Next Article in Journal
Interfacial Processes—The Key Steps of Phase Transfer Catalyzed Reactions
Next Article in Special Issue
Photocatalytic Porous Silica-Based Granular Media for Organic Pollutant Degradation in Industrial Waste-Streams
Previous Article in Journal
Fabrication of a Cu2O-Au-TiO2 Heterostructure with Improved Photocatalytic Performance for the Abatement of Hazardous Toluene and α-Pinene Vapors
Previous Article in Special Issue
Catalytic Pyrolysis as a Technology to Dispose of Herbal Medicine Waste
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A CeO2 Semiconductor as a Photocatalytic and Photoelectrocatalytic Material for the Remediation of Pollutants in Industrial Wastewater: A Review

by
Elzbieta Kusmierek
Institute of General and Ecological Chemistry, Faculty of Chemistry, Lodz University of Technology, Zeromskiego 116 St., 90-924 Lodz, Poland
Catalysts 2020, 10(12), 1435; https://doi.org/10.3390/catal10121435
Submission received: 22 November 2020 / Revised: 3 December 2020 / Accepted: 4 December 2020 / Published: 8 December 2020
(This article belongs to the Special Issue Catalytic Remediation for Industrial Wastes)

Abstract

:
The direct discharge of industrial wastewater into the environment results in serious contamination. Photocatalytic treatment with the application of sunlight and its enhancement by coupling with electrocatalytic degradation offers an inexpensive and green technology enabling the total removal of refractory pollutants such as surfactants, pharmaceuticals, pesticides, textile dyes, and heavy metals, from industrial wastewater. Among metal oxide—semiconductors, cerium dioxide (CeO2) is one of the photocatalysts most commonly applied in pollutant degradation. CeO2 exhibits promising photocatalytic activity. Nonetheless, the position of conduction bands (CB) and valence bands (VB) in CeO2 limits its application as an efficient photocatalyst utilizing solar energy. Its photocatalytic activity in wastewater treatment can be improved by various modification techniques, including changes in morphology, doping with metal cation dopants and non-metal dopants, coupling with other semiconductors, and combining it with carbon supporting materials. This paper presents a general overview of CeO2 application as a single or composite photocatalyst in the treatment of various pollutants. The photocatalytic characteristics of CeO2 and its composites are described. The main photocatalytic reactions with the participation of CeO2 under UV and VIS irradiation are presented. This review summarizes the existing knowledge, with a particular focus on the main experimental conditions employed in the photocatalytic and photoelectrocatalytic degradation of various pollutants with the application of CeO2 as a single and composite photocatalyst.

1. Introduction

Water resource contamination and energy crises constitute the most important issues in the modern world. Both problems affect not only the quality of human life, but also human health, considering the availability of drinking water and energy resources, as well as increasing contamination of the environment, which is especially noticeable with regards to climate change.
One of the main sources of environmental pollution is untreated industrial wastewater. Its direct release to surface water results in a serious contamination of fresh water. Photocatalytic and photoelectrocatalytic treatment with the application of sunlight offers an inexpensive and green technology for possible complete removal of refractory pollutants, such as surfactants, pharmaceuticals, pesticides, textile dyes, and heavy metals, from industrial wastewater. Furthermore, photoelectrochemical treatment of wastewater can be coupled with a simultaneous recovery of energy stored in the wastewater. This chemical energy is often wasted [1].
The demand for efficient and stable materials which can be applied as a photocatalyst or photoelectrocatalyst in wastewater treatment under visible light irradiation has increased during the last two decades. In particular, metal oxide–semiconductors have attracted great interest. Among these materials, cerium dioxide (CeO2) is one of the semiconductors most often applied, with the exception of titanium dioxide (TiO2). Metal oxides have been successfully applied as photocatalysts and photoelectrocatalysts in treatment processes of various pollutants [2,3,4,5,6,7,8,9,10,11,12,13,14,15,16]. They offer a number of ways to enhance the treatment process efficiency and energy efficiency with a simultaneous reduction of environmental pollution. TiO2 is the most commonly applied semiconductor photocatalyst due to its properties, such as its non-toxicity to living organisms, chemical stability and inertness, low-cost preparation, and photocatalytic properties resulting from its band gap energy and positions of conduction band (CB) and valence band (VB) [17,18,19]. This photocatalyst can be applied under UV irradiation, which is hazardous and constitutes only 2–5% of total solar light. TiO2 exhibits another limitation—a relatively high rate of recombination of photogenerated electron/hole (e/h+) pairs, which results in a decrease of its photocatalytic activity [20,21].
CeO2 is another semiconductor photocatalyst with various applications and similar properties to TiO2. However, its band gap is in the wide range of 2.6 to 3.4 eV, depending on the preparation method [22,23]. Furthermore, CeO2 exhibits promising photocatalytic activity. Nonetheless, the position of CB and VB limits its application as an efficient photocatalyst utilizing solar energy, even though CeO2 can absorb a larger fraction of the solar spectrum than TiO2 [24]. The photocatalytic and photoelectrocatalytic activity of CeO2 in wastewater treatment can be improved by various modification techniques, including changes in morphology, doping with metal cation dopants and non-metal dopants, coupling with other semiconductors, combining it with carbon supporting materials, etc. [24,25,26,27,28,29,30,31,32,33].
The main properties that make CeO2 significant as a photocatalyst and photoelectrode material applied in the degradation of various pollutants result from its high band gap energy, high refractive index, high optical transparency in the visible region, high oxygen storage capacity, and chemical reactivity [34,35,36,37,38,39,40,41]. The other properties of CeO2 which should be mentioned include its high thermal stability, high hardness, oxygen ion conductivity, special redox features, and easy conversion between Ce3+ and Ce4+ oxidation states [30,42,43].
The above-mentioned properties indicate the possible practical application of CeO2 as a photocatalyst or photoelectrocatalyst in wastewater treatment. CeO2 also has other versatile practical applications as a coating for corrosion protection for metals and alloys [44,45], as an oxygen ion conductor in solid oxide fuel cells [46,47], as glass-polishing material [48], as an electro-chromic thin film [39,49], as sunscreen in cosmetics [50,51], as a gas sensor [52,53], as an additive in ceramics [54,55], as anode material for the lithium ion battery system [56], in three-way catalysis (TWC) [57,58], in fluid catalytic cracking [59], in oxygen pumps and amperometric oxygen monitors [60], in water splitting for the generation of hydrogen [61], in solar cells [62], etc.
The aim of this paper is to present a general overview of CeO2 semiconductor materials applied as photocatalysts and photoelectrocatalytic materials in the treatment of various pollutants present in industrial wastewater. The characteristics of CeO2 semiconductors, as well as parameters characterizing the treatment process, are presented. They are followed by examples of CeO2 applied in the photocatalytic and photoelectrocatalytic treatment of different kinds of pollutants, in order to prove the extraordinary properties of CeO2. This review summarizes the existing knowledge, with a particular focus on the experimental conditions of pollutant degradation in photocatalytic and photoelectrocatalytic processes.

2. Ceria Characterization

Cerium dioxide (CeO2) is an n-type semiconductor material characterized by a relatively high dielectric constant value (ɛ = 23–52) [63] and a wide band gap of about 3.0 eV [22]. Although rare earth elements usually exist in a trivalent state, cerium exists not only in a trivalent (+3) state, but also in a tetravalent (+4) state. The Ce4+ oxidation state is regarded as more stable than Ce3+ due to Ce4+’s electronic structure ([Xe]4f0), which is more stable than that of Ce3+ ([Xe]4f1). Cerium oxide also exists in two forms, with ongoing transformation between the oxygen-rich CeO2 and the oxygen-poor Ce2O3 [64].
Generally, CeO2 exhibits a simple and stable cubic fluorite structure with the space group Fm3m. Cerium ions form a face-centered-cubic (fcc) structure, while oxygen ions form a cubic sublattice [65]. Each cerium cation is coordinated by the eight nearest oxygen ions and each oxygen anion is coordinated by the four nearest cerium cations [46,66,67]. The coordination of 8 to 4 enables oxide defects in CeO(2−x). Oxide defects play an important role in the catalytic and photocatalytic properties of CeO2 [68].
In bulk materials, both forms of cerium oxide are present, i.e., CeO2 and Ce2O3. In the case of Ce2O3, charge deficiency is compensated for by oxygen vacancies in the lattice, resulting in the formation of oxygen defects [69,70]. The transition between the two oxidation states of cerium ions, i.e., from Ce3+ to Ce4+, may generate neutral oxygen vacancies in ceria according to the following reaction [46]:
O 2 + 2 Ce 4 + 2 Ce 3 + + 1 / 2 O 2 ( g ) .
The transition between Ce3+ and Ce4+ accompanies the process of oxygen vacancy formation. In the case of a cubic fluorite structure, atoms of oxygen can quickly diffuse since they are all in one plane. The oxygen vacancies can be formed in the following reaction [46,71]:
O O X V O + 2 e + 1 / 2 O 2 ( g ) ,
where OOX denotes oxide ions in the lattice, VO•• represents a doubly charged oxygen vacancy, and e stands for electrons in CB. The conduction band is formed of Ce 4f energy states [67]. The number of oxygen vacancies is directly proportional to the concentration of Ce3+ ions. Furthermore, a reduction of the ceria particle size results in the formation of oxygen vacancies related to an increase in the levels of Ce3+ [72]. The catalytic activity of ceria strongly depends not only on the particle size, but also on the morphology, as proven in different oxidation and reduction processes, e.g., the oxidation of CO [73,74] and NO reduction [75,76]. Various shapes of CeO2 exhibit different exposures of lattice planes. Nanoparticles are often present in the form of octahedra or a truncated octahedral shape, and their most stable facet is {111} due to the minimization of surface energy. Nanorods expose {110} facets, while nanocubes expose {100} facets [66,76]. The stability of different facets is in the following order: {111} > {110} > {100}. Furthermore, the formation energy of oxygen vacancies on the {111} facet is higher than on {110} and {100} facets, which means that the number of oxygen vacancies depends on the facet. Therefore, different catalytic activity was determined in the case of ceria applied in the form of nanorods, nanocubes, and octahedra in CO oxidation [73,74,76]. The highest catalytic activity was determined for nanorods and the lowest for octahedra. This means that more oxygen vacancies are formed on the surface of nanorods and nanocubes.
If an organic pollutant is oxidized at ceria, surface lattice oxygen atoms act as an oxidant and an oxygen vacancy is formed, according to the Mars and van Krevelen (MvK) mechanism [71]. This mechanism explains oxidation reactions on catalysts consisting of metal oxides and based on redox reactions of active sites [77,78]. The oxygen vacancies in CeO2 are important not only because they enhance the activity and stability of supported catalysts, but also because they stimulate oxygen-containing bonds for hydrogen production with the application of a water-splitting reaction [79]. They act as electron scavengers and specific reaction sites in heterogeneous catalysis, and generate superoxide radicals [80,81,82,83]. Furthermore, oxygen vacancies bind adsorbates (e.g., molecules of pollutants) more strongly than normal oxide sites and also assist in their dissociation [84,85,86].
Although a higher concentration of Ce3+ in ceria means the formation of more defects, a higher concentration of Ce4+ ions results in a narrower bandgap of CeO2 and photocatalytic activity of ceria under visible light can be enhanced [87]. Cerium oxide exhibits strong absorption in the UV region, and high transparency in the VIS region and near IR region. In comparison with TiO2, the CeO2 absorption spectrum is shifted by 80 nm towards the VIS region [88,89]. Additionally, the photogenerated pairs e/h+ have much longer lifetimes than those generated in TiO2.
CeO2 is an n-type semiconductor with a relatively wide band gap. Its experimentally determined band gap energy is about 6 eV, and was measured from VB with a predominantly O 2p character to CB with a predominantly Ce 5d character [90,91,92,93]. However, localized and empty Ce 4f states lie within the gap, and the band gap in ceria can be described by the value of about 3 eV as the distance from VB to the bottom of the empty 4f states [90,94]. The unoccupied Ce 4f electronic states were proved to lie above the Fermi level [95]. The position of the CB edge and VB edge can be calculated at the point of zero charge, according to the following equations [96,97,98,99]:
E CB 0 ( CeO 2 ) = χ ( CeO 2 ) E c ½ E g ( CeO 2 ) ,
E VB 0 ( CeO 2 ) = E g ( CeO 2 ) E CB ( CeO 2 ) ,
where ECB and EVB denote conduction and valence band potentials, respectively; χ is the absolute electronegativity of the semiconductor, which is the geometric mean of the electronegativity of the constituent atoms (5.56 eV for CeO2) [97,100]; Eg is the band gap energy; and Ec is the energy of free electrons on the hydrogen scale (~4.5 eV). Therefore, the calculated values of ECB and EVB for CeO2 with the band gap energy of ~3 eV are equal to −0.44 and 2.56 eV, respectively. The positions of VB and CB in ceria compared with several selected redox potentials occurring at the semiconductor surface are presented in Figure 1.
The position of edges in CB and VB is one of the most important controlling factors in the production of reactive oxide species (ROS). Considering that the positions of CB and VB in CeO2 include the potentials of hydroxyl radical and superoxide anion radical formation (Figure 1), these reactive oxide species can be formed at the semiconductor surface and can participate in the degradation of pollutants present in industrial wastewater. Finally, water and carbon dioxide—the non-toxic products of pollutants degradation—are formed. Furthermore, the potential of CB in CeO2 is more negative than the potential of hydrogen evolution and the VB potential is more positive than the potential of oxygen evolution. This means that the requirements of hydrogen generation in a photocatalytic system are fulfilled. However, H2 and O2 formed during the water splitting process can undergo the reverse reaction, resulting in the generation of H2O.
Ceria can be applied not only as a photocatalyst, but also as an electrode material, in the photoelectrocatalytic degradation of pollutants. Then, two other parameters characterizing semiconductors—the Fermi level potential and flat band potential—should be considered. If a semiconductor is used as an electrode material, then the electrochemical potential or, equivalently, the Fermi level of electrons in the electrode (EF), can be adjusted by changing the applied potential (E) vs. the reference electrode [101]:
E F = eE + const . ,
where e denotes the elementary charge.
Considering an n-type semiconductor, if the Fermi level in the redox electrolyte lies below the Fermi level in a semiconductor, then electrons can be transferred from the semiconductor to the oxidized species until equilibrium is achieved. This results in the formation of a depletion layer in the semiconductor due to the removal of electrons and in a bending of bands upward, toward the surface (Figure 2) [102,103]. The difference between the Fermi levels in the semiconductor and the redox electrolyte determines the band bending. Under irradiation, the Fermi level rises up and undergoes splitting into two quasi-Fermi levels: EF,e−, representing the quasi-Fermi level of electrons for the CB, and EF,h+, representing the quasi-Fermi level of holes for the VB [104]. The quasi-Fermi level of the majority of charge carriers is close to the original Fermi level. The quasi-Fermi level of the minority of charge carriers clearly shifts away from the original Fermi level (Figure 2).
The Fermi level in the semiconductor can be controlled by the application of a bias potential, which changes the band bending [105,106,107]. If the applied potential causes flattening of the band edges, then it is called flat band potential (Efb). This results in a lack of differences between the potential at the surface and in the bulk of the semiconductor. If the applied potential is higher than Efb, the potential causes higher band bending [105,108]. Efb can be determined from a Mott–Schottky plot (1/C2 vs. E) by its extrapolation to 1/C2 = 0 resulting from the Mott–Schottky equation [22,23,101,109]:
1 C 2 = ( 2 e ε 0 ε S A 2 N D ) ( E E fb k B T e ) ,
where C denotes the space-charge capacitance, ND (cm−3) is the donor density, e represents the electronic charge (1.602 × 10−19 C), ε denotes the relative permittivity of the semiconductor, A (cm2) is the active geometric surface area, ε0 represents the vacuum permittivity (8.854 × 10−12 F m−1), E (V vs. the reference electrode) is the applied potential, kB indicates the Boltzmann constant (1.381 × 10−23 J K−1), and T is the temperature (298 K).
Furthermore, a positive slope of the Mott–Schottky equation indicates the n-type nature of a semiconductor, while a negative slope is characteristic of a p-type semiconductor [110].

3. Ceria as a Photocatalyst and Photoelectrocatalyst Applied in Pollutant Degradation

Photocatalytic materials, including metals, metal oxides, semiconductors, carbon-based structures, quantum-dots, metal–organic frameworks, magnetic cored dendrimers, etc., can be applied in the degradation of various pollutants present in industrial wastewater. CeO2 is a metal oxide and a semiconductor material. Generally, metal oxide photocatalysts can be divided into three generations [111]. The first generation includes single component oxides, sulfides, nitrides, and phosphates. The second generation was developed in order to improve the photocatalytic activity of catalysts belonging to the first generation and includes heterojunction materials. The photocatalytic materials deposited on different substrates constitute the third generation of photocatalysts. The last group of photocatalysts was created in order to immobilize the photocatalysts belonging to the first and second generation, and in order to avoid separation of the photocatalysts from wastewater after its treatment. CeO2 is a metal oxide and a semiconductor material and belongs to the first generation of photocatalysts. However, ceria can also be applied in the form of heterojunction photocatalysts, and independent of its form, should be immobilized on proper substrates or supports.
The degradation of pollutants in the photocatalytic process performed under UV or VIS irradiation depends on many parameters, including the pH of wastewater, effect of catalyst loading expressed by the ratio of catalyst to pollutant or catalyst to volume of wastewater, adsorption of a pollutant on a photocatalyst surface, source and light intensity, pollutant loading, and presence of interfering compounds [111,112].
A comparison of the performance of various photocatalysts is very difficult, because it depends on many conditions and, primarily, on the type of photocatalyst and type of pollutant. Nevertheless, two parameters can be included in such a comparison. The first one is the quantum yield (QY) of a photochemical reaction. The quantum yield is usually defined as the ratio of the number of molecules changed, formed, or destroyed to the number of photons of the specified wavelength adsorbed in the same period of time [113]. In the case of the photocatalytic degradation of a pollutant, QY can be calculated according to the following equation [111,114,115]:
QY = Rate   of   pollutant   degradation   ( molecules   per   second ) Flux   of   absorbed   photons   ( photons   per   second ) .
Anwer et al. introduced the second parameter, which enables a comparison of different photocatalysts and constitutes the supplementation of QY [111,116]. This parameter is called “a figure of merit” (FOM) and is based on important operation parameters:
F OM = Product   obtained   ( L ) Catalyst   dosage   ( g / L )   · Time   ( h ) · Energy   consumption   ( Wh / μ mole ) ( Wh μ mole ) .
However, the application of the above-described FOM requires a benchmark for the normalization of QY values determined for considered photocatalysts, for which the FOM value is established as 100. The relationship between QY and FOM values can be direct or reverse. A reverse relationship indicates a deficiency in at least one of the four parameters which are included in the formula describing FOM [111].

3.1. Pristine CeO2

The energy of the band gap in a semiconductor determines its possible application as a photocatalyst in wastewater treatment under UV or VIS/sunlight irradiation. Its value can be decreased by an application of CeO2 in the form of nanomaterials, e.g., nanoparticles, nanotubes, nanowires, or nanorods. A decrease in the band gap enables the performance of a photocatalytic degradation process under visible light irradiation or even sunlight, which is harmless and more economic than UV irradiation. Ceria nanomaterials can be fabricated in various ways, including by the sol-gel process, co-precipitation, hydrothermal synthesis, forced hydrolysis, electrochemical, solvothermal synthesis, reverse micelles route, sonochemical, or microwave methods [117]. The method of ceria preparation strongly affects the photocatalytic activity of CeO2. It was proved that an increase in the specific surface area, a decrease in the pore size, and an increase in the amount of oxygen vacancies resulted in a high rate of hydrogen production in the photocatalytic process carried out under visible and solar light [118].
During the photocatalytic degradation of pollutants, the photoexcitation of CeO2 in aqueous solution results in the generation of various radicals and charged species, according to the common route [102,119,120]:
CeO 2 + h ν CeO 2 ( e CB + h VB + ) ,
CeO 2 ( h VB + ) + H 2 O CeO 2 + H + + OH ,
CeO 2 ( h VB + ) + OH CeO 2 + OH ,
CeO 2 ( e CB ) + O 2 CeO 2 + O 2 ,
O 2 + H + HO 2 ,
2 HO 2 H 2 O 2 + O 2 ,
H 2 O 2 + CeO 2 ( e CB ) CeO 2 + OH + OH ,
where denotes the photon energy, VB is the valence band in CeO2, CB represents the conduction band in CeO2, h+ is a hole, and e denotes an electron.
Hydroxyl radicals (OH), as well as superoxide anion radicals (O2•−), are the main reactive oxygen species (ROS) generated in the photocatalytic system in water oxidation by photogenerated holes (h+) and oxygen reduction by photoexcited electrons (e) in CB. These species, along with hydroperoxide radicals (HO2), are highly reactive and can take part in photocatalytic degradation of the pollutants present in industrial wastewater. The reactivity of ROS can be compared considering their oxidation potentials, which are as follows: 2.80 V for OH; 1.78 V for H2O2; 1.68 V for HO2; and 1.23 V for O2 vs. the normal hydrogen electrode (NHE) [121,122].
The properties of CeO2 can be changed significantly when its particle size is reduced to a nanoscale and is applied in the form of nanostructures. This process affects CeO2 properties such as the lattice symmetry, cell parameters, and structural characteristics [123]. Phases in a bulk form are characterized by a high surface energy and are unstable. When the particle size is reduced to a nanoscale, the surface energy decreases significantly, and nanostructured materials exhibit a high stability [124]. Furthermore, ceria in the form of nanoparticles shows high catalytic activity, which results from the high mobility of surface oxygen vacancies [73,75,125].
CeO2 in the pristine form can be applied in the photocatalytic treatment of various pollutants (Table 1). It is mainly applied in the form of doped or composite materials. Since CeO2, similarly to TiO2, is characterized by a relatively high band gap energy and high rate of e/h+ pair recombination, its photoexcitation requires the application of UV irradiation. Although different methods of CeO2 preparation, especially in the form of nanomaterials, enable these issues to be avoided, pristine CeO2 is not commonly applied in the photocatalytic degradation of organic and inorganic pollutants. Examples of CeO2 application as a single photocatalyst in pollutant degradation are presented in Table 1.
Pristine CeO2 can be fabricated with the application of different methods (Table 1) affecting its properties. This photocatalyst is usually applied in the form of nanomaterials, which increases its photocatalytic activity, but simultaneously generates problems in its separation from the solution after treatment. The examples presented in Table 1 clearly show that nanoparticles, nanowires, and nanorods exhibit a band gap energy lower than 3.2 eV, except for hierarchical nanostructures [25]. However, the application of hierarchical nanorods and nanowires in methyl orange (MO) photodegradation resulted in a higher degradation rate due to a higher surface area and improved light harvesting ability.
It is worth noting that CeO2 can be applied in the photodegradation of organic pollutants, such as synthetic dyes or phenols, at low concentrations, but the duration of the process required for the complete removal of pollutants is relatively long—about 2–3 h or sometimes 10 h or more. Furthermore, the removal of organic pollutants does not mean their total degradation and total mineralization of the solutions. A decrease in total organic carbon (TOC) is rarely investigated along with pollutant removal. A significant decrease in TOC usually requires longer durations of photocatalytic degradation. However, CeO2 nanoparticles obtained in the dissolution and hydrolysis method [130], and applied in phenol derivative photodegradation, resulted in comparable and high degradation efficiencies (>96%) and TOC removal (>91%) during the same time of 180 min under solar irradiation, except for phenol.
CeO2 in the form of nanorods is also a promising material for obtaining an enhanced solar photocatalyst with a possible application in the photocatalytic production of H2 [118]. Therefore, the application of CeO2 as an electrode material in the photoelectrocatalytic degradation of pollutants under solar irradiation, combined with the simultaneous photoelectrocatalytic generation of H2, seems to be an interesting alternative to the commonly applied photocatalytic treatment of industrial wastewater.
The last example presented in Table 1 is related to CeO2 supported with sugarcane bagasse (SCB) adsorbent, in order to achieve higher photocatalytic activity in methylene blue (MB) photodegradation [134]. In fact, this photocatalyst constitutes a composite material showing the effect of a support on CeO2 characteristics, but it is not a composite with another photocatalyst. CeO2/SCB exhibits a higher BET (Brunauer-Emmett-Teller) surface area and more active sites, resulting in a three-times higher efficiency of MB photodegradation. Furthermore, the adsorption edge shifts towards longer wavelengths with a simultaneous decrease in the CeO2/SB band gap.

3.2. CeO2 Doped with Metals and Non-Metals

An enhancement of the photocatalytic activity and extension of light absorption into the visible region is possible by doping CeO2. The doping can be performed with metal dopants or non-metal dopants. In both cases, the superior photocatalytic performance of doped CeO2 is ascribed to a decrease in the band gap energy and bringing the absorption band from UV into the visible region, an increase in the specific surface area, and an increase in the number of oxygen vacancies.
The doping of CeO2 with metals or non-metals results in a change in the position of the top valence band level and bottom of the conduction band level. The effect of CeO2 doping with B, C, and N on the energy of CB and VB has been investigated [135]. In the case of B doping, the narrower band gap of 3.10 eV in comparison with pure CeO2 (3.20 eV) resulted from a simultaneous shift of the CB and VB to more negative energies in relation to the Fermi level. A similar effect was observed in the case of C doping, although the movement was a little bit smaller than that observed for B doping and resulted in the energy band gap of 3.02 eV. In contrast to B and C doping, N doping caused a shift of the CB and VB towards more positive values while the band gap energy of N-CeO2 was 3.01 eV [135].
CeO2 doping with metals resulted in higher photocatalytic activity due to a better separation of h+/e pairs [26,29]. Metal dopants can act as electron acceptors and/or hole donors and facilitate charge carrier localization (Figure 3).
CeO2 doped with metals and non-metals was successfully applied as a photocatalyst in the degradation of different pollutants which are present in wastewater (Table 2).
One of the non-metal dopants of CeO2 is nitrogen, which has drawn great attention due to its size similar to oxygen and its relatively low ionization energy [138]. Nitrogen can be doped to other metal oxides, especially those applied in the form of nanomaterials, with the application of a wet-chemical route, high-temperature treatment sintering, and ion implementation [148]. These methods were successfully applied in the case of TiO2. Ceria powder was doped with N by sintering in the presence of NH3 at the temperature of 700 °C and XPS analysis confirmed the introduction of N to a cubic fluorite-type structure of CeO2 [149]. The wet-chemical method was applied in the incorporation of N into CeO2 nanoparticles, resulting in the enhancement of visible light absorption [138]. The degradation of MB proceeded with a clearly higher efficiency under visible light in comparison with the undoped ceria.
Fluorine is another non-metal element applied in ceria doping, performed in order to enhance its photocatalytic activity. F-doped CeO2 nanoparticles were synthesized using low-temperature solution combustion followed by heat treatment in air [136], resulting in the formation of a smaller particle size in the form of nanocubes with a higher percentage of reactive facets. At the same time, the band gap energy was decreased in comparison with the undoped ceria.
The solvothermal method with hexamethylenetetramine as a precipitator was applied in the preparation of CeO2 nanomaterials co-doped with C and N [28]. These nanomaterials exhibited significantly higher photocatalytic activity when tested in acid orange 7 (AO7) photodegradation.
An enhancement of ceria’s photocatalytic activity can also be achieved by doping it with different transition metals characterized by valances lower than 4+ [142]. Such doping usually results in an increase in oxygen vacancies. Ceria nanomaterials doped with Ti, Co, Mn, or Fe were synthesized using a reverse co-precipitation method [150] and their photocatalytic activity was determined in MB photocatalytic degradation. The rate constant of MB degradation calculated for doped ceria nanoparticles (Ce-NP) decreased in the following order: Ce(Co)-NP > Ce(Mn)-NP > Ce(Ti)-NP > Ce(Fe)-NP > Ce-NP. This order is according to a decrease in band gap energy values determined for co-doped ceria materials in comparison to undoped ceria, i.e., 2.69 eV (Ce(Co)-NP), 2.77 eV (Ce(Ti)-NP), 2.78 eV (Ce(Mn)-NP), and 2.81 eV (Ce(Fe)-NP) in comparison to 3.07 eV determined for pristine Ce-NP.
Cobalt-doped ceria obtained in the co-precipitation method was tested in MB photodegradation under UV irradiation and sunlight [27]. In the case of ceria nanoparticles doped with 6% Co, only a slightly lower photodegradation efficiency of MB under sunlight (88.9%) was observed in comparison with the UV-assisted process (98.7%). The hydrothermal method was employed in the synthesis of alkaline metal ion-doped cerium oxide in the form of nanostructures [141]. Mg, Ca, Sr, and Ba were applied as dopants and significantly increased the photocatalytic activity of CeO2, which was proved by the at least two-times higher degradation degree of MB. Similar investigations were carried out with the application of CeO2 doped with Ag, Bi, Cd, and Pb in the form of nanoparticles obtained in a facile one-step precipitation method [142]. Doping resulted in a significantly (at least several times) higher photocatalytic degradation of MB. Among the tested doped ceria, CeO2/Ag exhibited the highest photocatalytic activity, resulting from the highest decrease in the band gap to 1.86 eV, a high stability, a large surface area, and high electrical conductivity.
Fe dopants were introduced to CeO2 in flame spray electrolysis. Fe3+ ions could easily substitute Ce ions or could be introduced to interstices of the crystal lattice of CeO2, leading to the formation of donor or acceptor levels between VB and CB [143]. This resulted in a decrease in the band gap from 3.19 eV (pristine CeO2) to 2.90 eV for Fe-doped CeO2 (5% Fe) and an increase in photocatalytic activity. The hierarchically nanoporous structure of CeO2 was also doped with Fe using facile solvothermal synthesis. Doped CeO2 revealed the band gap energy of 2.63 eV, which was lower than that for the undoped CeO2 (2.92 eV) [144]. Application of the doped CeO2 resulted in a higher photocatalytic degradation of methylene blue under VIS irradiation and proved the higher photocatalytic activity of Fe-doped CeO2. Fe-doped CeO2 synthesized with the yeast-templating method exhibited higher photocatalytic activity in acid orange 7 dye photodegradation in the presence of H2O2, which was attributed to more oxygen vacancies, a larger specific surface area, and a smaller band gap [145].
Yttrium-doped CeO2 in the form of nanorods is also worth mentioning. The photocatalysts were prepared with the application of a hydrothermal method [82]. The doped nanorods exhibited higher photocatalytic activity in dye photodegradation, which was attributed to a lower band gap energy and an increase in oxygen vacancies. In order to enhance the photocatalytic properties of CeO2, yttrium was also doped to a hedgehog-like hierarchical structure using a common hydrothermal process [144]. During the photocatalytic decomposition of acetaldehyde, the highest amount of CO2 was detected while the Y dopant amount was the highest (0.747 mmol). Although Y-doped CeO2 with a hedgehog-like structure exhibited a higher band gap energy (3.35 eV) than pure CeO2 (3.20 eV), it was characterized by a larger specific surface area and higher concentration of oxygen vacancies.
Lanthanum-doped CeO2 was found to have superior catalytic properties due to the replacement of Ce4+ ions with La3+ ions. The reactivity of aliovalent-doped ceria (e.g., La3+, Eu3+, and Sm3+) was reported to be significantly higher in CO oxidation reactions in comparison with isovalent (Zr4+)-doped CeO2, resulting from the enhancement in oxygen vacancies [151]. Therefore, CeO2 nanocrystals heavily doped with La3+ (10, 20, and 50%) were tested in the photocatalytic degradation of methylene blue under visible light irradiation [147]. La-doped CeO2 (10% La3+) presented the highest photocatalytic activity. In this case, the enhancement of CeO2 activity was attributed to an increase in the Ce3+ concentration. However, there is a final limit of Ce4+ to Ce3+ conversion which cannot be enhanced, even after increasing the dopant amount.
The results of the photocatalytic degradation of various pollutants which can be present in industrial wastewater prove the purposefulness of CeO2 doping with metal and non-metal dopants. The doping resulted in the following: (1) The formation of surface defects, which prevented electron-hole recombination or decreased recombination rates; (2) an increase in the surface area and a higher number of sites accessible for the adsorption of pollutants on CeO2 particles; (3) a decrease in the band gap energy, leading to visible light absorption; and (4) higher photocatalytic activity of pollutant degradation.
The amount of doped CeO2 catalyst applied in the photodegradation process usually does not exceed the value of 0.4 g/L, with a few exceptions, such as when this amount was 1 and 3.3 g/L in the case of Co-doped CeO2 applied in acid orange 7 degradation (Table 2). Almost all doped CeO2 photocatalysts were applied in the form of suspension. Only in a few cases were doped photocatalysts immobilized [26,140,144], eliminating a subsequent process of photocatalyst removal.
The duration of the photodegradation process is another parameter which attracts attention. Usually, the duration of the process required to achieve a significant decrease in the pollutant concentration is relatively long—about a couple of hours, with the exception of F-doped CeO2 [136], for which only a 6 min process was enough to remove 91.2% of the methylene blue. However, this was attributed to decolorization of the dye solution, not its mineralization. Total removal of the dye could be confirmed by a significant decrease in total organic carbon (TOC).
Therefore, further development of doped ceria photocatalysts should target not only visible light application in the process, but also a significant decrease in its duration and the elimination of photocatalyst application in the form of a suspension.

3.3. CeO2 Composite Photocatalysts

During the last two decades, various strategies have been employed to improve the photocatalytic activities of photocatalysts applied in the degradation of different pollutants present in industrial wastewater. These strategies include the formation of a semiconductor heterojunction via the combination of a semiconductor with metal or/and other semiconductors [152]. CeO2 is often combined with other semiconductor materials in order to decrease e/h+ pair recombination and enhance the utilization of sunlight in catalyst photoexcitation.
Generally, four types of heterojunction photocatalysts can be distinguished: (1) The semiconductor–semiconductor heterojunction; (2) the semiconductor–metal heterojunction; (3) the semiconductor–carbon group heterojunction; and (4) the multicomponent heterojunction [152]. The heterojunction between two or more semiconductors is the most common. The coupling of appropriate semiconductors results in an alignment of the band energy levels, which leads to the migration of charges and the creation of an electric field, resulting in an extended lifetime of photogenerated holes and electrons. Three combinations of n-type and p-type semiconductors are possible: (1) n–n heterojunctions; (2) p–p heterojunctions; and (3) p–n heterojunctions [153]. A comparison of CB and VB positions in two coupled semiconductors enables the division of conventional heterojunctions into three types with different gaps: A straddling gap (type I); staggered gap (type II); and broken gap (type III) [154,155].
CeO2 heterojunctions with other semiconductors usually belong to n–n and p–n heterojunctions, e.g., CeO2/TiO2 [156] and CeO2/CuO [157], respectively. Furthermore, CeO2 can form Z-scheme heterojunctions, e.g., Mn3O4/CeO2 [158] and g-C3N4/CeO2 [159]. A Z-scheme heterojunction is formed by two semiconductors with a staggered band structure (type II) and an electron acceptor/donor pair [154]. The photoinduced electrons are transferred to less negative CB and the holes to less positive VB in typical type II heterojunctions, resulting in a lower redox ability of electrons and holes that leads to a lower photocatalytic efficiency [160]. In the case of a Z-scheme heterojunction, the photoinduced electrons in less negative CB and holes in less positive VB can recombine at the heterojunction interface, while electrons and holes with a higher redox ability remain on two respective semiconductors [159]. This results in efficient charge separation and a high redox ability, leading to a higher photocatalytic efficiency.
The above-mentioned types of heterojunctions present in the case of CeO2 composites applied in the photodegradation of different pollutants are presented in the next three subsections.

3.3.1. CeO2 Coupled with TiO2

CeO2 is often coupled with TiO2 due to its oxygen storage ability and photocatalytic activity. The presence of Ce4+ and Ce3+ oxidation states results in excellent CeO2 characteristics manifesting themselves in transferring electrons and enhancement of the light absorption capability in near ultraviolet and ultraviolet ranges [161]. CeO2 composites with TiO2 also show a higher thermal stability and higher electrical conductivity [162]. In addition to these advantages, CeO2/TiO2 composites can be further improved considering their photocatalytic activity related to their agglomeration, specific surface area, and mass-transfer limitation of target pollutants [161]. Furthermore, the separation of CeO2/TiO2 composite photocatalysts from the solution after the treatment of pollutants requires the application of an additional post-treatment process, which is not easy and is energy consuming [163,164]. These disadvantages can be overcome by an application of CeO2/TiO2 composites in the form of nanofilms or nanotube arrays, which are characterized by a large surface area and more convenient separation of the suspensions.
TiO2/CeO2 composites can be prepared with the application of various methods, such as sol-gel techniques, magnetron sputtering, flame spray pyrolysis, electrodeposition, the hydrothermal method, etc. [31], which significantly affect their properties. In particular, the position of the bottom CB level and top VB level depends on the preparation method and determines the photocatalytic activity of the composite material. The possible photoexcitation of CeO2/TiO2 composites under UV and VIS irradiation is presented in Figure 4.
When a CeO2/TiO2 composite is irradiated by UV light, then both CeO2 and TiO2 are excited. This leads to the formation of photogenerated holes and electrons in the VB and CB of both semiconductors. Due to less negative CB energy in TiO2, photogenerated electrons in CB of CeO2 are easily transferred to the CB of TiO2 and form superoxide anion radicals (O2•−) which react with pollutant molecules. Simultaneously, photogenerated holes in the VB of TiO2 are moved to the VB of CeO2 and produce OH radicals which also react with the pollutant molecules. The formation of reactive radicals can be described by the following reactions [165,166]:
TiO 2 + h ν   e CB + h VB + ,
Ce 4 + + e CB Ce 3 + ,
Ce 3 + + O 2 Ce 4 + + O 2 ,
H 2 O + h VB + OH + H + ,
O 2 + 4 H + 2 OH .
Ce4+ species scavenge the excited photoelectrons in TiO2 and limit the recombination of h+/e pairs, leading to a higher photocatalytic activity of the composite materials.
When both semiconductors are irradiated with visible light, then CeO2 is predicted to absorb light and generate photoexcited h+/e pairs. The photogenerated electrons are transferred to the CB of TiO2 (Figure 4) [156,167]. This limits the recombination of h+/e pairs and improves the photocatalytic properties of the composite material.
Exemplary applications of CeO2/TiO2 composite materials in the photocatalytic degradation of various pollutants present in industrial wastewater with the main process parameters are presented in Table 3.
CeO2/TiO2 composites belong to n–n heterojunctions and can also be classified as staggered gap (type II) heterojunctions. Usually, the bang gap of the composite material displays a lower energy than TiO2 or CeO2 (Table 3). Nonetheless, in a few cases, composite materials exhibited a larger band gap than pristine semiconductors, even though their photocatalytic activity in the degradation of pollutants was higher [166,175]. The lower band gap energy enables photoexcitation of the composite materials by visible light irradiation.
The dose of CeO2/TiO2 composites applied in the photodegradation of various pollutants did not exceed the value of 1.3 g/L. The composite materials were applied in a form of suspension and their further removal from the solution was necessary. The duration of the photocatalytic processes was still relatively long—a few hours. Furthermore, the photocatalytic efficiency of the processes performed with the application of visible light was clearly lower than in the presence of more energetic UV light. Determination of TOC or the chemical oxygen demand (COD) proved that mineralization of the solution requires a significantly higher duration of the process than a decrease in the pollutant concentration, leading to the formation of intermediate degradation products [167,168]. This fact is especially important in the case of synthetic dye photodegradation, which is often related to the decolorization of dye solution and does not mean its total degradation.

3.3.2. CeO2-Based Carbon Materials

Although CeO2 displays unique photocatalytic activity and UV absorption, its optical properties can still be enhanced. One of the enhancement methods involves a combination of a semiconductor with carbon supporting materials, such as carbon black, carbon nanotubes, graphene, or graphitic carbon nitride (g-C3N4) [176,177,178]. In the case of CeO2, graphene oxide or reduced graphene oxide is often applied as a supporting material.
Graphene containing single layers of sp2-bonded carbon atoms packed into a 2D structure exhibits unique properties, such as a large specific surface area (~2600 m2/g), high adsorption ability, significant mobility of charge carriers (200,000 cm2/(V⋅s) at room temperature), high electrical conductivity (106 s/cm), strong mechanical properties, ability of chemical modification, and electrochemical stability [179,180]. Pure graphene exhibits weak absorption for light (~2.3% light over a broad wavelength range), which makes graphene suitable for some optoelectronic applications, but not suitable for efficiently collecting solar light [181]. It was reported that several graphene-based materials presented some photocatalytic activity due to the presence of defects. The common defects in graphene have been identified as heteroatoms (O, N, B, P, etc.) [182]. Small amounts of metal oxide nanoparticles can form a heterojunction that improves the photocatalytic activity of graphene.
Graphene-based materials are characterized by different contents of oxygen, and are formed in consecutive steps starting from graphite, involving deep chemical oxidation to graphite oxide, exfoliation to graphene oxide (GO), and a final partial reduction to reduced graphene oxide (r-GO) [182,183]. The content of O in GO can vary between 40 and 60%, depending on the conditions of chemical oxidation. Graphene oxide reduced to r-GO with the application of physical and chemical methods is characterized by a high density of defects in comparison with graphene. Graphene is regarded as a 0 band gap semiconductor due to the electronic band overlap, while GO and r-GO exhibit behavior of semiconductor materials [182]. GO is a p-type semiconductor with a lower electron mobility and a band gap with an energy in the range of 2.4 to 4.3 eV [180].
Graphitic carbon nitride (g-C3N3) is another carbon compound which can be combined with CeO2. Its structure is similar to graphite and composed of hexagonal rings of carbon atoms with sp2 hybrid bonds [184]. The g-C3N4 is a metal-free photocatalyst with a good chemical and thermal stability under ambient conditions, characterized by non-toxicity, low costs, and facile preparation [185]. Although its photocatalytic efficiency is comparable to TiO2, g-C3N4 suffers from a high recombination rate of h+/e pairs and low visible light utilization. Its band gap is characterized by the energy of 2.7 eV and the position of CB and VB at −1.4 and +1.3 eV, respectively [185,186]. The position of CB and VB in CeO2 and g-C3N4 indicates that type II (staggered gap) heterojunctions can be formed between these two semiconductors (Figure 5).
The application of CeO2 composites with GO, rGO, and g-C3N4 in the photodegradation of different pollutants is presented in Table 4.
CeO2 composites with graphene (G), GO, or r-GO are mainly prepared with the application of a hydrothermal method, in-situ growth strategy, or two-step wet-chemical route (Table 4). The composites show a lower energy of the band gap than pristine CeO2, facilitating better utilization of visible light. During irradiation, the photogenerated electrons are quickly transferred from CB in CeO2 to VB in graphene materials and react with oxygen dissolved in a solution, resulting in the formation of superoxide radicals (O2•−) according to the following reactions [119]:
CeO 2 + h ν CeO 2 ( h + + e ) ,
CeO 2 ( h + ) + H 2 O CeO 2 + OH + H + ,
CeO 2 ( h + ) + OH CeO 2 + OH ,
CeO 2 ( e ) + G CeO 2 + G ( e ) ,
G ( e ) + O 2 G + O 2 .
The superoxide and hydroxyl radicals can react with pollutant molecules and mineralize solutions with the formation of simple inorganic products, such as CO2, H2O, SO42−, NO3, etc. In this way, the recombination rate of photogenerated h+/e pairs in CeO2 is reduced. CeO2 composites with G, GO, and r-GO are characterized by a significantly higher photodegradation efficiency of different pollutants [119,177,187], but the duration of degradation processes is still long. These composite materials were mainly tested in the photocatalytic degradation of synthetic dyes at low concentrations. In industrial wastewater, the concentration of dyes can be significantly higher. Taking into consideration the fact that the photodegradation efficiency usually decreases with an in increase in the pollutant concentration, the degradation still requires improvement.
CeO2 composites with g-C3N4 are also promising photocatalytic materials with a lower band gap energy [194,195] and significantly higher photocatalytic efficiency in degradation processes [190,191,196]. Considering the position of CB and VB in CeO2 and g-C3N4 (Figure 5), the higher photocatalytic efficiency can be attributed to the transfer of photoexcited electrons and holes between CeO2 and g-C3N4, which suppresses the recombination of photogenerated h+/e pairs. During irradiation, photogenerated electrons on CB in g-C3N4 are transferred to CB in CeO2 and react with O2, while photogenerated holes on VB in CeO2 are transferred to VB in g-C3N4 and react with H2O according to the following reactions [192]:
g C 3 N 4 / CeO 2   + h ν   g   C 3 N 4 ( h + + e ) / CeO 2 ( h + + e ) ,
g   C 3 N 4 ( h + + e ) / CeO 2 ( h + + e ) g C 3 N 4 ( h + + h + ) / CeO 2 ( e + e ) ,
e + O 2 O 2 ,
h + + H 2 O OH + h + .
The superoxide and hydroxyl radicals formed in the above-presented reactions take part in the degradation of pollutants.
In the case of CeO2 composites with graphene substrates (G, GO, and r-GO) and g-C3N4, two problems have still not been resolved. The first one is related to the lower rates of TOC or COD decrease in wastewater in comparison with the degradation rate of pollutants [188,190]. The second one is attributed to the immobilization of a composite photocatalyst, which could eliminate the post-treatment process of photocatalyst removal from the wastewater.

3.3.3. CeO2 Composites with Other Materials

The photodegradation efficiency of CeO2 can be enhanced not only by doping or coupling with carbon materials, but also by forming heterojunctions with photocatalysts other than TiO2. CeO2 heterojunctions include n–n and p–n semiconductor heterojunctions belonging to the typical type II heterojunctions, and to more effective Z-scheme heterojunctions. Ag2O/CeO2 and CuO/CeO2 composites are examples of a p–n heterojunction. In the p–n heterojunction, holes—the predominant charge carriers in the p-type semiconductor—are transferred to the n-type semiconductor, while electrons—the predominant charge carriers in the n-type semiconductor—are transferred in the opposite way, i.e., to the p-type semiconductor [153]. When a p-type semiconductor is in contact with an n-type semiconductor, both sides around the interface become depletion layers. An electric field is formed from the n-side (positively charged) to the p-side (negatively charged) and increases the migration rate of photogenerated electrons and holes between semiconductors [197]. The coupling of a p-type semiconductor with an n-type semiconductor not only decreases the recombination of h+/e pairs, leading to a higher photocatalytic efficiency, but also enables the utilization of visible light. Furthermore, the p–n heterojunctions are more efficient than n–n heterojunctions due to a better separation of h+/e pairs.
Another way to improve CeO2’s photocatalytic activity is to construct a Z-scheme heterojunction. A Z-scheme heterojunction can be classified as a type II heterojunction, according to the position of bands, but with different paths of charge carrier migration. When both semiconductors in a Z-scheme heterojunction are irradiated, the photogenerated electrons in the less negative CB in the second semiconductor may migrate to the less positive VB in the first semiconductor and recombine with photogenerated holes (Figure 6). This process results in charge separation within each semiconductor. Simultaneously, the photogenerated electrons in the more negative CB in the first semiconductor maintain their highest redox potential, while the photogenerated holes in the more positive VB in the second semiconductor also maintain their highest oxidation ability [198].
Three types of Z-scheme heterojunctions can be distinguished, depending on the “mediator” facilitating the transfer of electrons between semiconductors [197]:
(1)
direct Z-scheme—mediator-free;
(2)
Z-scheme with a solid mediator;
(3)
Z-scheme with a redox pair mediator.
CeO2 composites with different photocatalysts were applied in the photodegradation of various pollutants and are presented along with the main process parameters in Table 5.
CeO2 composites with Bi2WO6, Bi2MoO6, BiVO4, Bi4Ti3O12, ZnO, and Ag2S belong to n–n heterojunctions which are prepared with the application of hydrothermal, solvothermal, or precipitation methods (Table 5). CeO2/Bi2WO6 composites obtained in the precipitation method coupled with a hydrothermal route were applied in rhodamine B (RhB) dye degradation. Although the same amount of catalyst was applied and the concentration of dye was the same, the degradation efficiency was 76% [32] and 54% [100], and it was achieved in a process lasting 75 min. The higher efficiency was observed in the case of the composite with the band gap energy of 3.15 eV, while a lower efficiency was achieved with the composite photocatalyst with 2.60 eV. The differences in the degradation efficiency can be attributed to the different contents of the two semiconductors, i.e., CeO2 and BiWO6 in the composite photocatalysts. The method of preparation also plays an important role in the composite photocatalyst activity. CeO2/Bi2WO6 prepared with a doctor blading method exhibited a band gap energy of 2.77 eV [199], but the degradation efficiency of RhB with a ten-fold lower concentration was only 44% in a process lasting 2 h. The lower degradation resulted not only from a different preparation method, but also from immobilization of the composite. The photocatalyst was applied in the form of a 3-layer film controlled by scotch tape, which eliminated the post-treatment process of photocatalyst removal. Nonetheless, the immobilization of a photocatalyst often results in its lower degradation efficiency.
CeO2 composites with ZrO2, Cu2S, CuO, CuBi2O4, Bi2O3, BiOI, Ag2O, and NiO are classified as p–n heterojunctions. CuO enables the utilization of solar energy due to its narrow band gap. It presented a relatively high photodegradation efficiency of 86.2% in RB dye treatment in only a 12 min process under visible light irradiation [157]. The degradation of Congo red dye with the application of CuBi2O4/CeO2 required UV irradiation, in order to achieve a comparable photodegradation efficiency in the process lasting 100 min [215]. The photodegradation of Rh 6G dye was performed under natural sunlight with the application of a Cu2S/CeO2 photocatalyst and resulted in a 44% degradation efficiency during a 4 h process. The results obtained in the presence of Cu2S/CeO2 were compared with the results of the process carried out in the presence of an n–n-type photocatalyst—Ag2S/CeO2. Under the same conditions, the photodegradation of Rh 6G was lower and totaled only 30% [213].
The photodegradation efficiency of a BiOI/CeO2 composite in the degradation process of MO was significantly higher (94%) in a 50 min process [204], while the lower removal of Orange II dye (55%) in the presence of Bi2O3/CeO2 required a process duration of 5 h [203]. However, it is worth noting that the concentration of Orange II was at least 3.5-times higher than the MO concentration and could limit the penetration of visible light inside the solution.
A CeO2/ZrO2 composite photocatalyst was tested in Congo red (CR) photodegradation of Congo red dye under UV, VIS, and sunlight irradiation. This photocatalyst exhibited a comparable degradation efficiency in the process under UV and sunlight irradiation, while the process under VIS irradiation resulted in a slightly higher efficiency (68%) [98]. The authors proved the possibility of simultaneous hydrogen generation.
The Z-scheme heterojunction seems to be an interesting alternative to n–n and p–n heterojunctions. The coupling of CeO2 with Mn3O4, Ag2CO3, AgBr, ZnO, Ag3PO4, or Bi2WO6 can result in the formation of a Z-scheme heterojunction under specific conditions. Z-scheme system photocatalysts with CeO2 were applied under visible light irradiation (Table 5) in the photodegradation of dyes and pharmaceuticals, in the photoreduction of CO2, and in hydrogen photocatalytic generation. The comparison of Z-scheme junction application in photodegradation processes indicates its higher photodegradation efficiency in comparison with pristine photocatalysts due to the better separation of h+/e pairs and the presence of electrons with a higher redox potential and holes with a higher oxidation ability. RhB dye photodegradation proceeded with almost a two-times higher efficiency in the presence of ZnO/CeO2 [158] and Mn3O4/CeO2 photocatalysts [222] in comparison with pristine CeO2, ZnO, and Mn3O4. Levofloxacin, which is a pharmaceutical, and Bisphenol A, which is a compound applied in the production of plastics and epoxy resins, were also photodegraded with a higher efficiency with the application of Ag2CO3/CeO2/AgBr and Tm3+:CeO2/palygorskite [212,223]. A high photodegradation efficiency was achieved in the process lasting about 1 h, except for Tm3+:CeO2/palygorskite and Mn3O4/CeO2 photocatalysts, which required a 3 h process to achieve about a 90% efficiency.
CeO2 composites are not only applied in the photodegradation of organic pollutants. Hollow magnetic microcapsules—CeO2@Bi2WO6—were applied in the photodegradation of Cr(VI) and cyanides [200]. The photocatalyst dose of 0.5 and 1 g/L was enough to achieve a 99.6% removal of Cr(VI) and 98.3% removal of cyanide, respectively, during a process lasting 1 h.
CeO2/Bi2MoO6 was applied in the photoreduction of CO2 into CH3OH and C2H5OH [202]. The yield of CH3OH and C2H5OH was 32.5 and 25.9 µmol/gcatal., respectively. A higher yield (40 and 30 µmol/gcatal.) was achieved in the case of the Ag/Ag3PO4/CeO2 photocatalyst used in CO2 photoreduction lasting 4 h, similar to the case of the previous composite material [209].
Another application of CeO2 composites occurs in the photocatalytic generation of hydrogen with the utilization of renewable resources. A CeO2/CdS composite photocatalyst was applied in the process performed under UV and VIS irradiation [224]. The hydrogen evolution rate was 3.5-times higher when CeO2/CdS was irradiated with visible light (λ > 420) nm). A significantly higher hydrogen evolution rate was observed in a process with the application of CeO2/CdS-DETA (DETA—diethylenetriamine) under visible irradiation [79] and was attributed to large active sites of oxygen vacancies in CeO2, excellent electron transfer between CdS-DETA and CeO2, and anti-photocorrosion. The possibility of hydrogen generation is important because this process is plausible in a performance with the simultaneous degradation of pollutants.
The examples of CeO2 composite materials applied in the photodegradation of various pollutants show a relatively high photocatalytic efficiency estimated as the degradation of a pollutant, but not the mineralization of a solution. Only in a few cases was the mineralization of a pollutant solution presented by the determination of TOC. The photodegradation efficiency of levofloxacin performed with the application of a ternary photocatalyst—Ag2CO3/CeO2/AgBr—resulted in an 88% degradation of the pollutant achieved in a 40 min process, while a 61% decrease in TOC required a process lasting two-times longer [212]. Furthermore, CeO2 composite materials were mainly tested in model solutions of different pollutants and not in real industrial wastewater, with a few exceptions. In the case of ZnO/CeO2 photocatalyst applied in the photodegradation of industrial textile effluent, a 90% decrease in TOC required a 6 h process, while the degradation of MO, MB, and phenol was above 96% in 2.5 h photodegradation [220]. Similarly, the application of CeO2/CuO composite material in the photodegradation of industrial textile effluent resulted in its 86% degradation, determined as a change in UV-VIS absorbance spectra during a 5 h process [214]. These examples show that the photodegradation of pollutant mixtures present in industrial wastewater is more difficult than in a model solution of a single pollutant, and mineralization (not only the removal of a pollutant) requires a significantly higher duration in terms of the treatment process.

3.4. CeO2 as a Photoelectrocatalyst Applied in Pollutant Degradation

Photocatalytic (PC) treatment techniques are tested in the degradation of model solutions of various inorganic and organic pollutants, but also in the case of industrial effluents. As was presented in the previous sections, the photocatalytic degradation of pollutants can be performed with the application of a pristine CeO2 photocatalyst or composite photocatalysts including CeO2. However, PC treatment requires a relatively long process duration in order to achieve a high removal efficiency. Usually, total removal of a pollutant does not mean its total degradation, but only its transformation to other less toxic or sometimes more toxic compounds in photocatalytic treatment. This effect is especially related to the PC degradation of synthetic dyes. Their PC treatment leads to the decolorization of wastewater, but not to complete mineralization. Complete mineralization is achieved when a significant decrease in TOC is observed. A TOC decrease requires a significantly longer duration in terms of the treatment process than decolorization. Furthermore, PC treatment requires an additional post-treatment process for photocatalyst removal. The duration of the process and the necessity of photocatalyst removal are the two main factors leading to the combination of PC with an electrocatalytic (EC) treatment. The combination of these two processes should result in the higher efficiency of pollutant degradation and the elimination of post-treatment removal of the photocatalyst. Photocatalysts can be immobilized on the surface of an electrode—an anode or cathode applied in the process.
Electrocatalytic techniques have been tested in industrial wastewater treatment containing organic and inorganic pollutants in the past several years. EC oxidation of pollutants seems to be more common than electroreduction and can be performed directly or indirectly [225]. The direct electrooxidation of pollutants usually results in a poor efficiency. Therefore, indirect electrooxidation by electrochemically generated oxidants, i.e., reactive oxygen species (ROS), especially hydroxyl radicals, is applied more often [226,227,228]. The EC treatment is also characterized by several disadvantages. The complete degradation of pollutants requires the so-called “deep electrooxidation” with the application of high voltages or current densities, resulting in high treatment costs [229]. Hence, it seems to be purposeful to combine the EC process with PC treatment. The photoelectrocatalytic (PEC) treatment of wastewater can be performed with a photoanode or photocathode—a photocatalyst (semiconductor) immobilized on a conducting support—and the concurrent application of UV or VIS irradiation simultaneously with the bias potential [230,231,232,233,234]. The treatment efficiency can also be increased by the application of commonly used electrode materials and the suspension of photocatalysts in wastewater. However, in this case, the post-treatment removal of photocatalysts is not eliminated.
The PEC treatment of pollutants present in industrial wastewater includes heterogeneous photocatalysis and simultaneous electrocatalysis with a biased potential or current applied to a photoelectrode. The pollutant removal depends on the different radicals formed in both photocatalytic and electrocatalytic processes. Therefore, the efficiency of the photoelectrocatalytic degradation process depends on the operation parameters affecting both photochemical and electrochemical processes. In addition to the photoelectrode type, these parameters include the (1) light source and its intensity, (2) external bias potential or current, (3) solution pH, (4) supporting electrolyte related to the solution conductivity, (5) concentration of a pollutant, (6) type of counter electrode, and (7) design of the photoelectrochemical reactor [102,105,235,236].
The degradation of various pollutants from industrial wastewater observed in a photoelectrocatalytic process can be monitored by control parameters, including [105,235]
(1)
The decolorization degree determined by UV/VIS spectrophotometry:
Decolorization ( % ) = A 0 A t A 0 × 100 ,
where A0 and At represent the initial and final absorbance determined in a treated solution respectively;
(2)
Pollutant removal:
Removal ( % ) = C 0 C t C 0 × 100 ,
where C0 and Ct represent the initial and final concentration of a pollutant, respectively;
(3)
TOC and COD removal:
TOC ( % ) = TOC 0 TOC t TOC c × 100 ,
COD ( % ) = COD 0 COD t COD c × 100 ,
where TOC0 and TOCt represent the initial and final values of total organic carbon, respectively, and COD0 and CODt represent the initial and final values of the chemical oxygen demand, respectively;
(4)
The electrical energy consumption per mass, which constitutes the main part of the operating costs:
E EM = P × t × 10 6 V × ( C 0 C t ) ,
where P is the rated power (kW) of the system; V is the volume of treated wastewater (L); t is the treatment time (h); and C0 and Ct represent the initial and final mass concentrations of a pollutant in mg/L, respectively. The decolorization degree is a parameter applied in the case of colored pollutant solutions, especially dye solution. Sometimes, absorbance values are used in the determination of the pollutant concentration. TOC and COD parameters are applied in the assessment of the mineralization of organic pollutant solution, which means pollutant conversion to CO2. The electrical energy consumption constitutes the main part of the operating costs.
Some operational and control parameters of photoelectrocatalytic processes performed with the application of CeO2 composites applied as photoelectrodes are presented in Table 6.
CeO2 composite materials immobilized on fluorine-doped tin oxide glass (FTO), indium tin oxide glass (ITO), or other substrates can be applied as photoelectrodes in the PEC process of hydrogen evolution. This process is important considering the possible simultaneous degradation of pollutants and H2 generation. Ce/Ce2O3/CeO2/TiO2 TiO2 nanotube arrays (TNAs) and TiO2/CeO2/Ti photoelectrodes were applied in hydrogen generation from different electrolytes under UV and Vis irradiation [109,245]. A higher hydrogen evolution rate was observed at the composite electrodes in contrast to single component electrodes. This effect was attributed to the better charge separation and improved light absorption ability. The CeO2/Cu2O composite (p–n heterojunction) deposited on ITO was also successfully applied in hydrogen evolution from 0.1 M NaOH solution in the PEC process [23] with the hydrogen evolution rate of 3.62 mL/h. The p–n composite—CeO2/CuO—deposited on the Cu substrate was used in CO2 photoelectrocatalytic reduction and methanol generation [246]. In this case, the PEC process exhibited a 4- and 2.4-times higher yield in comparison to the PC and EC process, respectively.
The CeO2 composite materials were also applied in the PEC degradation of different pollutants, even though PEC processes are not as common as PC degradation. Composites of CeO2 with exfoliated graphite (EG) and reduced graphene oxide co-modified with TiO2 were used in the PEC degradation of 2,4-dichlorophenol, bisphenol A, and tetrabromobisphenol A. The PEC degradation of 2,4-dichlorophenol exhibited a clear synergic effect of EC and PC, resulting in a high photodegradation degree of 99% being achieved in a 3 h process [238]. It is worth noting that the TOC decrease was only slight lower and totaled 93%, while such a high decrease in the TOC value observed in the PC process required a significantly longer duration for the PC process, as was presented in the previous section. An RGO-CeO2-TiO2 photoanode was applied in the PEC process combined with a Fenton reaction [239]. The EC process or Fenton reaction as a separate process resulted in a degradation efficiency of bisphenol A that was not higher than 30%. However, the combination of EC, PC, and the Fenton reaction increased the degradation efficiency to 82%. A similar composite photoelectrode was used in tetrabromobisphenol A treatment and the degradation efficiency was even higher (87%) without a Fenton reaction, in a process 20 min shorter than in the previous case [240]. The higher degradation efficiency was attributed to improved hole-electron separation by applying bias potential. Tetracycline was degraded in the PEC process with two novel photoelectrodes—CeO2@α-Fe2O3 and CeO2 QDs/Ag2Se—both representing a Z-scheme heterojunction [242,243]. The Z-scheme band structure in both composite materials promoted the separation and transfer of photogenerated charges and accelerated the surface redox reaction of tetracycline molecules. Furthermore, in the case of the CeO2@α-Fe2O3 photoelectrode, its photocorrosion was inhibited, resulting in an enhancement of its photocatalytic stability. The PEC degradation of methyl orange with the application of the CeO2/TiO2 photoelectrode confirmed that the PC process combined with EC degradation results in a higher efficiency of pollutant removal achieved in a shorter time. A degradation efficiency higher than 98% was achieved during the photoelectrocatalytic oxidation of methyl orange in the process lasting 1 h.
Although PEC processes represent a promising method of pollutant removal from industrial wastewater, they still require improvement in order to increase the stability of photoelectrodes, facilitate natural solar light utilization instead of UV or VIS light irradiation, and decrease process costs.

4. Conclusions and Future Perspectives

An overview of the CeO2 photocatalyst and its composites with other materials applied in the photocatalytic and photoelectrocatalytic treatment of various pollutants present in industrial wastewater has been presented. The characteristics of ceria as a photocatalyst have also been described. Examples of CeO2 and its composite materials with the main experimental conditions and results of pollutant degradation were included in the presentation of the advantages and disadvantages of photocatalytic and photoelectrocatalytic processes.
Photocatalysis and photoelectrocatalysis seem to be promising options for solving environmental pollution and energy shortage issues. Nonetheless, these processes require further enhancements. One of the possible development directions is associated with the application of properly selected photocatalytic or photoelectrocatalytic material. The application of CeO2 in the photocatalytic and photoelectrocatalytic treatment of industrial wastewater requires further investigations, which should focus on the following aspects:
(1)
The development of CeO2 composites which are highly efficient, stable, and visible light- or sunlight-active, and are characterized by a proper band gap structure and energy for redox reactions, a high photostability during long-term utilization, and scalability, making their commercial implementation possible;
(2)
The development of simple and low-cost procedures for CeO2-based photocatalyst manufacturing, with special attention given to morphology control, an increase in the active surface area, and CeO2 immobilization on a substrate in order to remove it from wastewater easily and ensure electrical conductivity in the case of photoelectrodes;
(3)
CeO2 application in combination with the photocatalytic or electrocatalytic treatment of wastewater with simultaneous electricity and hydrogen generation.
During the last two decades, many studies on the possible application of CeO2 and its composites in pollutant treatment have been reported. However, the concentrations of pollutants which have been considered have been relatively low and have not reflected the real concentrations observed in industrial wastewater. An increase in pollutant concentrations usually results in a degradation efficiency decrease. Furthermore, wastewater with a strong color, e.g., containing dyes, hinders the penetration of UV or VIS light irradiation inside solutions. These issues also need attention and problem-solving.
The optimization of photocatalytic and photoelectrocatalytic processes with CeO2 applied in wastewater treatment should involve not only the removal of pollutants, but also a COD and/or TOC decrease in the treated effluent containing organics. Toxicity control and the enhancement of wastewater biodegradability during treatment processes should also be considered.
The coupling of heterogeneous photocatalysis with electrocatalysis presents advantages over single processes. The photoelectrocatalytic treatment of pollutants proceeds with a higher efficiency in comparison to photocatalytic and electrocatalytic degradation performed under similar conditions. The proper choice of composition of CeO2-based photoelectrode materials is very important in order to achieve the optimal degradation of pollutant under mild conditions. The development of these photoelectrode materials should lead to a higher stability and extended lifetime. Furthermore, the strong adherence of CeO2 photocatalysts to conductive supports should be guaranteed.
The composition of industrial wastewater may vary, and some pollutants can poison photocatalysts and photoelectrocatalysts. Therefore, regeneration methods for CeO2-based photocatalyst surfaces must be taken under consideration.
Although there are many issues that should be overcome in photocatalytic and photoelectrocatalytic processes in wastewater treatment before these processes can be commercially implemented, CeO2 and its composites seem to be promising photocatalytic materials.

Funding

This research received no external funding.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Chen, X.; Chen, R.; Zhu, X.; Liao, Q.; Zhang, Y.; Ye, D.; Zhang, B.; Yu, Y.; Li, J. A solar responsive cubic nanosized CuS/Cu2O/Cu photocathode with enhanced photoelectrochemical activity. J. Catal. 2019, 372, 182–192. [Google Scholar] [CrossRef]
  2. Nadarajan, R.; Bakar, W.A.W.A.; Ali, R.; Ismail, R. Photocatalytic degradation of 1,2-dichlorobenzene using immobilized TiO2/SnO2/WO3 photocatalyst under visible light: Application of response surface methodology. Arab. J. Chem. 2018, 11, 34–47. [Google Scholar] [CrossRef] [Green Version]
  3. Trandafilović, L.V.; Jovanović, D.J.; Zhang, X.; Ptasińska, S.; Dramićanin, M.D. Enhanced photocatalytic degradation of methylene blue and methyl orange by ZnO: Eu nanoparticles. Appl. Catal. B Environ. 2017, 203, 740–752. [Google Scholar] [CrossRef] [Green Version]
  4. Intarasuwan, K.; Amornpitoksuk, P.; Suwanboon, S.; Graidist, P. Photocatalytic dye degradation by ZnO nanoparticles prepared from X2C2O4 (X = H, Na and NH4) and the cytotoxicity of the treated dye solutions. Sep. Purif. Technol. 2017, 177, 304–312. [Google Scholar] [CrossRef]
  5. Zhang, A.-Y.; Wang, W.-K.; Pei, D.-N.; Yu, H.-Q. Degradation of refractory pollutants under solar light irradiation by a robust and self-protected ZnO/CdS/TiO2 hybrid photocatalyst. Water. Res. 2016, 92, 78–86. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Liu, X.; Jin, A.L.; Jia, Y.S.; Xia, T.L.; Deng, C.X.; Zhu, M.H.; Chen, C.F.; Chen, X.S. Synergy of adsorption and visible-light photocatalytic degradation of methylene blue by a bifunctional Z-scheme heterojunction of WO3/g-C3N4. Appl. Surf. Sci. 2017, 405, 359–371. [Google Scholar] [CrossRef]
  7. Munoz-Batista, M.J.; Fernandez-Garcia, M.; Kubacka, A. Promotion of CeO2-TiO2 photoactivity by g-C3N4: Ultraviolet and visible light elimination of toluene. Appl. Catal. B Environ. 2015, 164, 261–270. [Google Scholar] [CrossRef]
  8. Munoz-Batista, M.J.; Nasalevich, M.A.; Savenije, T.J.; Kapteijn, F.; Gascon, J.; Kubacka, A.; Fernandez-Garcia, M. Enhancing promoting effects in g-C3N4-Mnn+/CeO2-TiO2 ternary composites: Photo-handling of charge carriers. Appl. Catal. B Environ. 2015, 176–177, 687–698. [Google Scholar] [CrossRef]
  9. Yu, Y.; Zhong, Q.; Cai, W.; Ding, J. Promotional effect of N-doped CeO2 supported CoOx catalysts with enhanced catalytic activity on NO oxidation. J. Mol. Catal. A Chem. 2015, 398, 344–352. [Google Scholar] [CrossRef]
  10. Xiao, F.-X.; Hung, S.-F.; Miao, J.; Wang, H.-Y.; Yang, H.; Liu, B. Metal-cluster-decorated TiO2 nanotube arrays: A composite heterostructure toward versatile photocatalytic and photoelectrochemical applications. Small 2015, 11, 554–567. [Google Scholar] [CrossRef]
  11. Fernandez-Domene, R.M.; Rosello-Marquez, G.; Sanchez-Tovae, R.; Lucas-Granados, B. Photoelectrochemical removal of chlorfenvinphos by using WO3 nanorods: Influence of annealing temperature and operation pH. Sep. Purif. Technol. 2019, 212, 458–464. [Google Scholar] [CrossRef]
  12. Huda, A.; Suman, P.H.; Torquato, L.D.M.; Silva, B.F.; Handoko, C.T.; Gulo, F. Visible light-driven photoelectrocatalytic degradation of acid yellow 17 using Sn3O4 flower-like thin films supported on Ti substrate (Sn3O4/TiO2/Ti). J. Photochem. Photobiol. A 2019, 376, 196–205. [Google Scholar] [CrossRef]
  13. Wang, X.; Wu, Q.; Ma, H.; Ma, C.; Yu, Z.; Fu, Y.; Dong, X. Fabrication of PbO2 tipped Co3O4 nanowires for efficient photoelectrochemical decolorization of dye (reactive brilliant blue KN-R) wastewater. Sol. Energy Mater. Sol. Cells 2019, 191, 381–388. [Google Scholar] [CrossRef]
  14. Goulart, L.A.; Alves, S.A.; Mascaro, L.H. Photoelectrochemical degradation of bisphenol A using Cu doped WO3 electrodes. J. Electroanal. Chem. 2019, 39, 123–133. [Google Scholar] [CrossRef]
  15. Fernandez-Domene, R.M.; Sanchez-Tovar, R.; Lucas-Granados, B.; Munoz-Portero, M.J. Elimination of pesticide atrazine by photoelectrocatalysis using a photoanode based on WO3 nanosheets. Chem. Eng. J. 2018, 350, 1114–1124. [Google Scholar] [CrossRef]
  16. Chang, K.-L.; Sun, Q.; Peng, Y.-P.; Lai, S.-W.; Sung, M.; Huang, C.-Y.; Kuo, H.-W.; Sun, J.; Lin, Y.-C. Cu2O loaded titanate nanotube arrays for simultaneously photoelectrochemical ibuprofen oxidation and hydrogen generation. Chemosphere 2016, 150, 605–614. [Google Scholar] [CrossRef]
  17. Kasinathan, K.; Kennedy, J.; Elayaperumal, M.; Henini, M.; Malik, M. Photodegradation of organic pollutants RhB dye using UV simulated sunlight on ceria based TiO2 nanomaterials for antibacterial applications. Sci. Rep. 2016, 6, 38064. [Google Scholar] [CrossRef] [Green Version]
  18. Yan, H.; Wang, X.; Yao, M.; Yao, X. Band structure design of semiconductors for enhanced photocatalytic activity: The case of TiO2. Prog. Nat. Sci. 2013, 23, 402–407. [Google Scholar] [CrossRef]
  19. Nakata, K.; Fujishima, A. TiO2 Photocatalysis: Design and applications. J. Photochem. Photobiol. C 2012, 13, 169–189. [Google Scholar] [CrossRef]
  20. Kang, X.; Liu, S.; Dai, Z.; He, Y.; Song, X.; Tan, Z. Titanium dioxide: From engineering to applications. Catalysts 2019, 9, 191. [Google Scholar] [CrossRef] [Green Version]
  21. Shen, L.; Xing, Z.; Zou, J.; Li, Z.; Wu, X.; Zhang, Y.; Zhu, Q.; Yang, S.; Zhou, W. Black TiO2 nanobelts/g-C3N4 nanosheets laminated heterojunctions with efficient visible-light-driven photocatalytic performance. Sci. Rep. 2017, 7, 41978. [Google Scholar] [CrossRef] [PubMed]
  22. Sharma, D.; Mehta, B.R. Nanostructured TiO2 thin films sensitized by CeO2 as an inexpensive photoanode for enhanced photoactivity of water oxidation. J. Alloys Compd. 2018, 749, 329–335. [Google Scholar] [CrossRef]
  23. Sharma, D.; Sastsangi, V.R.; Shrivastav, R.; Waghmare, U.V.; Dass, S. Understanding the photoelectrochemical properties of nanostructured CeO2/CuO2 heterojunction photoanode for efficient photoelectrochemical water splitting. Int. J. Hydrog. Energy 2016, 41, 18339–18350. [Google Scholar] [CrossRef]
  24. Yin, D.; Zhao, F.; Zhang, L.; Zhang, X.; Liu, Y.; Zhang, T.; Wu, C.; Chen, D.; Chen, Z. Greatly enhanced photocatalytic activity of semiconductor CeO2 by integrating with upconversion nanocrystals and graphene. RSC Adv. 2016, 6, 103795–103802. [Google Scholar] [CrossRef]
  25. Zhang, C.; Zhang, X.; Wang, Y.; Xie, S.; Liu, Y.; Lu, X.; Tong, Y. Facile electrochemical synthesis of CeO2 hierarchical nanorods and nanowires with excellent photocatalytic activities. New J. Chem. 2014, 38, 2581–2586. [Google Scholar] [CrossRef]
  26. Channei, D.; Inceesungvorn, B.; Wetchakun, N.; Ukritnukun, S.; Nattestad, A.; Chen, J.; Phanichphant, S. Photocatalytic degradation of methyl orange by CeO2 and Fe-doped CeO2 films under visible light irradiation. Sci. Rep. 2014, 4, 5757. [Google Scholar] [CrossRef]
  27. Saranya, J.; Ranjith, K.S.; Saravanan, P.; Mangalaraj, D.; Kumar, R.T.R. Cobalt-doped cerium oxide nanoparticles: Enhanced photocatalytic activity under UV and visible light irradiation. Mater. Sci. Semicond. Process. 2014, 26, 218–224. [Google Scholar] [CrossRef]
  28. Li, R.; Li, C.; Yin, S.; Fu, C.; Satom, T. Synthesis of C-N Co-doped nano-CeO2 and dye degradation under compact fluorescent lamp irradiation. J. Nanosci. Nanotechnol. 2012, 12, 2797–2801. [Google Scholar] [CrossRef]
  29. Mittal, M.; Gupta, A.; Pandey, O.P. Role of oxygen vacancies in Ag/Au doped CeO2 nanoparticles for fast photocatalysis. Sol. Energy 2018, 165, 206–216. [Google Scholar] [CrossRef]
  30. Xiu, Z.; Xing, Z.; Li, Z.; Wu, X.; Yan, X.; Hu, M.; Cao, Y.; Yang, S.; Zhou, W. Ti3+-TiO2/Ce3+-CeO2 nanosheet heterojunctions as efficient visible-light-driven photocatalysts. Mater. Res. Bull. 2018, 100, 191–197. [Google Scholar] [CrossRef]
  31. Qu, X.; Xie, D.; Gao, L.; Du, F. Synthesis and photocatalytic activity of TiO2/CeO2 core-shell nanotubes. Mater. Sci. Semicond. Process. 2014, 26, 657–662. [Google Scholar] [CrossRef]
  32. Issarapanacheewin, S.; Wetchakun, K.; Phanichphant, S.; Kangwansupamokon, W.; Wetchakun, N. A novel CeO2/Bi2WO6 composite with highly enhanced photocatalytic activity. Mater. Lett. 2015, 156, 28–31. [Google Scholar] [CrossRef]
  33. Huang, K.; Li, Y.H.; Liang, C.; Xu, X.; Zhou, Y.F.; Fan, D.Y.; Yang, H.J.; Lang, P.L.; Zhang, R.; Wang, Y.G.; et al. One-step synthesis of reduced graphene oxide-CeO2 nanocubes composites with enhanced catalytic activity. Mater. Lett. 2014, 124, 223–226. [Google Scholar] [CrossRef]
  34. Aboutaleb, W.A.; El-Salamony, R.A. Effect of Fe2O3-CeO2 nanocomposite synthesis method on the Congo red dye photodegradation under visible light irradiation. Mater. Chem. Phys. 2019, 236, 121724. [Google Scholar] [CrossRef]
  35. Wang, B.; Zhu, B.; Yun, S.; Zhang, W.; Xia, C.; Afzal, M.; Cai, Y.; Liu, Y.; Wang, Y.; Wang, H. Fast ionic conduction in semiconductor CeO2-δ electrolyte fuel cells. NPG Asia Mater. 2019, 11, 51. [Google Scholar] [CrossRef] [Green Version]
  36. Wang, W.; Zhu, Q.; Dai, Q.; Wang, X. Fe doped CeO2 nanosheets for catalytic oxidation of 1,2-dichloroethane: Effect of preparation method. Chem. Eng. J. 2017, 307, 1037–1046. [Google Scholar] [CrossRef]
  37. Zhang, N.; Zhang, G.; Chong, S.; Zhao, H.; Huang, T.; Zhu, J. Ultrasonic impregnation of MnO2/CeO2 and its application in catalytic sono-degradation of methyl orange. J. Environ. Manag. 2018, 205, 134–141. [Google Scholar] [CrossRef]
  38. Toro, R.G.; Malandrino, G.; Fragala, I.L. Relationship between the nanostructures and the optical properties of CeO2 thin films. J. Phys. Chem. B 2004, 108, 16357–16364. [Google Scholar] [CrossRef]
  39. Avellaneda, C.O.; Berton, M.A.C.; Bulhões, L.O.S. Optical and electrochemical properties of CeO2 thin film prepared by an alkoxide route. Sol. Energy Mater. Sol. Cells 2008, 92, 240–244. [Google Scholar] [CrossRef]
  40. Wongkaew, A. Effect of cerium oxide and zirconium oxide to activity of catalysts. Chiang Mai J. Sci. 2008, 35, 156–162. [Google Scholar]
  41. Chen, H.-I.; Chang, H.-Y. Synthesis and characterization of nanocrystalline cerium oxide powders by two-stage non-isothermal precipitation. Solid State Commun. 2005, 133, 593–598. [Google Scholar] [CrossRef]
  42. Munoz-Batista, M.J.; Gomez-Cerezo, M.N.; Kubacka, A.; Tudela, D.; Fernandez-Garcia, M. Role of interface contact in CeO2-TiO2 photocatalytic composite materials. ACS Catal. 2014, 4, 63–72. [Google Scholar] [CrossRef]
  43. Lei, W.; Zhang, T.; Gu, L.; Liu, P.; Rodriquez, J.A.; Liu, G.; Liu, M. Surface-structure sensitivity of CeO2 nanocrystals in photocatalysis and enhancing the reactivity with nanogold. ACS Catal. 2015, 5, 4385–4393. [Google Scholar] [CrossRef]
  44. Ramezanzadeh, B.; Bahlakeh, G.; Ramezanzadeh, M. Polyaniline-cerium oxide (PAni-CeO2) coated graphene oxide for enhancement of epoxy coating corrosion protection performance on mild steel. Corros. Sci. 2018, 137, 111–126. [Google Scholar] [CrossRef]
  45. Zhang, W.; Wang, H.; Lv, C.; Chen, X.; Zhao, Z.; Qu, Y.; Zhu, Y. Effects of CeO2 geometry on corrosion resistance of epoxy coatings. Surf. Eng. 2020, 36, 175–183. [Google Scholar] [CrossRef]
  46. Younis, A.; Chu, D.; Li, S. Cerium oxide nanostructures and their applications. Chapter 3. In Functionalized Nanomaterials; Farukh, M.A., Ed.; IntechOpen: London, UK, 2016; pp. 53–68. [Google Scholar]
  47. Melchionna, M.; Fornasiero, P. The role of ceria-based nanostructured materials in energy applications. Mater. Today 2014, 17, 349–357. [Google Scholar] [CrossRef]
  48. He, Q. Experimental study on polishing performance of CeO2 and nano-SiO2 mixed abrasive. Appl. Nanosci. 2018, 8, 163–171. [Google Scholar] [CrossRef] [Green Version]
  49. Porqueras, I.; Person, C.; Corbella, C.; Vives, M.; Pinyol, A.; Bertran, E. Characteristics of e-beam deposited electrochromic CeO2 thin films. Solid State Ion. 2003, 165, 131–137. [Google Scholar] [CrossRef]
  50. Parwaiz, S.; Khan, M.M.; Pradhan, D. CeO2-based nanocomposites: An advanced alternative to TiO2 and ZnO2 in sunscreens. Mater. Express 2019, 9, 185–202. [Google Scholar] [CrossRef]
  51. Li, Y.; Hou, X.; Yang, C.; Pang, Y.; Li, X.; Jiang, G.; Liu, Y. Photoprotection of cerium oxide nanoparticles against UVA radiation-induced senescence of human skin fibroblasts due to their antioxidant properties. Sci. Rep. 2019, 9, 2595. [Google Scholar] [CrossRef] [Green Version]
  52. Pislaru-Danescu, L.; Telipan, G.; Ion, I.; Marinescu, V. Prototyping a gas sensors using CeO2 as a matrix or dopant in oxide semiconductor systems. Introductory Chapter; In Cerium Oxide—Applications and Attributes; Khan, S.B., Akhtar, K., Eds.; IntechOpen: London, UK, 2019; pp. 1–34. [Google Scholar]
  53. Hu, J.; Sun, Y.; Xue, Y.; Zhang, M.; Li, P.; Lian, K.; Zhuiykov, S.; Zhang, W.; Chen, Y. Highly sensitive and ultra-fast gas sensor based on CeO2-loaded In2O3 hollow spheres for ppb-level hydrogen detection. Sens. Actuators B Chem. 2018, 257, 124–135. [Google Scholar] [CrossRef]
  54. Li, C.; Lu, Z. Effect of CeO2 and CuO sintering additives on the properties of KNN-based piezoelectric ceramics. Mater. Technol. 2018, 33, 474–478. [Google Scholar] [CrossRef]
  55. Souza, J.V.C.; de Macedo Silva, O.M.; do Carmo Andrade Nono, M.; Machado, J.P.B.; Pimenta, M.; Ribeiro, M.V. S3N4 ceramics cutting tool sintered with CeO2 and Al2O3 additives with AlCrN coating. Mater. Res. 2011, 14, 514–518. [Google Scholar] [CrossRef] [Green Version]
  56. Gong, Q.; Gao, T.; Huang, H.; Wang, R.; Cao, P.; Zhou, G. Double-shelled CeO2@C hollow nanospheres as enhanced anode materials for lithium-ion batteries. Inorg. Chem. Front. 2018, 5, 3197–3204. [Google Scholar] [CrossRef]
  57. Di Monte, R.; Kaspar, J. On the role of oxygen storage in three-way catalysis. Top. Catal. 2004, 28, 47–57. [Google Scholar] [CrossRef]
  58. Machida, M.; Fujiwara, A.; Yoshida, H.; Ohyama, J.; Asakura, H.; Hosokawa, S.; Tanaka, T.; Haneda, M.; Tomita, A.; Miki, T.; et al. Deactivation mechanism of Pd/CeO2-ZrO2 three-way Catalysts analysed by chassis-dynamometer tests and in situ diffuse reflectance spectroscopy. ACS Catal. 2019, 9, 6415–6424. [Google Scholar] [CrossRef]
  59. Jiang, L.; Wei, M.; Xu, X.; Lin, Y.; Lu, Z.; Song, J.; Duan, X. SOx oxidation and adsorption by CeO2/MgO: Synergistic effect between CeO2 and MgO in the fluid catalytic cracking process. Ind. Eng. Chem. Res. 2011, 50, 4398–4404. [Google Scholar] [CrossRef]
  60. Ramamoorthy, R.; Dutta, P.K.; Akbar, S.A. Oxygen sensors: Materials, methods, designs and applications. J. Mater. Sci. 2003, 38, 4271–4282. [Google Scholar] [CrossRef]
  61. Wu, T.; Vegge, T.; Hansen, H.A. Improved electrocatalytic water splitting reaction on CeO2(111) by strain engineering: A DFT+U study. ACS Catal. 2019, 9, 4853–4861. [Google Scholar] [CrossRef]
  62. Umale, S.V.; Tambat, S.N.; Sontakke, S.M. Combustion synthesized CeO2 as an anodic material in dye sensitized solar cells. Mater. Res. Bull. 2017, 94, 483–488. [Google Scholar] [CrossRef]
  63. Chiu, F.-C.; Lai, C.-M. Optical and electrical characterizations cerium oxide thin films. J. Phys. D Appl. Phys. 2010, 43, 075104. [Google Scholar] [CrossRef]
  64. Skorodumova, N.V.; Ahuja, R.; Simak, S.I.; Abrikosov, I.A.; Johansson, B.; Lundqvist, B.I. Electronic, bonding, and optical properties of CeO2 and Ce2O3 from first principles. Phys. Rev. B 2001, 64, 115108. [Google Scholar] [CrossRef]
  65. Schwarz, K. Materials design of solid electrolytes. Proc. Natl. Acad. Sci. USA 2006, 103, 3497. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  66. Zhang, D.; Du, X.; Shi, L.; Gao, R. Shape-controlled synthesis and catalytic application of ceria nanomaterials. Dalton Trans. 2012, 41, 14455–14475. [Google Scholar] [CrossRef]
  67. Sun, C.; Li, H.; Chen, L. Nanostructured ceria-based materials: Synthesis, properties, and applications. Energy Environ. Sci. 2012, 5, 8475–8505. [Google Scholar] [CrossRef]
  68. Zhang, J.; Kumagai, H.; Yamamura, K.; Ohara, S.; Takami, S.; Morikawa, A.; Shinjoh, H.; Kaneko, K.; Adschiri, T.; Suda, A. Extra-low-temperature oxygen storage capacity of CeO2 nanocrystals with cubic facets. Nano Lett. 2011, 11, 361–364. [Google Scholar] [CrossRef]
  69. Charbgoo, F.; Ramezani, M.; Darroudi, M. Bios-sensing applications of cerium oxide nanoparticles: Advantages and disadvantages. Biosens. Bioelectron. 2017, 96, 33–43. [Google Scholar] [CrossRef]
  70. Charbgoo, F.; Ahmad, M.B.; Darroudi, M. Cerium oxide nanoparticles: Green synthesis and biological applications. Int. J. Nanomed. 2017, 12, 1401–1413. [Google Scholar] [CrossRef] [Green Version]
  71. Campbell, C.T.; Peden, C.H.F. Oxygen vacancies and catalysis on ceria surfaces. Science 2005, 309, 713–714. [Google Scholar] [CrossRef]
  72. Gogoi, A.; Navgire, M.; Sarma, K.C.; Gogoi, P. Fe3O3-CeO2 metal oxide nanocomposite as a Fenton-like heterogeneous catalyst for degradation of catechol. Chem. Eng. J. 2017, 311, 153–162. [Google Scholar] [CrossRef]
  73. Wu, Z.; Li, M.; Overbury, S.H. On the structure dependence of CO oxidation over CeO2 nanocrystals with well-defined surface planes. J. Catal. 2012, 285, 61–73. [Google Scholar] [CrossRef]
  74. Sayle, T.X.T.; Cantoni, M.; Bhatta, U.M.; Parker, S.C.; Hall, S.R.; Möbus, G.; Molinari, M.; Reid, D.; Seal, S.; Sayle, D.C. Strain and architecture-tuned reactivity in ceria nanostructures; Enhanced Catalytic oxidation of CO to CO2. Chem. Mater. 2012, 24, 1811–1821. [Google Scholar] [CrossRef]
  75. Liu, L.; Cao, Y.; Sun, W.; Yao, Z.; Liu, B.; Gao, F.; Dong, L. Morphology and nanosize effects of ceria from different precursors on the activity for NO reduction. Catal. Today 2011, 175, 48–54. [Google Scholar] [CrossRef]
  76. Fernandez-Garcia, S.; Jiang, L.; Tinoco, M.; Hungria, A.B.; Han, J.; Blanco, G.; Calvino, J.J.; Chen, X. Enhanced hydroxyl radical scavenging activity by doping lanthanum in ceria nanocubes. J. Phys. Chem. C 2016, 120, 1891–1901. [Google Scholar] [CrossRef]
  77. Doornkamp, C.; Ponec, V. The universal character of the Mars and Van Krevelen mechanism. J. Mol. Catal. A Chem. 2000, 162, 19–32. [Google Scholar] [CrossRef]
  78. Mora-Briseno, P.; Jimenez-Garcia, G.; Castillo-Araiza, C.O.; Gonzales-Rodriguez, H.; Huirache-Acuna, R.; Maya-Yescas, R. Mars van Krevelen Mechanism for the selective partial oxidation of ethane. Int. J. Chem. React. Eng. 2018, 7, 20180085. [Google Scholar] [CrossRef]
  79. Li, Z.; Zhang, J.; Lv, J.; Lu, L.; Liang, C.; Dai, K. Sustainable synthesis of CeO2/CdS-diethylenetriamine composites for enhanced photocatalytic hydrogen evolution under visible light. J. Alloys Compd. 2018, 758, 162–170. [Google Scholar] [CrossRef]
  80. Wang, J.; Liu, P.; Fu, X.; Li, Z.; Han, W.; Wang, X. Relationship between oxygen defects and the photocatalytic property of ZnO nanocrystals in Nafion membranes. Langmuir 2009, 25, 1218–1223. [Google Scholar] [CrossRef]
  81. Tachikawa, T.; Yamashita, S.; Majima, T. Evidence for crystal-face-dependent TiO2 photocatalysis from single-molecule imaging and kinetic analysis. J. Am. Chem. Soc. 2011, 133, 7197–7204. [Google Scholar] [CrossRef]
  82. Liyanage, A.D.; Perera, S.D.; Tan, K.; Chabal, Y.; Balkus, K.J. Synthesis, characterization, and photocatalytic activity of Y-doped CeO2 nanorods. ACS Catal. 2014, 4, 577–584. [Google Scholar] [CrossRef]
  83. Chen, H.; Zhang, C.; Pang, Y.; Shen, Q.; Yu, Y.; Su, Y.; Wang, J.; Zhang, F.; Yang, H. Oxygen vacancy regulation in Nb-doped Bi2WO6 for enhanced visible light photocatalytic activity. RSC Adv. 2019, 9, 22559–22566. [Google Scholar] [CrossRef] [Green Version]
  84. Pu, Z.-Y.; Liu, X.-S.; Jia, A.-P.; Xie, Y.-L.; Lu, J.-Q.; Luo, M.-F. Enhanced activity of CO oxidation over Pr- and Cu-doped CeO2 catalysts: Effect of oxygen vacancies. J. Phys. Chem. C 2008, 12, 15045–15051. [Google Scholar] [CrossRef]
  85. Heydari, H.; Elahifard, M.; Behjatmanesh-Ardakani, R. Role of oxygen vacancy in the adsorption and dissociation of the water molecule on the surfaces of pure and Ni-doped rutile (110): A periodic full-potential DFT study. Surf. Sci. 2019, 679, 218–224. [Google Scholar] [CrossRef]
  86. Varilla, L.A.A.; Seriani, N.; Montoya, J.A. Molecular adsorption and dissociation of CO2 on TiO2 anatase (001) activated by oxygen vacancies. J. Mol. Model. 2019, 25, 231. [Google Scholar] [CrossRef]
  87. Kumar, S.; Kumar, A. Enhanced photocatalytic activity of rGO-CeO2 nanocomposites driven by sunlight. Mater. Sci. Eng. B 2017, 223, 98–108. [Google Scholar] [CrossRef]
  88. Özer, N. Optical properties and electrochromic characterization of sol-gel deposited ceria films. Sol. Energy Mater. Sol. Cells 2001, 68, 391–400. [Google Scholar] [CrossRef]
  89. Corma, A.; Atienzar, P.; Garcia, H.; Chane-Ching, J.-Y. Hierarchically mesostructured doped CeO2 with potential for solar-cell use. Nat. Mater. 2004, 3, 394–397. [Google Scholar] [CrossRef]
  90. Schmitt, R.; Nenning, A.; Kraynis, O.; Korobko, R.; Frenkel, A.I.; Lubomirsky, I.; Haile, S.M.; Rupp, J.L.M. A review of defect structure and chemistry in ceria and its solid solutions. Chem. Soc. Rev. 2019. [Google Scholar] [CrossRef] [Green Version]
  91. Castleton, C.W.M.; Kullgren, J.; Hermansson, K. Tuning LDA+U for electron localization and structure at oxygen vacancies in ceria. J. Chem. Phys. 2007, 127, 244704. [Google Scholar] [CrossRef]
  92. Keating, P.R.L.; Scanlon, K.D.O.; Morgan, B.J.; Galea, N.M.; Watson, G.W. Analysis of intrinsic defects in CeO2 using a Koopmans-like GGA+U approach. J. Phys. Chem. C 2012, 116, 2443–2452. [Google Scholar] [CrossRef]
  93. Song, Y.Q.; Zhang, W.W.; Yang, Q.H.; Liu, Y.L.; Li, Y.X.; Shah, L.R.; Zhu, H.; Xiao, J.Q. Electronic structure and magnetic properties of Co-doped CeO2: Based on first principle calculation. J. Phys. Condens. Matter 2009, 21, 125504. [Google Scholar] [CrossRef] [PubMed]
  94. Zacherle, T.; Schriever, A.; De Souza, R.A.; Martin, M. Ab initio analysis of the defect structure of ceria. Phys. Rev. B 2013, 87, 134104. [Google Scholar] [CrossRef]
  95. Nolan, M.; Grigoleit, S.; Sayle, D.C.; Parker, S.C.; Watson, G.W. Density functional theory studies of the structure and electronic structure of pure and defective low index surfaces of ceria. Surf. Sci. 2005, 576, 217–229. [Google Scholar] [CrossRef]
  96. Younis, A.; Chu, D.; Kaneti, Y.V.; Li, S. Tuning the surface oxygen concentration of {111} surrounded ceria nanocrystals for enhanced photocatalytic activities. Nanoscale 2016, 8, 378–387. [Google Scholar] [CrossRef] [PubMed]
  97. Raudonyte-Svirbutaviciene, E.; Neagu, A.; Vickackaite, V.; Jasulaitiene, V.; Zarkov, A.; Tai, C.-W.; Katelnikovas, A. Two-step photochemical inorganic approach to the synthesis of Ag-CeO2 nanoheterostructures and their photocatalytic activity on tributyltin degradation. J. Photochem. Photobiol. A 2018, 351, 29–41. [Google Scholar] [CrossRef]
  98. Hao, Y.; Li, L.; Zhang, J.; Luo, H.; Zhang, X.; Chen, E. Multilayer and open structure of dendritic crosslinked CeO2-ZrO2 composite: Enhanced photocatalytic degradation and water splitting performance. Int. J. Hydrog. Energy 2017, 42, 5916–5929. [Google Scholar] [CrossRef]
  99. Li, H.; Cui, Y.; Hong, W. High photocatalytic performance of BiOI/Bi2WO3 toward toluene and Reactive Brilliant Red. Appl. Surf. Sci. 2013, 264, 581–588. [Google Scholar] [CrossRef]
  100. Issarapanacheewin, S.; Wetchakun, K.; Phanichphant, S.; Kangwansupamonkon, W.; Wetchakun, N. Photodegradation of organic dyes by CeO2/Bi2WO6 nanocomposite and its physicochemical properties investigation. Ceram. Int. 2016, 42, 16007–16016. [Google Scholar] [CrossRef]
  101. Berger, T.; Monllor-Satoca, D.; Jankulovska, M.; Lana-Villarreal, T.; Gomez, R. The electrochemistry of nanostructured titanium dioxide electrodes. ChemPhysChem 2012, 13, 2824–2875. [Google Scholar] [CrossRef]
  102. Kusmierek, E. Semiconductor electrode materials applied in photoelectrocatalytic wastewater treatment—An overview. Catalysts 2020, 10, 439. [Google Scholar] [CrossRef] [Green Version]
  103. Peter, L.M. Photoelectrochemistry from basic principles to photocatalysis. Chapter 1. In Photocatalysis: Fundamentals and Perspectives; Schneider, J., Bahnemann, D., Ye, J., Puma, G.L., Dionysiou, D.D., Eds.; Royal Society of Chemistry: London, UK, 2016; pp. 3–28. [Google Scholar]
  104. Guijarro, N.; Prevot, M.S.; Sivula, K. Surface modification of semiconductor photoelectrodes. Phys. Chem. Chem. Phys. 2015, 17, 15655–15674. [Google Scholar] [CrossRef] [PubMed]
  105. Bessegato, G.G.; Guaraldo, T.T.; de Brito, J.F.; Brugnera, M.F.; Zanoni, M.V.B. Achievements and trends in photoelectrocatalysis: From environmental to energy applications. Electrocatalysis 2015, 6, 415–441. [Google Scholar] [CrossRef] [Green Version]
  106. Bessegato, G.G.; Guaraldo, T.T.; Zanoni, M.V.B. Enhancement of photoelectrocatalysis efficiency by using nanostructured electrodes. Chapter 10. In Modern Electrochemical Methods in Nano, Surface and Corrosion Science; Aliokhazraei, M., Ed.; IntechOpen: London, UK, 2014; pp. 271–319. [Google Scholar]
  107. Brugnera, M.F.; de Araujo Souza, B.C.; Zanoni, M.V.B. Advanced oxidation processes applied to actinobacterium disinfection. Chapter 15. In Antibacteria—Basics and Biotechnological Applications; Dhanasekaran, D., Ed.; IntechOpen: London, UK, 2016; pp. 353–376. [Google Scholar]
  108. Paramasivam, I.; Jha, H.; Liu, N.; Schmuku, P. A review of photocatalysis using self-organized TiO2 nanotubes and other ordered oxide nanostructure. Small 2012, 8, 3073–3103. [Google Scholar] [CrossRef]
  109. Tan, Y.; Zhang, S.; Shi, R.; Wang, W.; Liang, K. Visible light active Ce/Ce2O/CeO2/TiO2 nanotube arrays for efficient hydrogen production by photoelectrochemical water splitting. Int. J. Hydrog. Energy 2016, 41, 5437–5444. [Google Scholar] [CrossRef]
  110. Chaudhary, P.; Ingole, P.P. In-situ solid-state synthesis of 2D/2D interface between Ni/NiO hexagonal nanosheets supported on g-C3N4 for enhanced photo-electrochemical water splitting. Int. J. Hydrog. Energy 2020, 45, 16060–16079. [Google Scholar] [CrossRef]
  111. Anwer, H.; Mahmood, A.; Lee, J.; Kim, K.-H.; Park, J.-W.; Yip, A.C.K. Photocatalysts for degradation of dyes in industrial effluents: Opportunities and challenges. Nano Res. 2019, 12, 955–972. [Google Scholar] [CrossRef]
  112. Zangeneh, H.; Zinatizadeh, A.A.L.; Habibi, M.; Akia, M.; Isa, M.H. Photocatalytic oxidation of organic dyes and pollutants in wastewater using different modified titanium dioxides: A comparative study. J. Ind. Eng. Chem. 2015, 26, 1–36. [Google Scholar] [CrossRef]
  113. Kuhn, H.J.; Braslavsky, S.E.; Schmidt, R. Chemical actinometry (IUPAC Technical Report). Pure Appl. Chem. 2004, 76, 2105–2146. [Google Scholar] [CrossRef]
  114. Sun, L.; Bolton, J.R. Determination of the quantum yield for the photochemical generation of hydroxyl radicals in TiO2 suspension. J. Phys. Chem. 1996, 100, 4127–4134. [Google Scholar] [CrossRef]
  115. Ling, L.; Tugaoen, H.; Brame, J.; Sinha, S.; Li, C.; Schoepf, J.; Hristovski, K.; Kim, J.-H.; Shang, C.; Westerhoff, P. Coupling light emitting diodes with photocatalyst-coated optical fibers improves quantum yield of pollutant oxidation. Environ. Sci. Technol. 2017, 51, 13319–13326. [Google Scholar] [CrossRef]
  116. Anwer, H.; Park, J.-W. Near-infrared to visible photon transition by upconverting NaYF4: Yb3+, Gd3+, Tm3+ @ Bi2WO6 core@ shell composite for bisphenol A degradation in solar light. Appl. Catal. B 2019, 243, 438–447. [Google Scholar] [CrossRef]
  117. Goharshadi, E.K.; Samiee, S.; Nancarrow, P. Fabrication of cerium oxide nanoparticles: Characterization and optical properties. J. Colloid Interface Sci. 2011, 356, 473–480. [Google Scholar] [CrossRef] [PubMed]
  118. Dong, B.; Li, L.; Dong, Z.; Xu, R.; Wu, Y. Fabrication of CeO2 nanorods for enhanced solar photocatalysts. Int. J. Hydrog. Energy 2018, 43, 5275–5282. [Google Scholar] [CrossRef]
  119. Khan, M.E.; Khan, M.M.; Cho, M.H. Ce3+-ion, surface oxygen vacancy, and visible light-induced photocatalytic dye degradation and photocapacitive performance of CeO2-graphene nanostructures. Sci. Rep. 2017, 7, 5928. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  120. Wang, J.; Han, A.; Jaenicke, S.; Chuah, G.-K. Advances in sorbents and photocatalytic materials for water remediation. Chapter 6. In New and Future Developments in Catalysis—Catalysis for Remediation and Environmental Concerns; Suib, S.L., Ed.; Elsevier: Amsterdam, The Netherlands, 2013; pp. 127–153. [Google Scholar]
  121. Capodaglio, A.G. Contaminants of emerging concern removal by high-energy oxidation-reduction processes: State of the art. Appl. Sci. 2019, 9, 4562. [Google Scholar] [CrossRef] [Green Version]
  122. Armstrong, D.A.; Huie, R.E.; Lymar, S.; Koppenol, W.H.; Merenyi, G.; Neta, P.; Stanbury, D.M.; Steenken, S.; Wardman, P. Standard electrode potentials involving radicals in aqueous solution: Inorganic radicals (IUPAC Technical Report). Pure Appl. Chem. 2015, 87, 1139–1150. [Google Scholar] [CrossRef]
  123. Zamiri, R.; Ahangar, H.A.; Kaushal, A.; Zakaria, A.; Zamiri, G.; Tobaldi, D.; Ferreira, J.M.F. Dielectrical properties of CeO2 nanoparticles at different temperatures. PLoS ONE 2015, 10, e0122989. [Google Scholar]
  124. Renu, G.; Rani, V.V.D.; Nair, S.V.; Subramanian, K.R.V.; Lakshmanan, V.-K. Development of cerium oxide nanoparticles and its cytotoxicity in prostate cancer cells. Adv. Sci. Lett. 2012, 5, 1–9. [Google Scholar] [CrossRef]
  125. Tarnuzzer, R.W.; Colon, J.; Patil, S.; Seal, S. Vacancy engineered ceria nanostructures for protection from radiation-induced cellular damage. Nano Lett. 2005, 5, 2573–2577. [Google Scholar] [CrossRef] [PubMed]
  126. Phuruangrat, A.; Thongtem, S.; Thongtem, T. Microwave-assisted hydrothermal synthesis and characterization of CeO2 nanowires for using as a photocatalytic material. Mater. Lett. 2017, 196, 61–63. [Google Scholar] [CrossRef]
  127. Ji, P.; Zhang, J.; Chen, F.; Anpo, M. Study of adsorption and degradation of acid orange 7 on the surface of CeO2 under visible light irradiation. Appl. Catal. B Environ. 2009, 85, 148–154. [Google Scholar] [CrossRef]
  128. Ji, P.; Tian, B.; Chen, F.; Zhang, J. CeO2 mediated photocatalytic degradation studies of C.I. acid orange 7. Environ. Technol. 2012, 33, 467–472. [Google Scholar] [CrossRef] [PubMed]
  129. Tambat, S.; Umale, S.; Sontakke, S. Photocatalytic degradation of Milling Yellow dye using sol-gel synthesized CeO2. Mater. Res. Bull. 2016, 76, 466–472. [Google Scholar] [CrossRef]
  130. Aslam, M.; Qamar, M.T.; Soomro, M.T.; Ismail, I.M.I.; Salah, N.; Almeelbi, T.; Gondal, M.A.; Hameed, A. The effect of sunlight induced surface defects on the photocatalytic activity of nanosized CeO2 for the degradation of phenol and its derivatives. Appl. Catal. B Environ. 2016, 180, 391–402. [Google Scholar] [CrossRef]
  131. Khan, M.M.; Ansari, S.A.; Pradhan, D.; Han, D.H.; Lee, J.; Cho, M.H. Defect-induced band gap narrowed CeO2 nanostructures for visible light activities. Ind. Eng. Chem. Res. 2014, 53, 9754–9763. [Google Scholar] [CrossRef]
  132. Zheng, X.; Huang, S.; Yang, D.; Zhai, H.; You, Y.; Fu, X.; Yuan, J.; Zhou, X.; Wen, J.; Liu, Y. Synthesis of X-architecture CeO2 for the photodegradation of methylene blue under UV-light irradiation. J. Alloys Compd. 2017, 705, 131–137. [Google Scholar] [CrossRef]
  133. Amoresi, R.A.C.; Oliveira, R.C.; Marana, N.L.; de Almeida, P.B.; Prata, P.S.; Zaghete, M.A.; Longo, E.; Sambrano, J.R.; Simoes, A.Z. CeO2 nanoparticle morphologies and their corresponding crystalline planes for the photocatalytic degradation of organic pollutants. ACS Appl. Nano Mater. 2019, 2, 6513–6526. [Google Scholar] [CrossRef]
  134. Channei, D.; Nakaruk, A.; Phanichphant, S. Photocatalytic degradation of dye using CeO2/SCB composite catalysts. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2017, 183, 218–224. [Google Scholar] [CrossRef]
  135. Yang, K.; Huang, W.-Q.; Xu, L.; Luo, K.-W.; Yang, Y.-C.; Huang, G.-F. Insights into enhanced visible-light photocatalytic activity of CeO2 doped with nonmetal impurity from the first principles. Mater. Sci. Semicond. Process. 2016, 41, 200–208. [Google Scholar] [CrossRef]
  136. Miao, H.; Huang, G.-F.; Liu, J.-H.; Zhou, B.-X.; Pan, A.; Huang, W.-Q.; Huang, G.-F. Origin of enhanced photocatalytic activity of F-doped CeO2 nanocubes. Appl. Surf. Sci. 2016, 370, 427–432. [Google Scholar] [CrossRef]
  137. Wu, C. Solvothermal synthesis of N-doped CeO2 microspheres with visible light-driven photocatalytic activity. Mater. Lett. 2015, 139, 383–384. [Google Scholar] [CrossRef]
  138. Mao, C.; Zhao, Y.; Qiu, X.; Zhu, J.; Burda, C. Synthesis, characterization and computational study of nitrogen-doped CeO2 nanoparticles with visible-light activity. Phys. Chem. Chem. Phys. 2008, 10, 5633–5638. [Google Scholar] [CrossRef]
  139. Akbari-Fakhrabai, A.; Saravanan, R.; Jamshidijam, M.; Mangalaraja, R.V.; Gracia, M.A. Preparation of nanosized yttrium doped CeO2 catalyst used for photocatalytic application. J. Saudi Chem. Soc. 2015, 19, 505–510. [Google Scholar] [CrossRef] [Green Version]
  140. Xu, B.; Zhang, Q.; Yuan, S.; Zhang, M.; Ohno, T. Morphology control and photocatalytic characterization of yttrium-doped hedgehog-like CeO2. Appl. Catal. B Environ. 2015, 164, 120–127. [Google Scholar] [CrossRef] [Green Version]
  141. Murugan, R.; Kashinath, L.; Subash, R.; Sakhivel, P.; Byrappa, K.; Rajendran, S.; Ravi, G. Pure and alkaline metal ion (Mg, Ca, Sr, Ba) doped cerium oxide nanostructures for photo degradation of methylene blue. Mater. Res. Bull. 2018, 97, 319–325. [Google Scholar] [CrossRef]
  142. Murugadoss, G.; Ma, J.; Ning, X.; Kumar, M.R. Selective metal ions doped CeO2 nanoparticles for excellent photocatalytic activity under sun light and supercapacitor application. Inorg. Chem. Commun. 2019, 109, 107577. [Google Scholar] [CrossRef]
  143. Channei, D.; Inceesungvorn, B.; Wetchakun, N.; Phanichphant, S.; Nakaruk, A.; Koshy, P.; Sorrell, C.C. Photocatalytic activity under visible light of Fe-doped CeO2 nanoparticles synthesized by flame spray pyrolysis. Ceram. Int. 2013, 39, 3129–3134. [Google Scholar] [CrossRef]
  144. Arul, N.S.; Mangalaraj, D.; Han, J.I. Enhanced photocatalytic property of self-assembled Fe-doped CeO2 hierarchical nanostructures. Mater. Lett. 2015, 145, 189–192. [Google Scholar] [CrossRef]
  145. Zhao, B.; Shao, Q.; Hao, L.; Zhang, L.; Liu, Z.; Zhang, B.; Ge, S.; Guo, Z. Yeast-template synthesized Fe-doped cerium oxide hollow microspheres for visible photodegradation of acid orange 7. J. Colloid Interface Sci. 2018, 511, 39–47. [Google Scholar] [CrossRef]
  146. Arul, N.S.; Mangalaraj, D.; Chen, P.C.; Ponpandian, N.; Viswanathan, C. Self assembly of Co doped CeO2 microspheres from nanocubes by hydrothermal method and their photodegradation activity on AO7. Mater. Lett. 2011, 65, 3320–3322. [Google Scholar] [CrossRef]
  147. Singh, K.; Kumar, K.; Srivastava, S.; Chowdhury, A. Effect of rare-earth doping in CeO2 matrix: Correlations with structure, catalytic and visible light photocatalytic properties. Ceram. Int. 2017, 43, 17041–17047. [Google Scholar] [CrossRef]
  148. Chen, X.; Mao, S.S. Titanium dioxide materials: Synthesis, properties, modifications, and applications. Chem. Rev. 2007, 107, 2891–2959. [Google Scholar] [CrossRef]
  149. Jorge, A.B.; Fraxedas, J.; Cantarero, A.; Williams, A.J.; Rodgers, J.; Attfield, J.P.; Fuertes, A. Nitrogen doping ceria. Chem. Mater. 2008, 20, 1682–1684. [Google Scholar] [CrossRef]
  150. Yue, L.; Zhang, X.-M. Structural characterization and photocatalytic behaviours of doped CeO2 nanoparticles. J. Alloys Compd. 2009, 475, 702–705. [Google Scholar] [CrossRef]
  151. Vinodkumar, T.; Rao, B.G.; Reddy, B.M. Influence of isovalent and aliovalent dopants on the reactivity of cerium oxide for catalytic applications. Catal. Today 2015, 253, 57–64. [Google Scholar] [CrossRef]
  152. Wang, H.; Zhang, L.; Chen, Z.; Hu, J.; Li, S.; Wang, Z.; Liu, J.; Wang, X. Semiconductor heterojunction photocatalyst: Design, construction, and photocatalytic performances. Chem. Soc. Rev. 2014, 43, 5234–5244. [Google Scholar] [CrossRef]
  153. Peleyeju, M.G.; Arotiba, O.A. Recent trend in visible-light photoelectrocatalytic systems for degradation of organic contaminants in water/wastewater. Environ. Sci. Water Res. Technol. 2018, 4, 1389–1411. [Google Scholar] [CrossRef]
  154. Ge, J.; Zhang, Y.; Heo, Y.-J.; Park, S.-J. Advanced design and synthesis of composite photocatalyst for the remediation of wastewater: A review. Catalysts 2019, 9, 122. [Google Scholar] [CrossRef] [Green Version]
  155. Low, J.; Jaroniec, M.; Wageh, S.; Al-Ghamdi, A.A. Heterojunction photocatalysts. Adv. Mater. 2017, 29, 1601694. [Google Scholar] [CrossRef]
  156. Ameen, S.; Akhtar, M.S.; Seo, H.-K.; Shin, H.-S. Solution processed CeO2/TiO2 nanocomposite as potent visible light photocatalyst for the degradation of bromophenol dye. Chem. Eng. J. 2014, 247, 193–198. [Google Scholar] [CrossRef]
  157. Wang, N.; Pan, Y.; Lu, T.; Li, X.; Wu, S.; Wu, J. A new ribbon-ignition method for fabricating p-CuO/n-CeO2 heterojunction with enhanced photocatalytic activity. Appl. Surf. Sci. 2017, 403, 699–706. [Google Scholar] [CrossRef]
  158. Zhu, L.; Li, H.; Xia, P.; Liu, Z.; Xiong, D. Hierarchical ZnO decorated CeO2 nanoparticles as the direct Z-scheme heterojunctions for enhanced photocatalytic activity. ACS Appl. Mater. Interfaces 2018, 10, 39679–39687. [Google Scholar] [CrossRef] [PubMed]
  159. Qiao, Q.; Yang, K.; Ma, L.-L.; Huang, W.-Q.; Zhou, B.-X.; Pan, A.; Hu, W.; Fan, X.; Huang, G.-F. Facile in situ construction of mediator-free direct Z-scheme g-C3N4/CeO2 heterojunctions with highly efficient photocatalytic activity. J. Phys. D Appl. Phys. 2018, 51, 275302. [Google Scholar] [CrossRef]
  160. Li, H.; Tu, W.; Zhou, Y.; Zou, Z. Z-Scheme photocatalytic systems for promoting photocatalytic performance: Recent progress and future challenges. Adv. Sci. 2016, 3, 1500389. [Google Scholar] [CrossRef] [Green Version]
  161. Hao, C.; Li, J.; Zhang, Z.; Ji, Y.; Zhan, H.; Xiao, F.; Wang, D.; Liu, B.; Su, F. Enhancement of photocatalytic properties of TiO2 nanoparticles doped with CeO2 and supported on SiO2 for phenol degradation. Appl. Surf. Sci. 2015, 331, 17–26. [Google Scholar] [CrossRef]
  162. Li, G.; Zhang, D.; Yu, J.C. Thermally stable mesoporous CeO2/TiO2 visible light photocatalysts. Phys. Chem. Chem. Phys. 2009, 11, 3775–3782. [Google Scholar] [CrossRef]
  163. Liriano-Jorge, C.F.; Sohmen, U.; Özkan, A.; Gulyas, H.; Otterpohl, R. TiO2 photocatalyst nanoparticle separation: Flocculation in different matrices and use of powdered activated carbon as a precoat in low-cost fabric filtration. Adv. Mater. Sci. Eng. 2014, 602495. [Google Scholar]
  164. Szabo, T.; Veres, A.; Cho, E.; Khim, J.; Varga, N.; Dekany, I. Photocatalyst separation from aqueous dispersion using graphene oxide/TiO2 nanocomposites. Colloids Surf. A Physicochem. Eng. Asp. 2013, 433, 230–239. [Google Scholar] [CrossRef]
  165. Moongraksathum, B.; Chen, Y.-W. CeO2-TiO2 mixed oxide thin films with enhanced photocatalytic degradation of organic pollutants. J. Sol-Gel Sci. Technol. 2017, 82, 772–782. [Google Scholar] [CrossRef]
  166. Eskandarloo, H.; Badiei, A.; Behnajady, M.A. TiO2/CeO2 hybrid photocatalyst with enhanced photocatalytic activity: Optimization of synthesis variables. Ind. Eng. Chem. Res. 2014, 53, 7847–7855. [Google Scholar] [CrossRef]
  167. Tuyen, L.T.T.; Quang, D.A.; Toan, T.T.T.; Tung, T.Q.; Hoa, T.T.; Mau, T.X.; Khieu, D.Q. Synthesis of CeO2/TiO2 nanotubes and heterogeneous photocatalytic degradation of methylene blue. J. Environ. Chem. Eng. 2018, 6, 5999–6011. [Google Scholar] [CrossRef]
  168. Liu, H.; Wang, M.; Wang, Y.; Liang, Y.; Cao, W.; Su, Y. Ionic liquid-templated synthesis of mesoporous CeO2-TiO2 nanoparticles and their enhanced photocatalytic activities under UV or visible light. J. Photochem. Photobiol. A 2011, 223, 157–164. [Google Scholar] [CrossRef]
  169. Lu, X.; Li, X.; Qian, J.; Miao, N.; Yao, C.; Chen, Z. Synthesis and characterization of CeO2/TiO2 nanotube arrays and enhanced photocatalytic oxidative desulfurization performance. J. Alloys Compd. 2016, 661, 363–371. [Google Scholar] [CrossRef]
  170. Chen, F.; Ho, P.; 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]
  171. Yuan, B.; Long, Y.; Wu, L.; Liang, K.; Wen, H.; Luo, S.; Huo, H.; Yang, H.; Ma, J. TiO2@h-CeO2: A composite yolk-shell microsphere with enhanced photodegradation activity. Catal. Sci. Technol. 2016, 6, 6396–6405. [Google Scholar] [CrossRef]
  172. Zhang, L.; Zhang, J.; Jiu, H.; Zhang, X.; Xu, M. Preparation of hollow core/shell CeO2@TiO2 with enhanced photocatalytic performance. J. Mater. Sci. 2015, 50, 5228–5237. [Google Scholar] [CrossRef]
  173. Liu, Y.; Li, T.; Chen, W.; Guo, Y.; Liu, L.; Guo, H. Hierarchical hollow TiO2@CeO2 nanocube heterostructures for photocatalytic detoxification of cyanide. RSC Adv. 2015, 5, 11733–11737. [Google Scholar] [CrossRef]
  174. Sivalingam, J.R.; Kait, C.F.; Wilfred, C.D. CeO2-TiO2 photocatalyst: Ionic liquid-mediated synthesis, characterization, and performance for diisopropanolamine visible light degradation. Bull. Chem. React. Eng. Catal. 2018, 13, 170–178. [Google Scholar] [CrossRef] [Green Version]
  175. Vieira, G.B.; Jose, H.J.; Peterson, M.; Baldissarelli, V.Z.; Alvarez, P.; Moreira, R.F.P.M. CeO2/TiO2 nanostructures enhanced adsorption and photocatalytic degradation of organic compounds in aqueous suspension. J. Photochem. Photobiol. A 2018, 353, 325–336. [Google Scholar] [CrossRef]
  176. Velasco, L.F.; Parra, J.B.; Ania, C.O. Role of activated carbon features on the photocatalytic degradation of phenol. Appl. Surf. Sci. 2010, 256, 5254–5258. [Google Scholar] [CrossRef] [Green Version]
  177. Wang, M.Y.; Zhu, W.; Zhang, D.E.; Li, S.A.; Ma, W.X.; Tong, Z.W.; Chen, J. CeO2 hollow nanospheres decorated reduced graphene oxide composite for efficient photocatalytic dye-degradation. Mater. Lett. 2014, 137, 229–232. [Google Scholar]
  178. Lam, S.-M.; Sin, J.-C.; Mohamed, A.R. A review on photocatalytic application of g-C3N4/semiconductor (CNS) nanocomposites towards the erasure of dyeing wastewater. Mater. Sci. Semicond. Process. 2016, 47, 62–84. [Google Scholar] [CrossRef]
  179. Xiang, Q.; Yu, J.; Jaroniec, M. Graphene-based semiconductor photocatalysts. Chem. Soc. Rev. 2011. [Google Scholar] [CrossRef] [PubMed]
  180. Gusain, R.; Kumar, P.; Sharma, O.P.; Jain, S.L.; Khatri, O.P. Reduced graphene oxide-CuO nanocomposites for photocatalytic conversion of CO2 into methanol under visible light irradiation. Appl. Catal. B Environ. 2016, 181, 352–362. [Google Scholar] [CrossRef]
  181. Chang, H.; Wu, H. Graphene-based nanocomposites: Preparation, functionalization, and energy and environmental applications. Energy Environ. Sci. 2013, 6, 3483–3507. [Google Scholar] [CrossRef]
  182. Albero, J.; Mateo, D.; Garcia, H. Graphene-based materials as efficient photocatalysts for water splitting. Molecules 2019, 24, 906. [Google Scholar] [CrossRef] [Green Version]
  183. Dreyer, D.R.; Ruoff, R.S.; Bielawski, C.W. From conception to realization: An historial account of graphene and some perspectives for its future. Angew. Chem. Int. Ed. 2010, 49, 9336–9344. [Google Scholar] [CrossRef]
  184. Inagaki, M.; Tsumura, T.; Kinumoto, T.; Toyoda, M. Graphitic carbon nitrides (g-C3N4) with comparative discussion to carbon materials. Carbon 2019, 141, 580–607. [Google Scholar] [CrossRef]
  185. Darkwah, W.K.; Oswald, K.A. Photocatalytic applications of heterostructure graphitic carbon nitride: Pollutant degradation, hydrogen gas production (water splitting), and CO2 reduction. Nanoscale Res. Lett. 2019, 14, 234. [Google Scholar] [CrossRef]
  186. Samsudin, M.F.R.; Bacho, N.; Sufian, S. Recent development of graphitic carbon nitride-based photocatalyst for environmental pollution remediation. Chapter 3. In Nanocatalysts; Sinha, I., Shukla, M., Eds.; IntechOpen: London, UK, 2018; pp. 1–15. [Google Scholar]
  187. Ji, Z.; Shen, X.; Li, M.; Zhou, H.; Zhu, G.; Chen, K. Synthesis of reduced graphene oxide/CeO2 nanocomposites and their photocatalytic properties. Nanotechnology 2013, 24, 115603. [Google Scholar] [CrossRef]
  188. Murali, A.; Sarswat, P.K.; Free, M.L. Minimizing electron-hole pair recombination through band-gap engineering in novel ZnO-CeO2-rGO ternary nanocomposite for photoelectrochemical and photocatalytic applications. Environ. Sci. Pollut. Res. 2010, 27, 25042–25056. [Google Scholar] [CrossRef] [PubMed]
  189. Huang, T.; Wu, J.; Zhao, Z.; Zeng, T.; Zhang, J.; Xu, A.; Zhou, X.; Qi, Y.; Ren, J.; Zhou, R.; et al. Synthesis and photocatalytic performance of CuO-CeO2/graphene oxide. Mater. Lett. 2016, 185, 503–506. [Google Scholar] [CrossRef]
  190. Huang, L.; Li, Y.; Xu, H.; Xu, Y.; Xia, J.; Wang, K.; Li, H.; Cheng, X. Synthesis and characterization of CeO2/g-C3N4 composites with enhanced visible-light photocatalytic activity. RSC Adv. 2013, 3, 22269–22279. [Google Scholar] [CrossRef]
  191. Tian, N.; Huang, H.; Liu, C.; Dong, F.; Zhang, T.; Du, X.; Yu, S.; Zhang, Y. In situ co-pyrolysis fabrication of CeO2/g-C3N4 n-n type heterojunction for synchronously promoting photo-induced oxidation and reduction properties. J. Mater. Chem. A 2015, 3, 17120–17129. [Google Scholar] [CrossRef]
  192. Liu, W.; Zhou, J.; Hu, Z. Nano-sized g-C3N4 thin layer @ CeO2 sphere core-shell photocatalyst combined with H2O2 to degrade doxycycline in water under visible light irradiation. Sep. Purif. Technol. 2019, 227, 115665. [Google Scholar] [CrossRef]
  193. Jourshabani, M.; Shariatinia, Z.; Badiei, A. Facile one-pot synthesis of cerium oxide/sulfur-doped graphitic carbon nitride (g-C3N4) as efficient nanophotocatalysts under visible light irradiation. J. Colloid Interface Sci. 2017, 507, 59–73. [Google Scholar] [CrossRef]
  194. Li, X.; Zhu, W.; Lu, X.; Zuo, S.; Yao, C.; Ni, C. Integrated nanostructures of CeO2/attapulgite/g-C3N4 as efficient catalyst for photocatalytic desulfurization: Mechanism, kinetics and influencing factors. Chem. Eng. J. 2017, 326, 87–98. [Google Scholar] [CrossRef]
  195. Vignesh, S.; Suganthi, S.; Sundar, J.K.; Raj, V. Construction of α-Fe2O3/CeO2 decorated g-C3N4 nanosheets for magnetically separable efficient photocatalytic performance under visible light exposure and bacterial disinfection. Appl. Surf. Sci. 2019, 488, 763–777. [Google Scholar] [CrossRef]
  196. Yuan, Y.; Huang, G.-F.; Hu, W.-Y.; Xiong, D.-N.; Zhou, B.-X.; Chang, S.; Huang, W.-Q. Construction of g-C3N4/CeO2/ZnO ternary photocatalysts with enhanced photocatalytic performance. J. Phys. Chem. Solids 2017, 106, 1–9. [Google Scholar] [CrossRef]
  197. Stelo, F.; Kublik, N.; Ullah, S.; Wender, H. Recent advances in Bi2MoO6 based Z-scheme heterojunctions for photocatalytic degradation of pollutants. J. Alloys Compd. 2020, 829, 154591. [Google Scholar] [CrossRef]
  198. You, L.; Gao, M.; Li, T.; Guo, L.; Chen, P.; Liu, M. Investigation of the kinetics and mechanism of Z-scheme Ag3PO4/WO3 p-n junction photocatalysts with enhanced removal efficiency for RhB. New. J. Chem. 2019, 43, 17104. [Google Scholar] [CrossRef]
  199. Issarapanacheewin, S.; Wetchakun, K.; Phanichphant, S.; Kangwansupamonkon, W.; Wetchakun, N. Efficient photocatalytic degradation of Rhodamine B by a novel CeO2/Bi2WO6 composite film. Catal. Today 2016, 278, 280–290. [Google Scholar] [CrossRef]
  200. Lv, Z.; Zhou, H.; Liu, H.; Liu, B.; Liang, M.; Guo, H. Controlled assemble of oxygen vacant CeO2@Bi2WO6 hollow magnetic microcapsule heterostructures for visible-light photocatalytic activity. Chem. Eng. J. 2017, 330, 1297–1305. [Google Scholar] [CrossRef]
  201. Liu, Y.; Zhu, G.; Gao, J.; Hojamberdiev, M.; Lu, H.; Zhu, R.; Wei, X.; Liu, P. A novel CeO2/Bi4Ti3O12 composite heterojunction structure with an enhanced photocatalytic activity for bisphenol A. J. Alloys Compd. 2016, 688, 487–496. [Google Scholar] [CrossRef]
  202. Dai, W.; Hu, X.; Wang, T.; Xiong, W.; Luo, X.; Zou, J. Hierarchical CeO2/Bi2MoO6 heterostructured nanocomposites for photoreduction of CO2 into hydrocarbons under visible light irradiation. Appl. Surf. Sci. 2018, 434, 481–491. [Google Scholar] [CrossRef]
  203. Hsieh, S.-H.; Manivel, A.; Lee, G.-J.; Wu, J.J. Synthesis of mesoporous Bi2O3/CeO2 microsphere for photocatalytic degradation of Orange II dye. Mater. Res. Bull. 2013, 48, 4174–4180. [Google Scholar] [CrossRef]
  204. Wen, X.-J.; Niu, C.-G.; Zhang, L.; Zeng, G.-M. Novel p-n heterojunction BiOI/CeO2 photocatalyst for wider spectrum visible-light photocatalytic degradation of refractory pollutants. Dalton Trans. 2017, 46, 4982–4993. [Google Scholar] [CrossRef]
  205. Gu, S.; Li, W.; Wang, F.; Wang, S.; Zhou, H.; Li, H. Synthesis of buckhorn-like BiVO4 with a shell of CeOx nanodots: Effect of heterojunction structure on the enhancement of photocatalytic activity. Appl. Catal. B Environ. 2015, 170–171, 186–194. [Google Scholar] [CrossRef]
  206. Wetchakun, N.; Chaiwichain, S.; Inceesungvorn, B.; Pingmuang, K.; Phanichphant, S.; Minett, A.I.; Chen, J. BiVO4/CeO2 nanocomposites with high visible-light-induced photocatalytic activity. ACS Appl. Mater. Interfaces 2012, 4, 3718–3723. [Google Scholar] [CrossRef]
  207. Yang, Z.-M.; Huang, G.-F.; Huang, W.-Q.; Wei, J.-M.; Yan, X.-G.; Liu, Y.-Y.; Jiao, C.; Wan, Z.; Pan, A. Novel Ag3PO4/CeO2 composite with high efficiency and stability for photocatalytic applications. J. Mater. Chem. A 2014, 2, 1750–1756. [Google Scholar] [CrossRef]
  208. Lin, W.; Zhang, S.; Wang, D.; Zhang, C.; Sun, D. Ultrasound-assisted synthesis of high-efficiency Ag3PO4/CeO2 heterojunction photocatalyst. Ceram. Int. 2015, 41, 8956–8963. [Google Scholar] [CrossRef]
  209. Xiong, W.; Dai, W.; Hu, X.; Yang, L.; Wang, T.; Qin, Y.; Luo, X.; Zou, J. Enhanced photocatalytic reduction of CO2 into alcohols on Z-scheme Ag/Ag3PO4/CeO2 driven by visible light. Mater. Lett. 2018, 232, 36–39. [Google Scholar] [CrossRef]
  210. Wen, X.-J.; Niu, C.-G.; Ruan, M.; Zhang, L.; Zeng, G.-M. AgI nanoparticles-decorated CeO2 microsheets photocatalyst for the degradation of organic dye and tetracycline under visible-light irradiation. J. Colloid Interface Sci. 2017, 497, 368–377. [Google Scholar] [CrossRef] [PubMed]
  211. Wen, X.-J.; Niu, C.-G.; Zhang, L.; Liang, C.; Zeng, G.-M. A novel Ag2O/CeO2 heterojunction photocatalysts for photocatalytic degradation of enrofloxacin: Possible degradation pathways, mineralization activity and an in depth mechanism insight. Appl. Catal. B Environ. 2018, 221, 701–704. [Google Scholar] [CrossRef]
  212. Wen, X.-J.; Niu, C.-G.; Guo, H.; Liang, C.; Zeng, G.-M. Photocatalytic degradation of levofloxacin by ternary Ag2CO3/CeO2/AgBr photocatalyst under visible-light irradiation: Degradation pathways, mineralization ability, and an accelerated interfacial charge transfer process study. J. Catal. 2018, 358, 211–223. [Google Scholar] [CrossRef]
  213. Madkour, M.; Allam, O.G.; Nazeer, A.A.; Amin, M.O.; Al-Hetlani, E. CeO2-based nanoheterostructures with p-n and n-n heterojunction arrangements for enhancing the solar-driven photodegradation of rhodamine 6G dye. J. Mater. Sci. Mater. Electron. 2019, 30, 10857–10866. [Google Scholar] [CrossRef]
  214. Saravanan, R.; Joicy, S.; Gupta, V.K.; Narayanan, V.; Stephen, A. Visible light induced degradation of methylene blue using CeO2/V2O5 and CeO2/CuO catalysts. Mater. Sci. Eng. C 2013, 33, 4725–4731. [Google Scholar] [CrossRef]
  215. Elaziouti, A.; Laouedj, N.; Bekka, A.; Vannier, R.-N. Preparation and characterization of p-n heterojunction CuBi2O4/CeO2 and its photocatalytic activities under UVA light irradiation. J. King Saud Univ. Sci. 2015, 27, 120–135. [Google Scholar] [CrossRef] [Green Version]
  216. Valente, J.S.; Tzompantzi, F.; Prince, J. Highly efficient photocatalytic elimination of phenol and chlorinated phenols by CeO2/MgAl layered double hydroxides. Appl. Catal. B Environ. 2011, 102, 276–285. [Google Scholar] [CrossRef]
  217. Latha, P.; Prakash, K.; Karuthapandian, S. Enhanced visible light photocatalytic activity of CeO2/alumina nanocomposite; Synthesized via facile mixing-calcination method for dye degradation. Adv. Powder Technol. 2017, 28, 2903–2913. [Google Scholar] [CrossRef]
  218. Yang, Z.-M.; Hou, S.-C.; Huang, G.-F.; Duan, H.-G.; Huang, W.-Q. Electrospinning preparation of p-type NiO/n-type CeO2 heterojunctions with enhanced photocatalytic activity. Mater. Lett. 2014, 133, 109–112. [Google Scholar] [CrossRef]
  219. Usharani, S.; Rajendran, V. Optical, magnetic and visible light photocatalytic activity of CeO2/SnO2 nanocomposites. Eng. Sci. Technol. 2016, 19, 2088–2093. [Google Scholar] [CrossRef] [Green Version]
  220. Rajendra, S.; Khan, M.M.; Gracia, F.; Qin, J.; Gupta, V.K.; Arumainathan, S. Ce3+-ion-induced visible-light photocatalytic degradation and electrochemical activity of ZnO/CeO2 nanocomposite. Sci. Rep. 2016, 6, 31641. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  221. Prabhu, S.; Viswanathan, T.; Jothivenkatachalam, K.; Jeganathan, K. Visible light photocatalytic activity of CeO2-ZnO-TiO2 composites for the degradation of Rhodamine B. Indian J. Mater. Sci. 2014, 536123. [Google Scholar] [CrossRef] [Green Version]
  222. Ma, Y.; Jiang, J.; Zhu, A.; Tan, P.; Bian, Y.; Zeng, W.; Cui, H.; Pan, J. Enhanced visible-light photocatalytic degradation by Mn3O4/CeO2 heterojunction: A Z-scheme system photocatalyst. Inorg. Chem. Front. 2018, 5, 2579–2586. [Google Scholar] [CrossRef]
  223. Wang, Z.; Li, X.; Qian, H.; Zuo, S.; Yan, X.; Chen, Q.; Yao, C. Upconversion Tm3+: CeO2/palygorskite as direct Z-scheme heterostructure for photocatalytic degradation of bisphenol A. J. Photochem. Photobiol. A 2019, 372, 42–48. [Google Scholar] [CrossRef]
  224. Lu, X.-H.; Xie, S.-L.; Zhai, T.; Zhao, Y.-F.; Zhang, P.; Zhang, Y.-L.; Tong, Y.-X. Monodisperse CeO2/CdS heterostructured spheres: One-pot synthesis and enhanced photocatalytic hydrogen activity. RSC Adv. 2011, 1, 1207–1210. [Google Scholar] [CrossRef]
  225. Martinez-Huitle, C.A.; Ferro, S. Electrochemical oxidation of organic pollutants for the wastewater treatment: Direct and indirect processes. Chem. Soc. Rev. 2006, 35, 1324–1340. [Google Scholar] [CrossRef]
  226. Brillas, E.; Martinez-Huitle, C.A. Decontamination of wastewaters containing synthetic organic dyes by electrochemical methods. An updated review. Appl. Catal. B Environ. 2015, 166–167, 603–643. [Google Scholar] [CrossRef]
  227. Malpass, G.R.P.; Miwa, D.G.; Machado, S.A.S.; Motheo, A.J. SnO2-based materials for pesticide degradation. J. Hazard. Mater. 2010, 180, 145–151. [Google Scholar] [CrossRef]
  228. Peleyelu, M.G.; Umukoro, E.H.; Bablola, J.O.; Arotiba, O.A. Electrochemical degradation of an anthraquinone dye on an expanded graphite-diamond composite electrode. Electrocatalysis 2016, 7, 132–139. [Google Scholar] [CrossRef]
  229. Jüttner, K.; Galla, U.; Schmieder, H. Electrochemical approaches to environmental problems in the process industry. Electrochim. Acta 2000, 45, 2575–2594. [Google Scholar] [CrossRef]
  230. Umukoro, E.H.; Kumar, N.; Ngila, J.C.; Arotiba, O.A. Expanded graphite supported p-n MoS2-SnO2 heterojunction nanocomposite electrode for enhanced photo-electrocatalytic degradation of a pharmaceutical pollutant. J. Electroanal. Chem. 2018, 827, 193–203. [Google Scholar] [CrossRef]
  231. Kusmierek, E.; Chrzescijanska, E. Application of TiO2-RuO2/Ti electrodes modified with WO3 in electro- and photoelectrochemical oxidation of Acid Orange 7 dye. J. Photochem. Photobiol. A 2015, 302, 59–68. [Google Scholar] [CrossRef]
  232. Tantis, I.; Bousiakou, L.; Karikas, G.-A.; Lianos, P. Photocatalytic and photoelectrocatalytic degradation of the antibacterial agent ciprofloxacin. Photochem. Photobiol. Sci. 2015, 14, 603–607. [Google Scholar] [CrossRef]
  233. Xiao, S.; Song, Y.; Tian, Z.; Tu, X.; Hu, X.; Ruixia, L. Enhanced mineralization of antibiotic berberine by the photoelectrochemical process in presence of chlorides and its optimization by response surface methodology. Environ. Earth Sci. 2015, 73, 4947–4955. [Google Scholar] [CrossRef]
  234. Li, J.; Lv, S.; Liu, Y.; Bai, J.; Zhou, B.; Hu, X. Photoelectrocatalytic activity of an n-ZnO/p-Cu2O/n-TNA ternary heterojunction electrode for tetracycline degradation. J. Hazard. Mater. 2013, 262, 482–488. [Google Scholar] [CrossRef]
  235. Garcia-Segura, S.; Brillas, E. Applied photoelectrocatalysis on the degradation of organic pollutants in wastewater. J. Photochem. Photobiol. C 2017, 31, 1–35. [Google Scholar] [CrossRef]
  236. Zarei, E.; Ojani, R. Fundamentals and some applications of photoelectrocatalysis and effective factors on its efficiency: A review. J. Solid State Electrochem. 2017, 21, 305–336. [Google Scholar] [CrossRef]
  237. Yu, Y.; Yu, X.; Yang, S. Preparation and characterization of CeO2 decorated TiO2 nanotube arrays photoelectrode and its enhanced photoelectrocatalytic efficiency for degradation of methyl orange. J. Mater. Sci. Mater. Electron. 2015, 26, 5715–5723. [Google Scholar] [CrossRef]
  238. Mafa, P.J.; Mamba, B.B.; Kuvarega, A.T. Photoelectrocatalytic evaluation of EG-CeO2 photoanode on degradation of 2,4-dichlorophenol. Sol. Energy Mater. Sol. Cells 2020, 208, 110416. [Google Scholar] [CrossRef]
  239. Zhou, Q.; Xing, A.; Li, J.; Zhao, D.; Zhao, K.; Lei, M. Synergistic enhancement in photoelectrocatalytic degradation of bisphenol A by CeO2 and reduced graphene oxide co-modified TiO2 nanotube arrays in combination with Fenton oxidation. Electrochim. Acta 2016, 209, 379–388. [Google Scholar] [CrossRef]
  240. Zhou, Q.; Xing, A.; Zhao, D.; Zhao, K. Tetrabromobisphenol A photoelectrocatalytic degradation using reduced graphene oxide and cerium dioxide comodified TiO2 nanotube arrays as electrode under visible light. Chemosphere 2016, 165, 268–276. [Google Scholar] [CrossRef] [PubMed]
  241. Chen, C.; Zhao, D.; Zhou, Q.; Wu, Y.; Zhou, X.; Wang, H. Facile preparation and characterization of polyaniline and CeO2 co-decorated TiO2 nanotube array and its highly efficient photoelectrocatalytic activity. Nanoscale Res. Lett. 2019, 14, 60. [Google Scholar] [CrossRef] [Green Version]
  242. Li, L.; Feng, H.; Wei, X.; Jiang, K.; Xue, S.; Chu, P.K. Ag as cocatalyst and electron-hole medium in CeO2 QDs/Ag/Ag2Se Z-Scheme heterojunction enhanced the photo-electrocatalytic properties of photoelectrode. Nanomaterials 2020, 10, 253. [Google Scholar] [CrossRef] [Green Version]
  243. He, S.; Yan, C.; Chen, X.-Z.; Wang, Z.; Ouyang, T.; Guo, M.-L. Construction of core-shell heterojunction regulating α-Fe2O3 layer on CeO2 nanotube arrays enables highly efficient Z-scheme photoelectrocatalysis. Appl. Catal. B Environ. 2020, 276, 119138. [Google Scholar] [CrossRef]
  244. Wang, Y.; Hu, B.; Hu, C.; Zhou, X. Fabrication of a novel Ti/SnO2-Sb-CeO2@TiO2-SnO2 electrode and photoelectrocatalytic application in wastewater treatment. Mater. Sci. Semicond. Process. 2015, 40, 744–751. [Google Scholar] [CrossRef]
  245. Chawla, P.; Tripathi, M. Surface modification of semiconductor photoanode for photoelectrochemical water splitting. Int. J. Hydrog. Energy 2016, 41, 7987–7992. [Google Scholar] [CrossRef]
  246. Pan, Z.; Han, E.; Zheng, J.; Lu, J.; Wang, X.; Yin, Y.; Waterhouse, G.I.N.; Wang, X.; Li, P. Highly efficient photoelectrocatalytic reduction of CO2 to methanol by a p-n heterojunction CeO2/CuO/Cu catalyst. Nano-Micro Lett. 2020, 12, 18. [Google Scholar] [CrossRef] [Green Version]
Figure 1. Band position (valence band top and conduction band bottom) of cerium dioxide (CeO2) compared with several selected redox potentials of processes occurring at the semiconductor surface.
Figure 1. Band position (valence band top and conduction band bottom) of cerium dioxide (CeO2) compared with several selected redox potentials of processes occurring at the semiconductor surface.
Catalysts 10 01435 g001
Figure 2. Energy of a CeO2 semiconductor (n-type) and band positions in the semiconductor under equilibrium conditions after semiconductor–electrolyte contact and after irradiation.
Figure 2. Energy of a CeO2 semiconductor (n-type) and band positions in the semiconductor under equilibrium conditions after semiconductor–electrolyte contact and after irradiation.
Catalysts 10 01435 g002
Figure 3. A schematic diagram of the effects of CeO2 doping with metals (Me) on enhanced photogenerated hole (h+)/photoexcited electron (e) pair separation.
Figure 3. A schematic diagram of the effects of CeO2 doping with metals (Me) on enhanced photogenerated hole (h+)/photoexcited electron (e) pair separation.
Catalysts 10 01435 g003
Figure 4. A schematic diagram of CeO2/titanium dioxide (TiO2) composite photoexcitation under UV and VIS light irradiation.
Figure 4. A schematic diagram of CeO2/titanium dioxide (TiO2) composite photoexcitation under UV and VIS light irradiation.
Catalysts 10 01435 g004
Figure 5. A schematic diagram of h+/e pair separation in CeO2 composites with graphitic carbon nitride (g-C3N4) in a type II heterojunction.
Figure 5. A schematic diagram of h+/e pair separation in CeO2 composites with graphitic carbon nitride (g-C3N4) in a type II heterojunction.
Catalysts 10 01435 g005
Figure 6. A schematic illustration of different types of Z-scheme heterojunctions: I—Z-scheme with a redox pair mediator; II—Z-scheme with a solid mediator; and III—direct Z-scheme. SC—semiconductor; D—donor; A—acceptor; and SM—solid mediator.
Figure 6. A schematic illustration of different types of Z-scheme heterojunctions: I—Z-scheme with a redox pair mediator; II—Z-scheme with a solid mediator; and III—direct Z-scheme. SC—semiconductor; D—donor; A—acceptor; and SM—solid mediator.
Catalysts 10 01435 g006
Table 1. Photocatalytic degradation of various pollutants with the application of CeO2 as a semiconductor photocatalytic material.
Table 1. Photocatalytic degradation of various pollutants with the application of CeO2 as a semiconductor photocatalytic material.
CeO2 FormPreparation MethodBand GapCatalyst DosePollutant, ConcentrationLight SourceDegradation Efficiency (Process Time)Ref.
NanowiresMicrowave-assisted hydrothermal method2.65 eV0.2 g/LMO,
0.01 mM
250 W Hg lampNanowires—8%,
Nanorods—78%
(100 min)
[126]
Crystal phasePrecipitation method 1.0 g/LAO7,
70 mg/L
1000 W halogen lamp
(>420 nm)
AO7—97% (11 h),
TOC—60% (12 h)
[127]
Crystal phasePrecipitation followed by calcination 1.0 g/LAO7, 70 mg/L,
pH = 6.8
1000 W halogen lamp (>420 nm)AO7—92%,
TOC—40%
(10 h)
[128]
NanostructureSol-gel method2.52 eV1.0 g/LXylene Milling Yellow 6G, 10 ppm125 Hg lamp (365 nm)100% (30 min)[129]
NPs—nanoparticles,
HNRs—hierarchical nanorods,
HNWs—hierarchical nanowires
Electrodeposition method3.2 eV
(HNRs and HNWs)
0.2 g/LMO, 20 mg/L,
pH = 3
500 W Xe lamp (200–800 nm)NPs—68.8%,
HNRs—98.2%,
HNWs—99.3%
(180 min)
[25]
NPsDissolution and hydrolysis method2.94 eV1.3 g/LPhenol (Ph),
2-chlorophenol
(2-CP),
2-bromophenol
(2-BP),
2-nitrophenol
(2-NP),
50 mg/L
Sunlight,
900 × 102 lx
Ph—<35%,
2-CP—~98%,
2-BP—~96%,
2-NP—99%
(180 min),
TOC
Ph—~20%,
2-CP—~94%,
2-BP—91%,
2-NP—96%
(180 min)
[130]
NanostructureElectron beam method3.36 eV—NPs,
3.15 eV—30 kGy,
3.12 eV—90 kGy
0.1 g/Lp-nitrophenol
(4-NP), 5 mg/L,
MB, 10 mg/L
400 W lamp
(>420 nm)
4-NP
NPs—45%,
30 kGy—60%,
90 kGy—66%
(6 h)
MB
NPs—36%,
30 kGy—65%,
90 kGy—75%
(6 h)
[131]
CeO2
X-architecture particles
Hydrothermal route assisted by polyvinyl pyrrolidone 1 g/LMB, 40 mg/L500 W Hg lamp98%
(120 h)
[132]
CeO2
NPs
Microwave-assisted hydrothermal method2.54 eV—rods,
2.75 eV—beans,
2.76 eV—rods/cubes,
2.95 eV—hexagons
0.3 g/LRhB, 0.01 mM,
Ciprofloxacin (CIP)—0.01 mM
15 W Philips lamps,
9.55 mW/cm2
RhB
Hexagons—20.6%,
Rods/cubes (in the presence of ammonia oxalate)—86%,
CIP
Rods/cubes—53%
(60 min)
[133]
CeO2/SCB
Sugarcane bagasse
Heterogeneous precipitation2.80 eV—CeO2,
2.20 eV—CeO2/SCB
10.0 g/LMB, 20 ppm36 W lamp
(365 nm),
145 μW/cm2
CeO2—30%,
CeO2/SCB—90%
(120 min)
[134]
Abbreviations: MO—methyl orange; MB—methylene blue; AO7—acid orange 7; RhB—rhodamine B; TOC—total organic carbon; and NPs—nanoparticles.
Table 2. Photocatalytic degradation of various pollutants with the application of CeO2 doped with metals and non-metals as a photocatalytic material.
Table 2. Photocatalytic degradation of various pollutants with the application of CeO2 doped with metals and non-metals as a photocatalytic material.
PhotocatalystPreparation MethodBand Gap, eVCatalyst DosePollutant, ConcentrationLight SourceDegradation Efficiency (Process Time)Ref.
F-doped CeO2Low temp. solution combustionCeO2—3.16 eV,
F-CeO2—2.88 eV
0.375 g/LMB, 10 mg/L300 W UV lampCeO2—46%
(15 min),
F-CeO2—92.1%
(6 min)
[136]
N-CeO2
Microspheres
Solvothermal synthesisCeO2—2.98 eV,
N-CeO2—2.83 eV
0.1 g/LRhodamine 6G, 0.5 g/L1000 W Xe lamp
(>420 nm)
CeO2—8.7%,
N-CeO2—68.8%
(160 min)
[137]
N-CeO2Wet-chemical route MB150 W Xe lamp (>500 nm)CeO2—~9%,
N-CeO2—~18%
(320 min)
[138]
C-N-CeO2Solvothermal method with hexamethylenetetramine AO7, 0.2 mM, pH = 3100 W Hg lamp (HML),
10 W fluorescent lamp (CFL)
HML
TiO2—68.4%,
CeO2-TiO2—84%,
C-N-CeO2—94.4%,
CFL
TiO2—43%,
CeO2-TiO2—93.3%
C-N-CeO2—98.8%
(1 h)
[28]
Y-CeO2
Nanorods
Hydrothermal synthesisCeO2—2.68 eV,
Y-CeO2—2.62 eV
0.2 g/LCarmine indigo (CI)—15.55 mg/L,
RhB, 5 mg/L
450 W Hg lampCI
CeO2—39%,
Y-CeO2—58%
RhB
CeO2—15%,
Y-CeO2—35%
(100 min)
[82]
Y-CeO2Nitrate-fuel self-sustaining combustionCeO2—3.31 eV,
Y-CeO2—2.96 eV
1 g/LRhB, 1 g/L250 W halogen lamp (532 nm)99%
(3 h)
[139]
Y-CeO2
Hedgehog-like
hierarchical structure
Hydrothermal processCeO2—3.20 eV,
Y-CeO2—3.35 eV
20 mg spread on glass dishAcetylaldehyde, 500 ppmLED diode
(365 nm)
CO2 evolution
CeO2—~55 ppm,
Y-CeO2—~350 ppm
(24 h)
[140]
CeO2 (CP)
Mg-CeO2 (CM)
Ca-CeO2 (CC)
Sr-CeO2 (CS)
Ba-CeO2 (CB)
Hydrothermal method 0.05 g/LMB, 10 mg/L12 W UV lampCP—33%,
CM—60%,
CC—68%,
CS—75%,
CM—84%
(120 min)
[141]
CeO2
Ag-CeO2
Bi-CeO2
Cd-CeO2
Pb-CeO2
Facile one-step precipitation methodCeO2—2.84 eV,
Ag-CeO2—1.86 eV,
Bi-CeO2—2.60 eV,
Cd-CeO2—2.70 eV,
Pb-CeO2—2.25 eV
0.05 g/LMB, 30 mg/LDirect sunlight irradiationCeO2—10%,
Ag-CeO2—99.6%,
Bi-CeO2—~90%,
Cd-CeO2—~90%,
Pb-CeO2—~80%
(90 min)
[142]
Fe-CeO2Flame spray pyrolysisCeO2—3.18 eV,
Fe(5%)-CeO2—2.90 eV
50 mL, suspension in ethanolHCOOH,
500 μg C,
(COOH)2, 500 μg C
18 W fluorescent lamp
(>400 nm)
HCOOH–~100%,
(COOH)2–~100%
[143]
Fe-CeO2Solvothermal method followed by thermal annealingCeO2—2.92 eV,
Fe-CeO2—2.63 eV
1 cm × 1 cm thin filmMB, 0.5 mMUV light source (365 nm), 2 mW/cm2CeO2—57%,
Fe-CeO2—80%
(180 min)
[144]
Fe-CeO2
Hollow microspheres (HMs)
Co-precipitation routeCeO2(HMs)—3.21 eV,
Fe-CeO2(NPs)—3.37 eV,
Fe-CeO2(HMs)—3.10 eV
0.2 g/L,
with H2O2 (1 mM)
AO7, 20 mg/LXe lamp
(320–1100 nm)
CeO2(HMs)—~81%,
Fe-CeO2(NPs)—~63%,
Fe-CeO2(HMs)—~93%
(80 min)
[145]
Fe-CeO2Homogeneous precipitation with homogeneous/impregnation methodCeO2—~2.98 eV,
Fe(1.5%)-CeO2—~2.64 eV
thin filmMO, 0.025 mM50 W halogen lamp,
185 mW/cm2
CeO2—~55%,
Fe-CeO2—~14%
(120 min)
[26]
Co-CeO2Co-precipitation methodCeO2—3.13 eV,
Co(6%)-CeO2—2.84 eV
1 g/LMB, 15 ppm30 W lamp
(365 nm), sunlight
(~900 lm/cm2)
UV
CeO2—34.5%,
Co(6%)-CeO2—98.7%,
Sunlight
CeO2—34.5%
Co(6%)-CeO2—88.9%
(7 h)
[27]
Co-CeO2Hydrothermal technique 3.3 g/LAO7, 0.02 MUV light
(2 mW) source (365 nm)
96%
(8 h)
[146]
La-CeO2Co-precipitation route 0.25 g/L,
In the presence of NaBH4
MB, 0.25 mM500 W halogen lamp
(400–600 nm)
La(10%)-CeO2—99.99%
(180 min)
[147]
In-CeO2 NanocrystalsFacile, green hydrothermal methodCeO2—2.9 eV,
In(10%)-CeO2—2.7 eV
0.15 g/LMO, 15.5 mg/L,
MB, 15.5 mg/L
110 W UV light sourceMO
CeO2—~53%,
In-CeO2—~92%,
MB
CeO2—~38%,
In-CeO2—~65%
(100 min)
[96]
Ag-CeO2Co-impregnation and co-precipitation method 0.5 g/LTributyltin, 12 mg/L10 W LED lamp (390 nm)CeO2—~20%,
Ag-CeO2—~60%
(30 h)
[97]
Ag-CeO2
Au-CeO3
Co-precipitation routeCeO2—3.12 eV,
Ag(2%)-CeO2—2.90 eV,
Au(0.4%)-CeO2—3.04 eV
0.25 g/LRhB, 1 mg/LUv-Vis lightCeO2—80%,
Ag-CeO2—97.7%
(5 h)
CeO2—36%,
Au-CeO2—94.1%
(2.5 h)
[29]
Abbreviations: RhB—rhodamine B; MO—methyl orange; MB—methylene blue; AO7—acid orange 7; and NPs—nanoparticles.
Table 3. Photocatalytic degradation of various pollutants with the application of CeO2 composites with TiO2 as photocatalytic material.
Table 3. Photocatalytic degradation of various pollutants with the application of CeO2 composites with TiO2 as photocatalytic material.
CompositePreparation MethodBand Gap, eVCatalyst Dose Pollutant, ConcentrationLight SourceDegradation Efficiency (Process Time)Ref.
CeO2/TiO2
NPs
Hydrothermal methodTiO2—3.18 eV,
CeO2—2.88 eV,
CeO2(10%)/TiO2—2.30 eV
1 g/Lp-chlorophenol (p-CP), 0.47 mM1000 W tungsten lamp (>420 nm),
250 W Hg lamp (365 nm)
UV (CeO2/TiO2)
p-CP—95.3%,
TOC—85.3%
(2.5 h)
VIS (CeO2/TiO2)
p-CP—57.5%,
TOC—48.9%
(6 h)
[168]
CeO2/TiO2Hydrothermal methodTiO2—3.18 eV,
CeO2/TiO2—2.67 eV
Bromophenol dye300 W Xe lamp (>400 nm)TiO2—6%,
CeO2/TiO2—72%
(180 min)
[156]
CeO2/TiO2
NTAs
Anodization and microwave homogeneous synthesisTiO2—3.2 eV,
CeO2/TiO2—2.72 eV
1.1 g/L,
in the presence of H2O2
Benzothiophene in
n-octane, 200 ppm (sulfur content)
300 W Xe lamp with UV cut filterCeO2—35%,
TiO2—42%,
CeO2/TiO2—90%
(5 h—desulfurization)
[169]
CeO2@TiO2
Core-shell
Hydrothermal route with the Stöber methodCeO2—2.84 eV,
TiO2—2.91 eV,
CeO2@TiO2—2.73 eV
1 g/LRhB, 0.01 mM300 W Xe lamp (>400 nm)CeO2—19%,
TiO2—17%,
CeO2@TiO2—57%
(60 min)
[170]
CeO2/TiO2Hydrothermal synthesisCeO2—3.18 eV,
CeO2/TiO2—2.17 eV
0.5 g/LRhB, 1 mM15W/G15 T8 (<400 nm, Phillips), 0.15 W/m2,
18W/541M7 (>400 nm, Phillips, 14.5 W/m2
Solar light—99.89%
(8 h)
UV—93%
(2 h)
Antibacterial activity
[17]
CeO2/TiO2
(0.05:1)
Peroxo sol-gel method 0.1 g/LMB, 10 mg/L10 W UVC lamp (254 nm),
18 W Germicidal lamp(>420 nm)
UVC
TiO2—77%,
CeO2/TiO2—98%,
VIS
TiO2—55%,
CeO2/TiO2—82%
(6 h)
[165]
CeO2-TiO2/SiO2Co-precipitation methodTiO2—2.92 eV,
CeO2-TiO2/SiO2—2.51 eV
0.1 g/LPhenol, 30 mg/L350 W solar simulator (300–2500 nm)TiO2—38%,
CeO2-TiO2/SiO2—96.5%
(180 min)
[161]
TiO2/CeO2
Core-shell NTs
Sol-gel technique 15 L/m2MO, 5 mg/L,
pH = 3
300 W Hg lamp (254 nm)CeO2—~30%,
TiO2—~70%,
TiO2/CeO2—~80%
(180 min)
[31]
Ti3+-TiO2/Ce3+-CeO2Hydrothermal route combined with wet-chemical deposition precipitation and in-situ solid-state chemical reductionCeO2—3.00 eV, TiO2—3.20 eV,
Ti3+-TiO2/Ce3+-CeO2—2.70 eV
1 g/LMO, 10 mg/L,
MB, 10 mg/L
300 W Hg lamp (>420 nm)MO—93.3%,
MB—97.1%
(180 min)
[30]
TiO2@h-CeO2
Yolk-shell microspheres
Calcination methodTiO2—3.22 eV,
h-CeO2—2.95 eV
h-CeO2—hollow CeO2
0.1 g/LMB, 0.01 mM125 W Hg lampMB—50%
h-CeO2—14 min,
TiO2—13 min,
TiO2@h-CeO2—7 min
[171]
CeO2@TiO2Precipitation-co-hydrothermal methodTiO2—3.15 eV,
CeO2—2.73 eV,
CeO2@TiO2—2.42 eV
1 g/LRhB, 10 mg/L500 W Xe lamp,
100 W Hg lamp
UV
CeO2—60%,
TiO2—82%,
CeO2@TiO2—95%
(60 min)
UV
CeO2—58%,
TiO2—50%,
CeO2@TiO2—75%
(180 min)
[172]
TiO2@CeO2 0.4 g/LCN, 4.7 mM,
pH = 12.5
300 W Hg lamp (>420 nm)CeO2—6.4%,
TiO2—25.5%,
TiO2@CeO2—96.2%
(90 min)
[173]
CeO2/TiO2Evaporation-induced self-assembly methodTiO2—3.10 eV,
CeO2/TiO2—1.85 eV
1.33 g/LMB, 10 ppm,
4-chlorophenol (4-CP), 0.1 mM
300 W tungsten lamp (400–660 nm)MB
TiO2—9%,
CeO2/TiO2—75%
(3 h)
4-CP
TiO2—0%,
CeO2/TiO2—66%
(5 h)
[162]
CeO2-TiO2Co-precipitation methodTiO2—2.82 eV,
CeO2-TiO2—2.30 eV
0.1 g/LDiisopropanolamine (DIPA), 1000 ppm
With the addition of H2O2
500 W halogen lampDIPA
TiO2—66%,
CeO2-TiO2—82%
COD
TiO2—43.2%,
CeO2-TiO2—54.8%
(5 h)
[174]
CeO2/TiO2-NTsHydrothermal methodTiO2-NTs—3.08 eV,
CeO2—2.93 eV
CeO2/TiO2-NTs—2.64 eV
0.8 g/LMB, 15 mg/L250 W Hg lampMB
CeO2—10%,
TiO2-NTs—92%
CeO2/TiO2-NTs—~100%
COD
CeO2/TiO2-NTs—62%
(2 h)
[167]
CeO2/TiO2Hydrothermal routeTiO2—3.20 eV,
CeO2/TiO2—3.23 eV
0.5 g/LPolyvinylpyrrolidone (PVP), 300 mg/L,
MB, 20 mg/L
Visible light (400–800 nm),
Hg lamp, 41 W/m2
PVP—VIS—4 h
TiO2—4%,
CeO2/TiO2—2%,
PVP—UV—4 h
TiO2—23%,
CeO2-TiO2—7%
MB—VIS—1.5 h
TiO2—6%,
CeO2/TiO2—92%
MB—UV—1.5 h
TiO2—99%,
CeO2/TiO2—90%
[175]
TiO2-CeO2Calcination methodCeO2—2.01 eV,
TiO2—3.26 eV,
TiO2/CeO2—3.28 eV
0.4 g/LPhenazopyridine (PhP), 12 mg/L15 W Hg lamp (254 nm)CeO2—17%,
TiO2—55%,
TiO2/CeO2—66%
(20 min)
[166]
Abbreviations: RhB—rhodamine B; MO—methyl orange; MB—methylene blue; TOC—total organic carbon; COD—chemical oxygen demand; NPs—nanoparticles; NTs—nanotubes; and NTAs—nanotube arrays.
Table 4. Photocatalytic degradation of various pollutants with the application of CeO2 composites with carbon materials.
Table 4. Photocatalytic degradation of various pollutants with the application of CeO2 composites with carbon materials.
CompositePreparation MethodBand Gap, eVCatalyst DosePollutant, ConcentrationLight SourceDegradation Efficiency (Process Time)Ref.
rGO-CeO2Hydrothermal methodCeO2 NPs—2.94 eV,
rGO-CeO2—2.91 eV
1 g/LMB,
0.01 mM
Direct sunlightrGO—3.5%,
CeO2 NPs—35%,
rGO-CeO2—~72%
(50 min)
[87]
rGO/CeO2In-situ growth and self-assembly approach 0.3 g/LMB, 20 mg/L,
with addition of 1 mL 30% H2O2
500 W tungsten lampRGO—7.8%,
CeO2—10%,
RGO/CeO2—74.8%
(140 min)
[187]
CeO2HS/rGO
HS—hollow nanospheres
Hydrothermal method 0.5 g/LMO, 30 mg/L
pH = 6.5
800 W Hg lampCeO2HS—16%,
rGO—22%,
GO—25%,
CeO2HS/rGO—97%
(50 min)
[177]
ZnO-CeO2-rGO
Possible H2 generation in PEC
Hydrothermal processZnO-rGO—3.04 eV,
ZnO-CeO2-rGO—2.20 eV
1 g/LMB, 100 mM,
pH = 3
Xe lamp (AM 1.5G filter),
100 mW/cm2
MB
ZnO-rGO—20%,
ZnO-CeO2-rGO—90%,
TOC
ZnO-rGO—11%,
ZnO-CeO2-rGO—72%
(90 min)
[188]
rGO-CeO2 NCsHydrothermal processCeO2NWs—2.88 eV,
CeO2NCs—2.76 eV
0.375 g/LMB, 0.01 mM500 W Hg lamp, 40 mW/cm2CeO2NWs—58%,
CeO2NCs—65%,
RGO-CeO2NCs—87%
(90 min)
[33]
CeO2/Graphene NPsHydrothermal methodCeO2NPs—3.16 eV,
CeO2-Graphene—2.69 eV
0.08 g/LCR, 5 mg/L
MB, 5 mg/L
400 W lamp, (>500 nm)CR
Graphene—~16%,
CeO2-Graphene—~94.5%
MB
Graphene—~20%,
CeO2-Graphene—~98%
(180 min)
[119]
UCNCs@SiO2@CeO2:Tm/GN
GN-graphene
Core/shell structure upconversion nanocrystals—UCNCs
Two-step wet-chemical route 0.4 g/LRhB, 20 mg/L500 W Xe lamp
(300–2500 nm)
CeO2—46%,
CeO2:Tm—55%,
UCNCs@SiO2@CeO2:Tm—82%,
UCNCs@SiO2@CeO2:Tm/GN—95%,
(210 min)
[24]
CuO-CeO2/GO
GO—graphene oxide
Dip-molding and ultrasound-assisted processCeO2—3.28 eV,
CuO—2.61 eV
0.5 g/LMO, 10 mg/L,
with addition of 0.2 mL 3% H2O2
300 W Xe lamp (>400 nm)CuO—65.7%,
CuO/GO—79.8%,
CuO-CeO2/GO—97.8%
(150 min)
[189]
CeO2/g-C3N4Mixing-calcination techniqueCeO2—2.89 eV,
g-C3N4—2.70 eV
1 g/LMB, 10 mg/L
4-CP, 5 mg/L
300 W Xe lamp
(>400 nm)
MB
CeO2—2.8% (TOC—16%),
g-C3N4—75% (TOC—49%),
CeO2/g-C3N4—95% (TOC—81%)
(3 h)
4-CP
CeO2—15.1%,
g-C3N4—2.3%,
CeO2/g-C3N4—45%
(5 h)
[190]
CeO2/g-C3N4In-situ co-pyrolysis methodCeO2—2.82 eV,
g-C3N4—2.70 eV
1 g/LPhenol, 10 mg/L500 W Xe lamp
(>420 nm)
CeO2—1%,
g-C3N4—3%,
CeO2/g-C3N4—55%,
(5 h)
[191]
g-C3N4/CeO2Calcination methodCeO2—2.98 eV,
g-C3N4—2.78 eV,
g-C3N4/CeO2—2.6 eV
0.06 g/LMB, 20 mg/L50 W fluorescent lamp
(>400 nm)
CeO2—8%,
g-C3N4—20%,
g-C3N4/CeO2—70%
(4 h)
[159]
g-C3N4@CeO2
core-shell structure
Hydrothermal methodg-C3N4—2.82 eV,
CeO2—2.76 eV
1 g/LDoxycycline hydrochloride, 0.01 g/L,
with addition of 100 μL H2O2
150 W Xe lamp (>400 nm)g-C3N4—66.7%,
CeO2—71.7%,
g-C3N4@CeO2—84%
(1 h)
[192]
CeO2/S-doped g-C3N4
(CeO2/CNS)
One-pot thermal condensation methodCNS—2.55 eV,
CeO2—2.80 eV
1 g/LMB, 10 mg/L300 W halogen lamp (>400 nm), 21.9 mW/cm2CeO2—39%,
CNS—54%,
CeO2/CNS)—91.4%
(150 min)
[193]
CeO2/ATP/g-C3N4
ATP—attapulgite
Electrostatic-induced self-assembly methodCeO2/ATP—3.2 eV,
g-C3N4—2.7 eV,
CeO2/ATP/g-C3N4—2.55 eV
m(catal.)/m(DBT) = 1:10Dibenzothiophene (DBT), 200 ppm (sulfur conc.),
with addition of 30% H2O2
300 W Xe lamp
(>420 nm)
Desulfurization
g-C3N4—42%,
CeO2/g-C3N4—83%,
CeO2/ATP/g-C3N4—98%
(3 h)
[194]
g-C3N4/CeO2/ZnOPyrolysis and subsequent exfoliation method 0.375 mg/LMB, 10 mg/L5 W fluorescent lamp (>400 nm),
300 W UV lamp
UV
ZnO—50%,
g-C3N4—40.1%,
g-C3N4/CeO2—62%,
g-C3N4/CeO2/ZnO—98.9%
(25 min),
VIS
ZnO—22%,
g-C3N4—20%,
g-C3N4/CeO2—32%,
g-C3N4/CeO2/ZnO—52%
(4 h)
[195]
g-C3N4/α-Fe2O3/CeO2Hydrothermal techniqueg-C3N4—2.7 eV,
α-Fe2O3—1.98 eV,
CeO2—2.81 eV,
g-C3N4/α-Fe2O3—2.51 eV,
g-C3N4/α-Fe2O3/CeO2—2.34 eV
50 mgMB, 30 ppm500 W Xe lamp (>420 nm)g-C3N4—36%,
g-C3N4/α-Fe2O3—59%,
g-C3N4/α-Fe2O3/CeO2—97.5%
(120 min)
[196]
Abbreviations: RhB—rhodamine B; MO—methyl orange; MB—methylene blue; CR—Congo red; TOC—total organic carbon; NPs—nanoparticles; NCs—nanocubes; rGO—reduced graphene oxide; and g-C3N4—graphitic carbon nitride.
Table 5. Photocatalytic degradation of various pollutants with the application of CeO2 composites with other materials.
Table 5. Photocatalytic degradation of various pollutants with the application of CeO2 composites with other materials.
CompositePreparation MethodBand Gap, eVCatalyst DosePollutant, ConcentrationLight SourceDegradation Efficiency (Process Time)Ref.
CeO2/Bi2WO6Homogeneous precipitation coupled with the hydrothermal methodCeO2—2.58 eV,
Bi2WO6—3.1 eV,
0.4CeO2/0.6Bi2WO6—3.15 eV
1 g/LRhB, 0.02 mM50 W halogen lamp—simulated solar lightCeO2—10.26%,
Bi2WO6—61.46%,
0.4CeO2/0.6Bi2WO6—75.94%
(75 min)
[32]
CeO2/Bi2WO6Precipitation coupled with the hydrothermal methodCeO2—2.93 eV,
Bi2WO6—2.86 eV,
CeO2/Bi2WO—2.60 eV
1 g/LRhB, 0.02 mM,
MB, 0.02 mM
50 W halogen lamp
(>400 nm)
RhB
CeO2—7.7%,
Bi2WO6—9.7%,
CeO2/Bi2WO6—54.1%
MB
CeO2—8.2%,
Bi2WO6—10.7%,
CeO2/Bi2WO6—15.6%
(75 min)
[100]
CeO2/Bi2WO6A doctor blading methodCeO2—2.70 eV,
Bi2WO6—2.97 eV,
0.4CeO2/0.6Bi2WO6—2.77 eV
3-layer films controlled by scotch tapeRhB, 0.002 mM50 W halogen lampCeO2—6.2%,
Bi2WO6, 37.9%,
CeO2/Bi2WO6—44.4%
(120 min)
[199]
CeO2@Bi2WO6
Hollow magnetic microcapsules
Template-assisted synthesis followed by H2 reductionCeO2—2.92 eV,
Bi2WO—2.68 eV,
0.4CeO62@0.6Bi2WO6—2.75 eV
0.5 g/L (Cr),
1 g/L (CN)
Cr(VI), 8 mg/L (pH = 5.7),
CN, 4.78 mM (pH = 12.5)
300 W Xe lamp (>420 nm)Cr(VI)
CeO2—85.2%,
Bi2WO6—91.2%,
CeO2@Bi2WO6—99.6%
(1 h),
CN
CeO2—6%,
Bi2WO6—8%,
CeO2@Bi2WO6—98.3%
(1 h)
[200]
CeO2/Bi4Ti3O12
(C-BTO)
Molten salt method and ion-impregnation methodCeO2—2.61 eV,
BTO—2.91 eV
1 g/LBisphenol A, 10 mg/L400 W halogen lamp,
580 mW/cm2
CeO2—54.8%,
BTO—70%,
C-BTO—96.8%
(60 min)
[201]
CeO2/Bi2MoO6
Hierarchical heterostructure microspheres
Solvothermal routeCeO2—2.47 eV,
Bi2MoO6—2.86 eV,
CeO2/Bi2MoO6—2.76 eV
1 g/LCO2 bubbled in ultrapure water300 W Xe lamp (>420 nm)Production
CH3OH—32.5 µmol/gcatal.,
C2H5OH—25.9 µmol/gcatal.
(4 h)
[202]
Bi2O3/CeO2Hydrothermal methodCeO2—3.1 eV,
Bi2O3—2.84 eV,
Bi2O3/CeO2—2.39 eV
Orange II dye, 0.1 mM150 W Xe lamp (>400 nm)CeO2—~26%,
Bi2O3—~27%,
Bi2O3/CeO2—~55%
(5 h)
[203]
BiOI/CeO2In-situ chemical bath methodCeO2—2.81 eV,
BiOI—1.70 eV
1 g/LMO, 10 mg/L,
Bisphenol A (BPA)—10 mg/L
300 W Xe lamp (>420 nm)MO
CeO2—11.9%,
BiOI—55.1%,
BiOI/CeO2—93.8%,
(50 min)
BPA
CeO2—28.1%,
BiOI—55.8%,
BiOI/CeO2—92.0%
(90 min)
[204]
CeOx/BiVO4Hydrothermal and ion-impregnation methodBiVO4—2.52 eV,
CeOx(5.7%)/BiVO4—2.48 eV
1 g/LMB, 0.03 mM100 W lamp
(>420 nm)
BiVO4—24%,
CeOx/BiVO4—61%
(150 min)
[205]
BiVO4/CeO2Homogeneous precipitation coupled with the hydrothermal methodCeO2—2.76 eV,
BiVO4—2.51 eV,
BiVO4/CeO2—2.46 eV
1 g/LMB, 0.02 mMHalogen lamp (>400 nm),
185 mW/cm2
CeO2—~20%,
BiVO4—~50%
BiVO4/CeO2—80%
(30 min)
[206]
Ag3PO4/CeO2Low-temperature solution combustion followed by annealingCeO2—3.2 eV,
Ag3PO4—2.45 eV
0.375 g/LMB, 10 mg/L
with addition of H2O2
300 W UV lamp, 55 W fluorescent lamp
(>400 nm)
UV
CeO2—40.6%,
Ag3PO4—92.4%,
Ag3PO4/CeO2—98.9%
(6 min)
VIS
CeO2—20.8%,
Ag3PO4—82.1%,
Ag3PO4/CeO2—98%
(60 min)
[207]
Ag3PO4/CeO2Ultrasound-assisted methodCeO2—2.98 eV,
Ag3PO4—2.45 eV,
Ag3PO4/CeO2—2.20 eV
1 g/LMB, 10 mg/L300 W Xe lampCeO2—20%,
Ag3PO4—88%,
Ag3PO4/CeO2—95%
(40 min)
[208]
Ag/Ag3PO4/CeO2Solvothermal method combined with co-precipitation and photoreductionCeO2—2.54 eV,
Ag3PO4—2.42 eV
1 g/LCO2 bubbled in ultrapure water300 W Xe lamp (>420 nm)Production
CH3OH—40 μmol/gcatal.,
C2H5OH—30 μmol/gcatal.
(4 h)
[209]
AgI/CeO2Sol-gel auto-combustion methodCeO2—2.81 eV,
AgI—2.76 eV
0.5 g/LRhB, 20 mg/L,
Tetracycline (TC), 20 mg/L
300 W Xe lamp (>420 nm)RhB
CeO2—20%,
AgI 48%,
AgI/CeO2—98%
(20 min)
TC
CeO2—46%,
AgI—74%,
AgI/CeO2—94%
(1 h)
[210]
Ag2O/CeO2Hydrolysis and calcination followed by thermal decompositionCeO2—2.72 eV,
Ag2O—1.30 eV
1 g/LEnrofloxacin (EFA), 10 mg/L300 W Xe lamp (>420 nm)EFA
CeO2—11.7%,
Ag2O—43%,
Ag2O/CeO2—87.1%
(120 min)
TOC
Ag2O/CeO2—66.8%
(120 min)
[211]
Ag2CO3/CeO2/AgBrHydrolysis and calcination followed by hydrobromic acid corrosionCeO2—2.72 eV,
Ag2CO3—2.43 eV,
AgBr—2.58 eV
0.4 g/LLevofloxacin, 10 mg/L300 W Xe lamp (>420 nm)CeO2—~11%,
Ag2CO3—46%,
Ag2CO3/CeO2—~69%
(60 min)
AgBr—~45%,
CeO2/AgBr—~53%,
Ag2CO3/CeO2/AgBr—88%
(40 min)
TOC—60.98%
(80 min)
[212]
Fe2O3-CeO2Precipitation methodCeO2—2.82 eV,
Fe2O3-CeO2—2.1 eV
2 g/LCR, 25 mg/L100 W tungsten lamp (>400 nm)CeO2—89%,
Fe2O3-CeO2—96% (adsorption—91%)
[34]
p-CuO/n-CeO2Combination of ribbon-ignition and calcination methodsCeO2—2.94 eV,
CuO—1.54 eV,
CuO/CeO2—2.59 eV
0.5 g/LRhB, 10 mg/L, with addition of 2 mL 30% H2O2350 W Xe lamp (>420 nm)CeO2—~70%,
CuO—~77%,
CuO/CeO2—86.2%
(12 min)
[157]
Cu2S/CeO2
Ag2S/CeO2
Precipitation methodCeO2—3.39 eV,
Ag2S—2.00 eV,
Ag2S/CeO2—2.51 eV
Cu2S—1.75 eV,
Cu2S/CeO2—2.62 eV
1 g/LRhodamine 6G—0.01 mMnatural sunlightCeO2 NPs—25%,
Cu2S/CeO2—44%,
Ag2S/CeO2—30%
(240 min)
[213]
CeO2/V2O5,
CeO2/CuO
Thermal decomposition methodCeO2—3.28 eV,
V2O5—2.66 eV,
CuO—1.83 eV,
CeO2/V2O5—2.62 eV,
CeO2/CuO—2.59 eV
1 g/LMB, 0.03 mM,
Textile effluent
250 W Philips lamp (532 nm)MB
CeO2—6.1%,
V2O5—27.5%,
CuO—33.4%,
CeO2/V2O5—64.%,
CeO2/CuO—70.1%
(210 min)
Textile effluent
CeO2/V2O5—76.9%,
CeO2/CuO—85.7%
(300 min)
[214]
CuBi2O4/CeO2Solid state methodCeO2—3.18 eV,
CuBiO4—1.38 eV
CuBiO4(30%)/CeO2(70%)—3.14 eV
0.5 g/LCR, 20 mg/L,
pH = 7
6 W UV lamp (365 nm)CeO2—14.92%
CuBi2O4—3.13%,
CuBi2O4/CeO2—83.05%
(100 min)
[215]
CeO2/MgAlDispersion of insoluble metal oxidesCeO2/MgAl—3.2eV1 g/LPhenol (Ph), 80 ppm,
4-chlorophenol (4-CP), 100 ppm
2,4,6-trichlorophenol (2,4,6-TCP), 100 ppm
Pen Ray Power Supply lamp (254 nm), 4.4 mW/cm2Ph—50% (7 h),
4-CP—96% (5 h),
2,4,6-TCP—90% (2 h)
[216]
CeO2/Al2O3Wet-chemical methodCeO2—3.0 eV,
CeO2/Al2O3—2.85 eV
50 mg in 5 ppm dye solutionCR, 5 ppm,
MO, 5 ppm
pH = 11
300 W tungsten lampCR—90%
(120 min)
MO—92%,
TOC—92%
(90 min)
[217]
NiO/CeO2Electrospinning techniqueCeO2—3.2 eV,
NiO—3.5 eV
0.375 g/LMB, 10 mg/L300 W UV lampNiO—47.3%,
CeO2—69.2%,
NiO/CeO2—96%
(40 min)
[218]
CeO2-ZrO2
3DOM—3 dimensionally ordered microporous structure
Sol-gel method combined with the decompression filling methodCeO2—2.73 eV,
ZrO2—3.25 eV,
CeO2-ZrO2—2.44 eV
1.7 g/LCR, 50 mg/L400W Xe lamp (>410 nm),
1000 W Xe lamp—simulated sunlight
CeO2
VIS—55%,
Sunlight—53%,
UV—55%,
ZrO2
VIS—24%,
Sunlight—52%,
UV—40%,
CeO2-ZrO2
VIS—72%,
Sunlight—68%,
UV—68%
(120 min)
[98]
CeO2/SnO2Wet-chemical methodCeO2-SnO2—3.4 eV0.1 g/LMB, 0.03 mM250 W lamp (532 nm)80%
(150 min)
[219]
ZnO/CeO2Thermal decomposition methodZnO—3.2 eV,
CeO2—3.25 eV
1 g/LMO, 0.03 mM,
MB, 0.03 mM,
Phenol (Ph), 0.03 mM,
Industrial textile effluent (Ite)
250 W lamp (532 nm)ZnO
MO—4.0%,
MB—4.7%,
Ph—1.9%
CeO2
MO—4.2%,
MB—5.5%,
Ph—2.9%
ZnO/CeO2
MO—95.9%,
MB—97.4%,
Ph—96.2%
(150 min)
Ite—ZnO/CeO2
TOC—90.2%
(6 h)
[220]
ZnO/CeO2Wet chemistry method with the calcination techniqueZnO—3.13 eV,
CeO2—2.86 eV
0.5 g/LRhB, 10 mg/L300 W Xe lamp (>400 nm), 0.95 mW/cm2CeO2—56%,
ZnO—38%,
ZnO/CeO2—96%
(80 min)
[158]
CeO2-ZnO-TiO2Sol-gel methodCeO2-ZnO-TiO2—3.13 eV0.2 g/LRhB, 5 mg/L300 W tungsten lamp (8500 lumen)80% (180 min)[221]
Mn3O4/CeO2One-step hydrothermal methodCeO2—2.94 eV,
Mn3O4—2.02 eV,
Mn3O4/CeO2—2.25 eV
0.67 g/LRhB, 10 mg/L300 W Xe lampCeO2—60%,
Mn3O4—50%,
Mn3O4/CeO2—93%
(180 min)
[222]
Tm3+:CeO2/palygorskite
Pal—palygorskite
Hydrothermal-deposition methodTm3+:CeO2—2.92 eV,
Pal—3.5 eV
0.5 g/LBisphenol A300 W Xe lamp (>420 nm)Pal—32%,
CeO2/Pal—33%,
Tm3+:CeO2/Pal—86%
(180 min)
[223]
CeO2/CdS-DETA
DETA—diethylenetriamine
Two-step hydrothermal methodCeO2—3.07 eV,
CdS-DETA—2.35 eV
0.5 g/L,
with 0.6% Pt as co-catalyst
0.35 M Na2S + 0.25 M Na2SO3300 W Xe lamp (>420 nm)H2 production,
14.84 mmol/(g⋅h)
[79]
CeO2/CdSElectrochemical method 0.5 g/L0.43 M Na2S + 0.5 M Na2SO4300 W Xe lamp, UV or VIS (>420 nm)H2 formation
UV
0.782 mmol/(g⋅h),
VIS
0.223 mmol/(g⋅h)
[224]
Abbreviations: RhB—rhodamine B; MO—methyl orange; MB—methylene blue; CR—Congo red; and TOC—total organic carbon.
Table 6. Application of CeO2 composites as photoelectrodes in photoelectrocatalytic (PEC) processes.
Table 6. Application of CeO2 composites as photoelectrodes in photoelectrocatalytic (PEC) processes.
PhotoelectrodePreparation MethodCounter ElectrodePollutant, ConcentrationSupporting ElectrolyteLight Source, Applied Voltage/CurrentDegradation Efficiency (Process Time)Ref.
CeO2/TiO2 NTAsElectrodeposition methodPtMO, 5 mg/L0.1 M Na2SO4350 W Xe lamp,
2.0 V vs. SCE
EC—23.2%,
PC—56.5%,
PEC—98.1%
(60 min)
[237]
EG-CeO2Impregnation ultrasonic agitationPt foil2,4-DCP, 50 mg/L0.1 M Na2SO4,
pH = 6.2
350 W Xe lamp (AM 1.5G filter), 0.1 W/cm2,
8 mA/cm2
PEC—98.7%,
TOC—92.6%
(3 h, 8 mA/cm2)
P—11.2%,
EC—43.9%,
PEC- 85.7%
(3 h, 2 mA/cm2)
[238]
rGO-CeO2-TNAsGalvanostatic methodPt foilBPA, 10 mg/L0.05 M Na2SO4500 W Xe lamp (>365 nm), 110 mW/cm2,
9 V
Fenton—~30%,
EC-Fenton—~30%,
P-Fenton—55%,
PC-Fenton—72%,
PEC-Fenton—82%
(120 min)
[239]
rGO-CeO2-TiO2 NTAsGalvanostatic methodPt foilTetrabromobisphenol A, 10 mg/L0.05 M Na2SO4500 W Xe lamp, simulated solar light—110 mW/cm2,
9 V vs. SCE
EC—~42%,
PC—~65%,
PEC—87%
(100 min)
[240]
PANI/CeO2/TiO2 NTAs
PANI—polyaniline
Electrochemical methodPt foilTetrabromobisphenol A, 10 mg/L0.05 M Na2SO4500 W Xe lamp, 120 mW/cm2,
9 V vs. SCE
TiO2—85.3%,
CeO2/TiO2—90.3%, PANI/TiO2—86.8%,
PANI/CeO2/TiO2—94.0%
(2 h)
[241]
CeO2 QDs/Ag2SePrecipitation routePt plateTetracycline, 0.02 g/L0.1 M Na2SO48 W halogen lamp (400–790 nm),
80 lm/W,
0.5 V vs. SCE
EC—27.7%,
PC—92.3%
(90 min)
PEC—95.8%
(75 min)
[242]
CeO2@α-Fe2O3 NTAsElectrodeposition methodPt foilTetracycline, 30 mg/L,
pH = 13
0.1 M NaOH300 W Xe lamp (AM 1.5G filter), 100 mW/cm2,
1.5 V vs. Ag/AgCl
PC—15.6%,
PEC—88.6%
(1 h)
[243]
Ti/SnO2-Sb-CeO2@TiO2-SnO2Sol-gel routePtMB, 20 ppm10 g/L Na2SO4500 W Xe lamp, 60.2 mW/cm2,
1.5 V
EC—63.1%,
PC—84.2%,
PEC—95.8%
(2 h)
[244]
Ce/Ce2O3/CeO2/TiO2 TNAsElectrochemical anodizationPt grid 0.1 M Na2SO4 + 10vol% ethylene glycol450 W Xe lamp, 100 mW/cm2,
0.7 V vs. OCP
H2 generation
UV
TNTs—2.6 mL (h⋅cm2),
TNTs-Ce-CeOx—5.0 mL/(h⋅cm2)
VIS
TNTs—1.2 mL/(h⋅cm2),
TNTs-Ce-CeOx—2.9 mL/(h⋅cm2)
[109]
ns-TiO2/CeO2/Ti
ns-nanostructured
Sol-gel processPt H2SO4 + K2SO4100W Xe lamp,
450 mW/cm2
H2 generation
ns-TiO2—8.2 l/(h⋅m2),
ns-TiO2-CeO2—13.8 l/(h⋅m2)
[245]
CeO2/CuO/CuElectrochemical methodPtCO2,
Flow rate—40 mL/min
0.1 M KHCO3500 W Xe lamp (420–800 nm),
100 mW/cm2,
−1.0 V vs. SCE
Methanol yield in μmol/(cm2)
PC—5.53,
EC—9.51,
PEC—22.32
(6.5 h)
[246]
CeO2/Cu2ODeposition routePt gauze 0.1 M NaOH150 W Xe lamp,
0.7 V vs. SCE
H2 generation
3.62 mL/h
[23]
Abbreviations: MO—methyl orange; MB—methylene blue; BPA—bisphenol A; TOC—total organic carbon; DCP—dichlorophenol; NTAs—nanotube arrays; OCP—open circuit potential; TNAs—TiO2 nanotube arrays; rGO—reduced graphene oxide; P—photolysis; PC—photocatalysis; EC—electrocatalysis; and PEC—photoelectrocatalysis.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Kusmierek, E. A CeO2 Semiconductor as a Photocatalytic and Photoelectrocatalytic Material for the Remediation of Pollutants in Industrial Wastewater: A Review. Catalysts 2020, 10, 1435. https://doi.org/10.3390/catal10121435

AMA Style

Kusmierek E. A CeO2 Semiconductor as a Photocatalytic and Photoelectrocatalytic Material for the Remediation of Pollutants in Industrial Wastewater: A Review. Catalysts. 2020; 10(12):1435. https://doi.org/10.3390/catal10121435

Chicago/Turabian Style

Kusmierek, Elzbieta. 2020. "A CeO2 Semiconductor as a Photocatalytic and Photoelectrocatalytic Material for the Remediation of Pollutants in Industrial Wastewater: A Review" Catalysts 10, no. 12: 1435. https://doi.org/10.3390/catal10121435

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