed. This is because of the decrease in the number of active sites exposed to the aqueous matrix due to the immobilization of the photocatalyst on a substrate. Degradation rates as low as 30% in 15 h for of azo dyes have been reported using TiO2-InVO4 thin films [66], although the efficiency can be improved by the deposition of noble metal nanoparticles on the film surface. Conversely, in other study [67], the complete degradation of methyl orange was achieved in 8 h of visible light irradiation by using BiOCl-TiO2 thin films. This study reveals the importance of the charge carrier transference in the immobilized photocatalyst material. When this factor is taken into account, the reactivity of the thin film surface increases, leading to a higher photocatalytic activity and overcoming the mass transference hindrance.

In this sense, the arrangement of the heterostructure components is of high relevance since some approaches may favor the transfer of photo-holes or photoelectrons to the surface of the thin films. In this sense, Monfort et al. [68] tested the transfer of charge carriers in the BiVO4-TiO2 heterostructure, noting the occurrence of oxidation reactions by photo-holes when BiVO4 was located on the surface of the thin films, while reduction reactions given by photoelectron were prominent when TiO2 was in contact with the aqueous matrix.

Advertisement

5. Conclusions

Even when TiO2 displays an outstanding performance as photocatalyst, its limitation to absorb light and photoactivate under visible light irradiation makes necessary to develop a set of strategies to overcome this handicap. Coupling of low band gap semiconductors with TiO2 nanoparticles is an auspicious approach not only to redshift the light absorption of the composite, but to reduce the recombination rate of the hole-electron pair by the transference of the charge carriers from one semiconductor to another, increasing the photocatalytic performance. This leads to the generation of materials with high photoactivity and stability under visible light and sunlight irradiation.

In order to obtain functional heterostructures, care must be taken in the selection of the composite components, in order to get the best alignment of the semiconductors bands and thus the optimal transference of the charge carriers from one component to the other. Type II and III heterostructures have shown the highest efficiency in the separation of the hole-electron pairs. The defects formed in the heterounion act as transference sites for the charge carriers; although it only works when tiny loadings of semiconductor particles are deposited on TiO2, and recombination centers appear in the heterounion when the optimal loading is surpassed. The p-n heterostructures, specially the all solid Z schemes, have shown not only the efficient separation of the charge carriers in the composite, but the generation of highly oxidant photo-holes, which opens the opportunity to photodegrade highly recalcitrant organic pollutants in water. To date, very few TiO2-based Z schemes have been explored; thus, investigation should aim to develop new schemes, beyond the TiO2-Au-CdS approach, using nontoxic semiconductors, such as BiOI, ZnS, Ag2O or graphene.

TiO2-based heterostructures have shown a notable high performance in the photocatalytic removal of organic and inorganic pollutants in water under visible light irradiation. In some cases, removal yields surpass those observed for the individual components of the heterostructure, indicating the occurrence of a synergistic effect. New challenges are in the development of functional heterostructures in the form of thin films, which will optimize the energy and space consumption in photocatalytic water treatment systems.

The photocatalysis processes have the potential to move toward sustainability, through the development of sunlight active heterostructures, which simultaneously perform the oxidation of organic pollutants and the reduction of water molecule for hydrogen generation. This will lead to energy autonomous treatment systems based on sunlight-driven photocatalysis. Lastly, investigations should aim to the development of green synthesis methods, which optimize the energy consumption and minimize the use of harmful reagents.

Advertisement

Acknowledgments

The authors want to than to Secretaria de Ciencia, Tecnología e Innovación de la Ciudad de México and Dirección General de Asuntos del Personal Académico de la UNAM for the financial supporting given to this work through projects SECITI/047/2016 and IA101916, respectively.

References

  1. 1. Richardson SD, Ternes TA. Water analysis: Emerging contaminants and current issues. Analytical Chemistry. 2014;86:2813-2848. DOI: 10.1021/ac500508t
  2. 2. Murray KE, Thomas SM, Bodour AA. Prioritizing research for trace pollutants and emerging contaminants in the freshwater environment. Environmental Pollution. 2010;158:3462-3471. DOI: 10.1016/j.envpol.2010.08.009
  3. 3. Oller I, Malato S, Sánchez-Pérez JA. Combination of advanced oxidation processes and biological treatments for wastewater decontamination–A review. Science of the Total Environment. 2011;409:4141-4166. DOI: 10.1016/j.scitotenv.2010.08.061
  4. 4. Andreozzi R, Vincenzo C, Insola A, Marotta R. Advanced oxidation processes (AOP) for water purification and recovery. Catalysis Today. 1999;53:51-59. DOI: 10.1016/S0920-5861(99)00102-9
  5. 5. Matilainen A, Sillanpää M. Removal of natural organic matter from drinking water by advanced oxidation processes. Chemosphere. 2010;80:351-365. DOI: 10.1016/j.chemosphere.2010.04.067
  6. 6. Klavarioti M, Mantzavinos D, Kassinos D. Removal of residual pharmaceuticals from aqueous systems by advanced oxidation processes. Environment International. 2009;35:402-417. DOI: 10.1016/j.envint.2008.07.009
  7. 7. Deng Y, Zhao R. Advanced oxidation processes (AOPs) in wastewater treatment. Current Pollution Reports. 2015;1:167-176. DOI: 10.1007/s40726-015-0015-z
  8. 8. Padmanaban VC, Giri Nandagopal MS, Madhangi Priyadharshini G, Maheswari N, Janani Sree G, Selvaraju N. Advanced approach for degradation of recalcitrant by nanophotocatalysis using nanocomposites and their future perspectives. International Journal of Environmental Science and Technology. 2016;13:1591-1606. DOI: 10.1007/s13762-016-1000-9
  9. 9. Hoffmann MR, Martin S, Choi W, Bahnemann DW. Environmental applications of semiconductor photocatalysis. Chemical Reviews. 1995;95:69-96. DOI: 10.1021/cr00033a004
  10. 10. Kumar SG, Rao KSRK. Comparison of modification strategies towards enhanced charge carrier separation and photocatalytic degradation activity of metal oxide semiconductors (TiO2, WO3 and ZnO). Applied Surface Science. 2017;391:124-148. DOI: 10.1016/j.apsusc.2016.07.081
  11. 11. Fagan R, McCormack DE, Dionysiou DD, Pillai SC. A review of solar and visible light active TiO2 photocatalysis for treating bacteria, cyanotoxins and contaminants of emerging concern. Materials Science in Semiconductor Processing. 2016;42:2-14. DOI: 10.1016/j.mssp.2015.07.052
  12. 12. Linsebigler AL, Lu G, Yates JT. Photocatalysis on TiO2 surfaces: Principles, mechanisms, and selected results. Chemical Reviews. 1995;95:735-758
  13. 13. Sclafani A, Herrmann JM. Comparison of the photoelectronic and photocatalytic activities of various anatase and rutile forms of titania in pure liquid organic phases and in aqueous solutions. The Journal of Physical Chemistry. 1996;100:13655-13661. DOI: 10.1021/jp9533584
  14. 14. Priecel P, Adekunle Salami H, Padilla RH, Zhong Z, Lopez-Sanchez JA. Anisotropic gold nanoparticles: Preparation and applications in catalysis. Chinese Journal of Catalysis. 2016;37:1619-1650. DOI: 10.1016/S1872-2067(16)62475-0
  15. 15. Daghrir R, Drogui P, Robert D. Modified TiO2 for environmental photocatalytic applications: A review. Industrial and Engineering Chemistry Research. 2013;52:3581-3599. DOI: 10.1021/ie303468t
  16. 16. Park H, Park Y, Kim W, Choi W. Surface modification of TiO2 photocatalyst for environmental applications. Journal of Photochemistry and Photobiology C: Photochemistry Reviews. 2013:1-20. DOI: 10.1016/j.jphotochemrev.2012.10.001
  17. 17. Kumar SG, Devi LG. Review on modified TiO2 photocatalysis under UV/visible light: Selected results and related mechanisms on interfacial charge carrier transfer dynamics. Journal of Physical Chemistry A. 2011;115:13211-13241. DOI: 10.1021/jp204364a
  18. 18. Marschall R. Semiconductor composites: Strategies for enhancing charge carrier separation to improve photocatalytic activity. Advanced Functional Materials. 2014;24:2421-2440. DOI: 10.1002/adfm.201303214
  19. 19. Zhang Z, Yates JT. Band bending in semiconductors: Chemical and physical consequences at surfaces and interfaces. Chemical Reviews. 2012;112:5520-5551. DOI: 10.1021/cr3000626
  20. 20. Primo A, Corma A, García H. Titania supported gold nanoparticles as photocatalyst. Physical Chemistry Chemical Physics. 2011;13:886-910. DOI: 10.1039/C0CP00917B
  21. 21. Zhang X, Chen YL, Liu R-S, Tsai DP. Plasmonic photocatalysis. Reports on Progress in Physics. 2013;76:46401. DOI: 10.1088/0034-4885/76/4/046401
  22. 22. Jovic V, Chen WT, Sun-Waterhouse D, Blackford MG, Idriss H, Waterhouse GIN. Effect of gold loading and TiO2 support composition on the activity of Au/TiO2 photocatalysts for H2 production from ethanol-water mixtures. Journal of Catalysis. 2013;305:307-317. DOI: 10.1016/j.jcat.2013.05.031
  23. 23. Bowker M, James D, Stone P, Bennett R, Perkins N, Millard L, Greaves J, Dickinson A. Catalysis at the metal-support interface: Exemplified by the photocatalytic reforming of methanol on Pd/TiO2. Journal of Catalysis. 2003;217:427-433. DOI: 10.1016/S0021-9517(03)00074-5
  24. 24. Eustis S, El-Sayed MA. Why gold nanoparticles are more precious than pretty gold: Noble metal surface plasmon resonance and its enhancement of the radiative and nonradiative properties of nanocrystals of different shapes. Chemical Society Reviews. 2006;35:209-217. DOI: 10.1039/B514191E
  25. 25. Zhou X, Liu G, Yu J, Fan W. Surface plasmon resonance-mediated photocatalysis by noble metal-based composites under visible light. Journal of Materials Chemistry. 2012;22:21337. DOI: 10.1039/c2jm31902k
  26. 26. Kamat PV. Photophysical, photochemical and photocatalytic aspects of metal nanoparticles. Journal of Physical Chemistry B. 2002;106:7729-7744. DOI: 10.1021/jp0209289
  27. 27. Wu L, Yu JC, Fu X. Characterization and photocatalytic mechanism of nanosized CdS coupled TiO2 nanocrystals under visible light irradiation. Journal of Molecular Catalysis A: Chemical. 2006;244:25-32. DOI: 10.1016/j.molcata.2005.08.047
  28. 28. Zhou P, Yu J, Jaroniec M. All-solid-state Z-scheme photocatalytic systems. Advanced Materials. 2014;26:4920-4935. DOI: 10.1002/adma.201400288
  29. 29. Li X, Yu J, Wageh S, Al-Ghamdi AA, Xie J. Graphene in photocatalysis: A review. Small. 2016:1-57. DOI: 10.1002/smll.201600382
  30. 30. Obregón S, Colón G. A ternary Er3+-BiVO4/TiO2 complex heterostructure with excellent photocatalytic performance. RSC Advances. 2014;4:6920-6926. DOI: 10.1039/c3ra46603e
  31. 31. Hu Y, Li D, Zheng Y, Chen W, He Y, Shao Y, Fu X, Xiao G. BiVO4/TiO2 nanocrystalline heterostructure: A wide spectrum responsive photocatalyst towards the highly efficient decomposition of gaseous benzene. Applied Catalysis B: Environmental. 2011;104:30-36. DOI: 10.1016/j.apcatb.2011.02.031
  32. 32. Zhang L, Tan G, Wei S, Ren H, Xia A, Luo Y. Microwave hydrothermal synthesis and photocatalytic properties of TiO2/BiVO4 composite photocatalysts. Ceramics International. 2013;39:8597-8604. DOI: 10.1016/j.ceramint.2013.03.106
  33. 33. Pingmuang K, Chen J, Kangwansupamonkon W, Wallace GG, Phanichphant S, Nattestad A. Composite photocatalysts containing BiVO4 for degradation of cationic dyes. Scientific Reports. 2017;7:1-11. DOI: 10.1038/s41598-017-09514-5
  34. 34. Leghari SAK, Sajjad S, Chen F, Zhang J. WO3/TiO2 composite with morphology change via hydrothermal template-free route as an efficient visible light photocatalyst. Chemical Engineering Journal. 2011;166:906-915. DOI: 10.1016/j.cej.2010.11.065
  35. 35. Ramos-Delgado NA, Hinojosa-Reyes L, Guzman-Mar IL, Gracia-Pinilla MA, Hernández-Ramírez A. Synthesis by sol-gel of WO3/TiO2 for solar photocatalytic degradation of malathion pesticide. Catalysis Today. 2013;209:35-40. DOI: 10.1016/j.cattod.2012.11.011
  36. 36. Mou F, Xu L, Ma H, Guan J, Chen D, Wang S. Facile preparation of magnetic γ-Fe2O3/TiO2 Janus hollow bowls with efficient visible-light photocatalytic activities by asymmetric shrinkage. Nanoscale. 2012;4:4650-4657. DOI: 10.1039/c2nr30733b
  37. 37. Zhou W, Yin Z, Du Y, Huang X, Zeng Z, Fan Z, Liu H, Wang J, Zhang H. Synthesis of few-layer MoS2 nanosheet-coated TiO2 nanobelt heterostructures for enhanced photocatalytic activities. Small. 2013;9:140-147. DOI: 10.1002/smll.201201161
  38. 38. Serpone N, Borgarello E, Grätzel M. Visible light induced generation of hydrogen from H2S in mixed semiconductor dispersions; improved efficiency through inter-particle electron transfer. Journal of the Chemical Society, Chemical Communications. 1984;6:342-344. DOI: 10.1039/C39840000342
  39. 39. Yang L, Luo S, Liu R, Cai Q, Xiao Y, Liu S, Su F, Wen L. Fabrication of CdSe nanoparticles sensitized long TiO2 nanotube arrays for photocatalytic degradation of anthracene-9-carbonxylic acid under green monochromatic light. Journal of Physical Chemistry C. 2010;114:4783-4789. DOI: 10.1021/jp910489h
  40. 40. Toyoda T, Yindeesuk W, Kamiyama K, Katayama K, Kobayashi H, Hayase S, Shen Q. The electronic structure and photoinduced electron transfer rate of CdSe quantum dots on single crystal rutile TiO2: Dependence on the crystal orientation of the substrate. The Journal of Physical Chemistry C. 2016;120:2047-2057. DOI: 10.1021/acs.jpcc.5b09528
  41. 41. Yang L, Sun W, Luo S, Luo Y. White fungus-like mesoporous Bi2S3 ball/TiO2 heterojunction with high photocatalytic efficiency in purifying 2,4-dichlorophenoxyacetic acid/Cr(VI) contaminated water. Applied Catalysis B: Environmental. 2014;156-157:25-34. DOI: 10.1016/j.apcatb.2014.02.044
  42. 42. Ho W, Yu JC, Lin J, Yu J, Li P. Preparation and photocatalytic behavior of MoS2 and WS2 nanocluster sensitized TiO2. Langmuir. 2004;20:5865-5869. DOI: 10.1021/la049838g
  43. 43. Jianhua L, Rong Y, Songmei L. Preparation and characterization of the TiO2-V2O5 photocatalyst with visible-light activity. Rare Metals. 2006;25:636-642
  44. 44. Yang J, Dai J, Li J. Visible-light-induced photocatalytic reduction of Cr(VI) with coupled Bi2O3/TiO2 photocatalyst and the synergistic bisphenol A oxidation. Environmental Science and Pollution Research. 2013;20:2435-2447. DOI: 10.1007/s11356-012-1131-6
  45. 45. Helaïli N, Bessekhouad Y, Bouguelia A, Trari M. Visible light degradation of Orange II using xCuyOz/TiO2 heterojunctions. Journal of Hazardous Materials. 2009;168:484-492. DOI: 10.1016/j.jhazmat.2009.02.066
  46. 46. Bessekhouad Y, Robert D, Weber J-V. Photocatalytic activity of Cu2O/TiO2, Bi2O3/TiO2 and ZnMn2O4/TiO2 heterojunctions. Catalysis Today. 2005;101:315-321. DOI: 10.1016/j.cattod.2005.03.038
  47. 47. Yang H, Zhang K, Shi R, Tang A. Sol-gel synthesis and photocatalytic activity of CeO2/TiO2 nanocomposites. Journal of the American Ceramic Society. 2007;90:1370-1374. DOI: 10.1111/j.1551-2916.2007.01540.x
  48. 48. Heng H, Gan Q, Meng P, Liu X. The visible-light-driven type III heterojunction H3PW12O40/TiO2-In2S3: A photocatalysis composite with enhanced photocatalytic activity. Journal of Alloys and Compounds. 2017;696:51-59. DOI: 10.1016/j.jallcom.2016.11.116
  49. 49. Yang H, Shi R, Zhang K, Hu Y, Tang A, Li X. Synthesis of WO3/TiO2 nanocomposites via sol-gel method. Journal of Alloys and Compounds. 2005;398:200-202. DOI: 10.1016/j.jallcom.2005.02.002
  50. 50. Xu H, Li G, Zhu G, Zhu K, Jin S. Enhanced photocatalytic degradation of rutile/anatase TiO2 heterojunction nanoflowers. Catalysis Communications. 2015;62:52-56. DOI: 10.1016/j.catcom.2015.01.001
  51. 51. Zhu J, Wang S, Wang J, Zhang D, Li H. Highly active and durable Bi2O3/TiO2 visible photocatalyst in flower-like spheres with surface-enriched Bi2O3 quantum dots. Applied Catalysis B: Environmental. 2011;102:120-125. DOI: 10.1016/j.apcatb.2010.11.032
  52. 52. Li N, Li Y, Li W, Ji S, Jin P. One-step hydrothermal synthesis of TiO2@MoO3 Core-Shell nanomaterial: Microstructure, growth mechanism, and improved photochromic property. Journal of Physical Chemistry C. 2016;120:3341-3349. DOI: 10.1021/acs.jpcc.5b10752
  53. 53. Liu L, Yang W, Sun W, Li Q, Shang JK. Creation of Cu2O@TiO2 composite photocatalysts with p—N heterojunctions formed on exposed Cu2O facets, their energy band alignment study, and their enhanced photocatalytic activity under illumination with visible light. ACS Applied Materials and Interfaces. 2015;7:1465-1476. DOI: 10.1021/am505861c
  54. 54. Yin H, Wang X, Wang L, Nie Q, Zhang Y, Wu W. Cu2O/TiO2 heterostructured hollow sphere with enhanced visible light photocatalytic activity. Materials Research Bulletin. 2015;72:176-183. DOI: 10.1016/j.materresbull.2015.07.030
  55. 55. Perales-Martínez IA, Rodríguez-González V, Lee SW, Obregón S. Facile synthesis of InVO4/TiO2 heterojunction photocatalysts with enhanced photocatalytic properties under UV-vis irradiation. Journal of Photochemistry and Photobiology A: Chemistry. 2015;299:152-158. DOI: 10.1016/j.jphotochem.2014.11.021
  56. 56. Maeda K, Ishimaki K, Okazaki M, Kanazawa T, Lu D, Nozawa S, Kato H, Kakihana M. Cobalt oxide nanoclusters on rutile titania as bifunctional units for water oxidation catalysis and visible light absorption: Understanding the structure-sctivity relationship. ACS Applied Materials and Interfaces. 2017;9:6114-6122. DOI: 10.1021/acsami.6b15804
  57. 57. Zhou W, Fu H, Pan K, Tian C, Qu Y, Lu P, Sun C-C. Mesoporous TiO2/α-Fe2O3: Bifunctional composites for effective elimination of arsenite contamination through simultaneous photocatalytic oxidation and adsorption. Journal of Physical Chemistry C. 2008;112:19584-19589. DOI: 10.1021/jp806594m
  58. 58. Peng L, Xie T, Lu Y, Fan H, Wang D. Synthesis, photoelectric properties and photocatalytic activity of the Fe2O3/TiO2 heterogeneous photocatalysts. Physical Chemistry Chemical Physics. 2010;12:8033-8041. DOI: 10.1039/c002460k
  59. 59. Uddin MT, Babot O, Thomas L, Olivier C, Redaelli M, D’Arienzo M, Morazzoni F, Jaegermann W, Rockstroh N, Junge H, Toupance T. New insights into the photocatalytic properties of RuO2/TiO2 mesoporous heterostructures for hydrogen production and organic pollutant photodecomposition. Journal of Physical Chemistry C. 2015;119:7006-7015. DOI: 10.1021/jp512769u
  60. 60. Shifu C, Lei C, Shen G, Gengyu C. The preparation of coupled WO3/TiO2 photocatalyst by ball milling. Powder Technology. USA. 2005;160:198-202. DOI: 10.1016/j.powtec.2005.08.012
  61. 61. Yu J, Dai G, Huang B. Fabrication and characterization of visible-light-driven plasmonic photocatalyst Ag/AgCl/TiO2 nanotube arrays. The Journal of Physical Chemistry C. 2009;113:16394-16401. DOI: 10.1021/jp905247j
  62. 62. Barbero N, Vione D. Why dyes should not be used to test the photocatalytic activity of semiconductor oxides. Environmental Science and Technology. 2016;50:2130-2131. DOI: 10.1021/acs.est.6b00213
  63. 63. Emeline AV, Ryabchuk VK, Serpone N. Dogmas and misconceptions in heterogeneous photocatalysis. Some enlightened reflections. Journal of Physical Chemistry B. 2005;109:18515-18521. DOI: 10.1021/jp0523367
  64. 64. Reemtsma T, Jekel M. Organic Pollutants in the Water Cycle : Properties, Occurrence, Analysis and Environmental Relevance of Polar Compounds. Wiley-VCH; 2006
  65. 65. Miège C, Choubert JM, Ribeiro L, Eusèbe M, Coquery M. Fate of pharmaceuticals and personal care products in wastewater treatment plants—Conception of a database and first results. Environmental Pollution. 2009;157:1721-1726. DOI: 10.1016/j.envpol.2008.11.045
  66. 66. Ge L, Xu M, Fang H. Synthesis and characterization of the Pd/InVO4-TiO2 co-doped thin films with visible light photocatalytic activities. Applied Surface Science. 2006;253:2257-2263. DOI: 10.1016/j.apsusc.2006.04.023
  67. 67. Yang J, Wang X, Lv X, Xu X, Mi Y, Zhao J. Preparation and photocatalytic activity of BiOX-TiO2 composite films (X=Cl, Br, I). Ceramics International. 2014;40:8607-8611. DOI: 10.1016/j.ceramint.2014.01.077
  68. 68. Monfort O, Roch T, Gregor M, Satrapinskyy L, Raptis D, Lianos P, Plesch G. Photooxidative properties of various BiVO4/TiO2 layered composite fi lms and study of their photocatalytic mechanism in pollutant degradation. Journal of Environmental Chemical Engineering. 2017;5:5143-5149. DOI: 10.1016/j.jece.2017.09.050
  69. 69. Bera S, Rawal SB, Kim HJ, Lee WI. Novel coupled structures of FeWO4 /TiO2 and FeWO4/TiO2/CdS designed for highly efficient visible-light photocatalysis. Applied Materials & Interfaces. 2014;6:9654-9663. DOI: 10.1021/am502079x
  70. 70. Bessekhouad Y, Brahimi R, Hamdini F, Trari M. Cu2S/TiO2 heterojunction applied to visible light Orange II degradation. Journal of Photochemistry and Photobiology A: Chemistry. 2012;248:15-23. DOI: 10.1016/j.jphotochem.2012.08.013
  71. 71. Bian Z, Zhu J, Wang S, Cao Y, Qian X, Li H. Self-assembly of active Bi2O3/TiO2 visible photocatalyst with ordered mesoporous structure and highly crystallized anatase. Journal of Physical Chemistry C. 2008;112:6258-6262. DOI: 10.1021/jp800324t
  72. 72. Challagulla S, Nagarjuna R, Ganesan R, Roy S. Acrylate-based polymerizable sol-gel synthesis of magnetically recoverable TiO2 supported Fe3O4 for Cr(VI) photoreduction in aerobic atmosphere. ACS Sustainable Chemistry & Engineering. 2016;4:974-982. DOI: 10.1021/acssuschemeng.5b01055
  73. 73. Etacheri V, Michlits G, Seery MK, Hinder SJ, Pillai SC. A highly efficient TiO2– x Cx nano-heterojunction photocatalyst for visible light induced antibacterial applications. ACS Applied Materials & Interfaces. 2013;5:1663-1672. DOI: 10.1021/am302676a
  74. 74. Kontic R, Patzke GR. Synthetic trends for BiVO4 photocatalysts: Molybdenum substitution vs. TiO2 and SnO2 heterojunctions. Journal of Solid State Chemistry. 2012;189:38-48. DOI: 10.1016/j.jssc.2011.11.050
  75. 75. Li W, Li D, Meng S, Chen W, Fu X, Shao Y. Novel approach to enhance photosensitized degradation of rhodamine B under visible light irradiation by the ZnxCd1-xS/TiO2 nanocomposites. Environmental Science and Technology. 2011;45:2987-2993. DOI: 10.1021/es103041f
  76. 76. Liu L, Yang W, Li Q, Gao S, Shang JK. Synthesis of Cu2O nanospheres decorated with TiO2 nanoislands, their enhanced photoactivity and stability under visible light illumination, and their post-illumination catalytic memory. ACS Applied Materials and Interfaces. 2014;6:5629-5639. DOI: 10.1021/am500131b
  77. 77. Zhang M, Shao C, Mu J, Zhang Z, Guo Z, Zhang P, Liu Y. One-dimensional Bi2MoO6/TiO2 hierarchical heterostructures with enhanced photocatalytic activity. Crystal Engineering Communication. 2012;14:605-612. DOI: 10.1039/C1CE05974B
  78. 78. Liu S, Hu Q, Qiu J, Wang F, Lin W, Zhu F, Wei C, Zhou N, Ouyang G. Enhanced photocatalytic degradation of environmental pollutants under visible irradiation by a composite coating. Environmental Science and Technology. 2017;51:5137-5145. DOI: 10.1021/acs.est.7b00350
  79. 79. Sood S, Mehta SK, Sinha ASK, Kansal SK. Bi2O3/TiO2 heterostructures: Synthesis, characterization and their application in solar light mediated photocatalyzed degradation of an antibiotic, ofloxacin. Chemical Engineering Journal. 2016;290:45-52. DOI: 10.1016/j.cej.2016.01.017
  80. 80. Sun M, Chen G, Zhang Y, Wei Q, Ma Z, Du B. Efficient degradation of azo dyes over Sb2S3/TiO2 heterojunction under visible light irradiation. Industrial & Engineering Chemistry Research. 2012;51:2897-2903. DOI: 10.1021/ie2025882
  81. 81. Tang A, Jia Y, Zhang S, Yu Q, Zhang X. Synthesis, characterization and photocatalysis of AgAlO2/TiO2 heterojunction with sunlight irradiation. Catalysis Communications. 2014;50:1-4. DOI: 10.1016/j.catcom.2014.02.015
  82. 82. Wang M, Sun L, Lin Z, Cai J, Xie K, Lin C. p–n Heterojunction photoelectrodes composed of Cu2O-loaded TiO2 nanotube arrays with enhanced photoelectrochemical and photoelectrocatalytic activities. Energy & Environmental Science. 2013;6:1211-1220. DOI: 10.1039/c3ee24162a
  83. 83. Wang X, Utsumi M, Yang Y, Li D, Zhao Y, Zhang Z, Feng C, Sugiura N, Cheng JJ. Degradation of microcystin-LR by highly efficient AgBr/Ag3PO4/TiO2 heterojunction photocatalyst under simulated solar light irradiation. Applied Surface Science. 2015;325:1-12. DOI: 10.1016/j.apsusc.2014.10.078
  84. 84. Xia Y, Yin L. Core–shell structured α-Fe2O3@TiO2 nanocomposites with improved photocatalytic activity in the visible light region. Physical Chemistry Chemical Physics. 2013;15:18627-18634. DOI: 10.1039/c3cp53178c
  85. 85. Xu X, Gao Z, Cui Z, Liang Y, Li Z, Zhu S, Yang X, Ma J. Synthesis of Cu2O octadecahedron/TiO2 quantum dot heterojunctions with high visible light photocatalytic activity and high stability. ACS Applied Materials & Interfaces. 2016;8:91-101. DOI: 10.1021/acsami.5b06536
  86. 86. Xue C, Zhang T, Ding S, Wei J, Yang G. Anchoring tailored low-index faceted BiOBr nanoplates onto TiO2 nanorods to enhance the stability and visible-light-driven catalytic activity. ACS Applied Materials and Interfaces. 2017;9:16091-16102. DOI: 10.1021/acsami.7b00433
  87. 87. Yang L, Luo S, Li Y, Xiao Y, Kang Q, Cai Q. High efficient photocatalytic degradation of p-nitrophenol on a unique Cu2O/TiO2 p-n heterojunction network catalyst. Environmental Science and Technology. 2010;44:7641-7646. DOI: 10.1021/es101711k
  88. 88. Yu C, Wei L, Chen J, Xie Y, Zhou W, Fan Q. Enhancing the photocatalytic performance of commercial TiO2 crystals by coupling with trace narrow-band-gap Ag2CO3. Industrial and Engineering Chemistry Research. 2014;53:5759-5766. DOI: 10.1021/ie404283d

Written By

Raquel Del Angel, Juan C. Durán-Álvarez and Rodolfo Zanella

Reviewed: 15 March 2018 Published: 05 November 2018