Next Article in Journal
Structural, Morphological and Optical Properties of MoS2-Based Materials for Photocatalytic Degradation of Organic Dye
Next Article in Special Issue
Mechanistic Insights into Graphene Oxide Driven Photocatalysis as Co-Catalyst and Sole Catalyst in Degradation of Organic Dye Pollutants
Previous Article in Journal / Special Issue
Progress in the Photoreforming of Carboxylic Acids for Hydrogen Production
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Solid State Nanostructured Metal Oxides as Photocatalysts and Their Application in Pollutant Degradation: A Review

by
Carlos Díaz
1,*,
Marjorie Segovia
2 and
Maria Luisa Valenzuela
3,*
1
Departamento de Química, Facultad de Ciencias, Universidad de Chile, Las Palmeras 3425, Santiago 7750000, Chile
2
Facultad de Ingeniería, Pontificia Universidad Católica de Chile, Vicuña Mackenna 4860, Santiago 8331150, Chile
3
Grupo de InvestigaciÓn en Energía y Procesos Sustentables, Instituto de Ciencias Químicas Aplicadas, Facultad de Ingeniería, Universidad Autónoma de Chile, Av. El Llano Subercaseaux 2801, Santiago 8910060, Chile
*
Authors to whom correspondence should be addressed.
Photochem 2022, 2(3), 609-627; https://doi.org/10.3390/photochem2030041
Submission received: 2 June 2022 / Revised: 22 July 2022 / Accepted: 23 July 2022 / Published: 5 August 2022

Abstract

:
Most dyes used in various industries are toxic and carcinogenic, thus posing a serious hazard to humans as well as to the marine ecosystem. Therefore, the impact of dyes released into the environment has been studied extensively in the last few years. Heterogeneous photocatalysis has proved to be an efficient tool for degrading both atmospheric and aquatic organic contaminants. It uses the sunlight in the presence of a semiconductor photocatalyst to accelerate the remediation of environmental contaminants and the destruction of highly toxic molecules. To date, photocatalysis has been considered one of the most appealing options for wastewater treatment due to its great potential and high efficiency by using sunlight to remove organic pollutants and harmful bacteria with the aid of a solid photocatalyst. Among the photocatalysts currently used, nanostructured metal oxide semiconductors have been among the most effective. This review paper presents an overview of the recent research improvements on the degradation of dyes by using nanostructured metal oxide semiconductors obtained by a solid-state method. Metal oxides obtained by this method exhibited better photocatalytic efficiency than nanostructured metal oxides obtained using other solution methods in several cases. The present review discusses examples of various nanostructured transition metal oxides—such as TiO2, Fe2O3, NiO, ReO3, IrO2, Rh2O3, Rh/RhO2, and the actinide ThO2—used as photocatalysts on methylene blue. It was found that photocatalytic efficiency depends not only on the bandgap of the metal oxide but also on its morphology. Porous nanostructured metal oxides tend to present higher photocatalytic efficiency than metal oxides with a similar band gap.

1. Introduction

Currently, researchers and scientists are becoming increasingly interested in metal oxide nanostructures due to their important technological applications in electronic and optoelectronic devices, sensors, medicines, and renewable energy sources. Reducing the size of materials (metal oxides) to nano-level imparts properties different from the bulk or crystalline forms; these nanoparticles demonstrate behaviors characteristic of isolated atoms and molecules [1,2,3,4,5].
Industrial plants generate increasing amounts of wastewater, which often causes severe environmental problems [6,7]. Wastewater produced in many industrial processes typically contains organic compounds that are toxic and not amenable to biological treatments [7]. There are several different types of organic pollutants, including dyes, phenols, biphenyls, pesticides, fertilizers, hydrocarbons, plasticizers, detergents, oils, greases, pharmaceuticals, proteins, carbohydrates, and so on [8].
Among the various physical, chemical, and biological technologies used in pollution control, advanced oxidation processes such as photocatalysis are being increasingly adopted in the destruction of organic contaminants due to their high efficiency, simplicity, good reproducibility, and ease of handling [8]. Heterogeneous photocatalysis possesses some critical advantages that have feasible applications in wastewater treatments, including: ambient operating temperatures and pressures, and the complete mineralization of contaminants and their intermediate compounds without leaving secondary pollutants. However, for large scale municipal applications, current photocatalytic water treatment systems are less attractive because they are more time consuming and have higher costs than other existing advanced oxidation techniques such as UV/H2O2, O3/H2O2, and UV/O3 technologies [9].
Therefore, semiconducting metal oxides have high potential due to their capability to generate charge carriers when stimulated with the required amount of energy, and to their applications in environmental remediation and electronics. Some semiconducting metal oxide synthesis methods are the chemical vapor deposition technique, the hydrothermal method, the laser ablation technique, and the electro-deposition method [10,11]. Out of these techniques, the hydrothermal method is the most user-friendly because of its reduced cost and ease of handling as well as being chemically reactive at low temperatures. However, other methods including the thermal decomposition of solids such as molecular complexes and macromolecular complexes [12] are gaining relevance due to their ease of preparation and final product purity.
An example of environmental remediation processes is the degradation of organic dyes [13] and the generation of clean energies such as “green hydrogen” [14].
In this context, one of the main applications of nanostructured metal oxides is the photocatalytic degradation of organic pollutants, which is in the field of environmental remediation.
The main characteristics that a suitable metal oxide photocatalytic system must have are [15]:
  • An adequate bandgap;
  • Suitable morphology;
  • High surface area;
  • Stability and reusability.
Semiconductor metal oxides with a bandgap near 3.2 eV are UV-light active, while semiconductor metal oxides with a bandgap near 2.7 eV are Visible-light active [16,17]. Metal oxides exhibiting these features—such as oxides of vanadium, chromium, titanium, zinc, tin, and cerium—follow similar primary photocatalytic processes such as light absorption, which induces a charge separation process with the formation of positive holes that can oxidize organic substrates. In this process, a metal oxide is activated with either UV light, visible light, or a combination of both, and photoexcited electrons are promoted from the valence band to the conduction band, forming an electron/hole pair (e/h+). The photogenerated pair (e/h+) can reduce and/or oxidize a compound adsorbed on the photocatalyst surface. A schematic representation of this process is shown in Figure 1.
As Figure 1 shows, the bandgap of the metal oxides dictates their photocatalytic activity; crucially, it establishes whether their energy matches or exceeds the band gap energy of the semiconductor. In Table 1, a summary of the bandgap values for the most common metal oxides along the periodic table is displayed. It is important to note that the bandgap values vary with the size of the nanoparticle, preparation method, doping, etc.
The photocatalytic activity of metal oxides comes from two sources [13]:
(i)
Generation of OH radicals by oxidation of OH anions;
(ii)
Generation of O2 radicals by reduction of O2.
Nanostructured metal oxide semiconductors have been widely used in photocatalytic redox processes because of their filled valence band (VB) and empty conduction band (CB) electronic configurations. When exposed to a photon with energy exceeding the bandgap (hν > Eg), it generates an electron-hole pair with one electron in VB pumped into CB, leaving a hole behind in VB. The holes generated in VB have great oxidation capability, while the electrons in the CB have high reduction potential.
These highly-reactive electrons and holes participate in the photocatalytic organic degradation.
As mentioned above, the factors relevant for a photocatalyst to be efficient include adequate bandgap, suitable morphology, high surface area, stability, and reusability.
Whether a given nanostructured metal oxide meets these characteristics depends on its preparation method. For instance, TiO2 is one of the most used and efficient metal oxides for the photocatalytic degradation of several organic dye pollutants [23]. However, its relative efficiency depends on the preparation method, which in turn determines its bandgap, morphology, surface area, stability and reusability.
For instance, the dependence of the catalytic effectiveness of TiO2 on the preparation method is illustrated in Table 2.
The Eg bandgap also exhibits an inverse dependence on particle size. For instance, theoretical studies have explained this relationship with the following expression [25]:
E g * E g b u l k + 2 π 2 2 e r 2 1 m e + 1 m h 1.8 e 2 4 π ε ε 0 r
where Eg bulk is the bulk energy gap, r is the particle radius, me is the effective mass of the electrons, mh is the effective mass of the holes, ε is the relative permittivity, ε0 is the permittivity of free space, is Plancks constant divided by 2π, and e is the charge of the electron.
The experimental inverse, Eg vs. particle size, has been observed for several nanostructured metal oxide nanoparticles [25,26,27], including ZnO, Fe2O3, Co3O4, NiO and SnO2.
On the other hand, the bandgap can also be modified by doping. In general, doping decreases the bandgap, as is reported in the case of ZnO/Mn2+, ZnO/Cu and TiO2/doping, among others [24]. Some of the most studied are the TiO2/dopant and ZnO/dopant. Table 3 and Table 4 summarize some of the results of MxOy/dopant systems. One of the explanations for this decrease in bandgap after doping is the increase in surface area. The increased surface area of the photocatalyst also increases the photodegradation of dye pollutants. This is because materials with a large surface area have more active sites than materials with a low surface area. This is illustrated in Table 3.
As mentioned before, ZnO is one of the most common and effective photocatalysts for the photodegradation of organic dye pollutants. ZnO is insoluble in water, existing in the form of a white powder, and it is extracted from the mineral Zincite through synthetic methods. Plastic, rubber, glass, ceramic, and cement industries use ZnO in various processes. ZnO is an n-type semiconductor with high bandgap energy that belongs to group II–IV of the periodic table [26,27]. ZnO is chemically stable, nontoxic, and inexpensive. It possesses high radiation hardness, effective transparency, remarkable optical absorption in the UV range, and outstanding thermal properties. Due to these characteristics, ZnO is used in a number of applications, such as optoelectronics, solar cells, medicine, catalysis, photocatalysis, and as an antibacterial agent. Doping in ZnO produces a decrease in the bandgap, which enhances photocatalytic efficiency. In Table 4, some doped ZnO are shown.
Titanium dioxide, TiO2, is also known as Titania or titanium (IV) oxide. TiO2 is the most promising photocatalyst since it possesses long-term chemical and physical stability along with low production costs [28]. It has a wide bandgap (3.0–3.2 eV) in the UV light range and exhibits three polymorphs, namely rutile, anatase, and brookite. Among these three phases, rutile is the most stable phase; the others are metastable, but they can transform into rutile phase irreversibly at elevated temperatures. Furthermore, the TiO2 photocatalyst has important uses in water treatment, lithium-ion batteries, sensors, catalysis, antibacterial, and anticancer applications [29,30,31]. Table 5 shows some examples of TiO2/dopants and their photocatalytic activity toward different organic dye pollutants.

2. Materials and Methods

Nanostructured Metal Oxides Preparation Methods

Several solution methods have been developed for synthesizing metal or metal oxide nanoparticles, such as the solvothermal, hydrothermal, and sol-gel methods [10,11]. Several of these nanostructured metal oxides have been suitable for photocatalytic degradation of organic dye pollutants. Usually, the preparation of nanostructured metal oxides for photocatalysis involves some solution method. However, for some of these methods, drawbacks like the aggregation of the NPs during the adsorption stage cause a reduction in the efficiency of the desired application [32,33,34]. When the metal or metal oxide is obtained as a colloid, as is commonly observed, their separation by centrifugation often produces agglomeration of the NPS [34]. To solve this type of problem, a solid-state method is used in which the metal nanoparticles are generally obtained as solid pure phase material. Therefore, the solventless synthesis of nanostructures is highly significant due to its economical, eco-friendly, and industrially viable nature. However, the development of new solid-state methods to prepare metallic nanostructured materials is a constant challenge. We have previously described a new solid-state method to synthesize metallic nanostructured nanomaterials from the pyrolysis of metallic and organometallic derivatives of poly- and oligo-phosphazene under air and at 800 °C [35,36,37,38]. Metallic nanostructured materials of the type M, MxOy, and MxPyOz are obtained depending on the nature of the metal. Another method—used when the respective metallic or organometallic derivative is not possible to prepare—uses MLn/[NP(O2C12H8)]3 mixtures [39,40,41]. However, in several of these systems, the M or MxOy phase is accompanied by a phosphate phase. These methods have been discussed in detail in several papers [35,36,37,38,39,40,41] and books [32,33,42,43,44]. When we used a polymer that does not contain phosphorus in its polymeric chain—such as the chitosan MXn and PS-co-4-PVP MXn macromolecular precursors (MXn = metallic salt)—and then subjected it to solid-state pyrolysis under air and at 800 °C (see Figure 2), pure phase nanostructured MxOy were obtained [45,46,47,48,49,50,51,52,53,54,55,56,57]. Now we will discuss the use of these nanostructured metal oxides MxOy obtained by this method in the degradation of organic dye pollutants.

3. Results

3.1. TiO2

The titania to be used as photocatalyst was obtained by pyrolysis of the (Cp2TiCl2)(Quitosano) (I),(Cp2TiCl2)(PS-co-4-PVP) (II), (TiOSO4)(Chitosan) (III), (TiOSO4)(PS-co-4-PVP) (IV), (TiO(acac)2)(Chitosan) (V), and (TiO(acac)2)(PS-co-4-PVP) (VI) macromolecular precursors under air and at several temperatures (500 °C, 600 °C, 700 °C, and 800 °C [23]. The bandgaps of the TiO2 obtained using these precursors are displayed in Table 6. The values range from 3.21 to 3.72 eV, which are adequate for photocatalytic degradation of organic dyes using UV-visible irradiation.
The results of the photocatalytic degradation of methylene blue are shown in Table 7.
As can be seen in Table 7, the most efficient catalytic degradation arises from the TiO2 produced from the TiO2-A-(III)p(800) system precursors, with 87% degradation after 25 min. Based on this table, the photocatalytic efficiency could be expressed by the following multifactorial relationship:
PE = a bandgap + b particle morphology + c particle size + d Titania phase + e pyrolysis temperature
where PE is photocatalytic efficacy and the Titania phase can be anatase, rutile, or brookita. In the case of TiO2 obtained by our solid-state method using TiO2-A-(III)p(800) as the precursor, factor b is highly porous (see Figure 3) since, as it is known, the anatase form is the most efficient phase as a photocatalyst and, in solid-state, the temperature of the thermal treatment is crucial to determine morphology and particle size. In this case, factor a does not appear to be very important since, as shown in Table 7, the values are similar.
The results shown in Table 7 were obtained at pH 6.5; however, a pH effect study showed that, at alkaline pH 9.5, the degradation was 98% [23].

3.2. Fe2O3

Fe2O3 was prepared by pyrolysis of the Chitosan(FeCl2)y, Chitosan(FeCl3)y, PS-b-4-PVP(FeCl2)y, and PS-b-4-PVP(FeCl3)y macromolecular complexes under air and at 800 °C [47]. The product was Fe2O3 in the hematite phase in all the cases. The bandgap of the as-prepared Fe2O3 ranged from 1.83 eV to 2.12 eV, which indicated an effective photocatalyst behavior under UV-visible irradiation (see Table 8).
Some parameters for the degradation experiments of MB with α-Fe2O3 obtained from PS-co-4-PVP and chitosan macromolecular precursors are summarized in Table 9. The highest extent of degradation of MB (98.6 %) at 150 min of irradiation time was achieved for α-Fe2O3 obtained from the PS-co-4-PVP(FeCl2)y precursor in a 1:1 ratio.
The size and morphology of the two systems were determined by electron microscopy studies, showing fused nanoparticles and 3D networks with average sizes of 150–200 nm for α-Fe2O3 from the precursor Chitosan(FeCl2) 1:1 and 55–100 nm for α-Fe2O3 from the precursor PS-co-4-PVP(FeCl2) 1:1. These morphologies show a random network of the surface area, quasi-linear nanoparticle chains that fold into larger porous powder particles [47]. It appears to be that for these Fe systems, the multifactorial relationship that holds this behavior could be the following:
PE= a bandgap + b particle morphology + c particle size + d (Fe precursor)
where the (Fe precursor) factor refers to the FeCl2 or FeCl3 metal salt source linked to the polymer. Once again, the most important factor dictating the photocatalyst efficiency of α-Fe2O3 toward the degradation of methylene blue is the b factor, more specifically, the material porosity. Table 9 shows that the bandgaps of α-Fe2O3 are similar; however, factor a could be less important. On the other hand, factors c and d could be somewhat significant.

3.3. NiO

NiO is a p-type semiconductor with Eg = 3.5 eV with multiple practical applications [57,58,59,60]. However, its bandgap can be modified by doping with other metal oxide semiconductors or by formation of nanocomposites in which one of the components would be in less quantity (mainly its photocatalyst properties) [59,60]. NiO has been used in catalysis, battery cathodes, fuel-cell electrodes, electrochromic films, electrochemical supercapacitors, and magnetic materials [59,60]. As these applications depend on its bandgap value and, in turn, the bandgap properties of superconductors depend on the environment of the metal oxide materials [61,62], there are no systematic, detailed studies on the effect of the medium on the bandgap. Although some MxOy/M’xO’y have been prepared—where MxOy is a nanostructured metal oxide and M’xO’y is a metal oxide matrix—no systematic effect of the matrix on the photocatalytic efficiency toward the degradation of organic dyes has been reported. By preparing the NiO/SiO2, NiO/TiO2, NiO/Al2O3, and NiO/glasses nanocomposites through a solid-state method, we have investigated the effect of the SiO2, TiO2, Al2O3, and Na4.2Ca2.8(Si6O18)(glass) matrices on the photocatalytic properties of NiO toward the degradation of methylene blue [57]. The composites were prepared by solid-state thermal treatment of the Chitosan(NiCL2)x/M’xO’and PS-co-4-PVP(NiCL2)x//M’xO’y precursors, where M’xO’y could be SiO2, TiO2, and Al2O3 glasses. Kinetic data for the photodegradation process of MB with NiO and with the NiO/SiO2, NiO/TiO2, NiO/Al2O3, and NiO/glasses composites are displayed in Table 10.
As can be observed in Table 6, the NiO from the precursor containing the chitosan as a solid-state template produces much higher activity than the NiO sample arising from the PSP-4-PVP(NiO)x precursor. The polymer precursor clearly influences the photocatalytic activity. In this sense, using chitosan as a solid-state template produces much higher activity than the NiO sample arising from the PSP-4-PVP(NiO)x precursor. On the other hand, the most efficient photocatalytic activity was observed for the NiO/TiO2 composite, with 90% degradation of methylene blue in 5 h.
In general, the photocatalytic activity shown in Table 10 is probably related to a matrix effect of NiO inside the different SiO2, TiO2, and Al2O3 matrices. Different matrices can influence NiO in different ways. For example, when the matrix is TiO2, a p-n junction p-NiO/TiO2 can be formed; this reduces the recombination rate of the photogenerated electron-hole pairs, which is known to enhance the photocatalytic activity of TiO2 (see Figure 4) [60]. In the case of the NiO/TiO2 composite, which has the highest catalytic behavior, the matrix effect may arise from the enhanced activity of TiO2 by NiO. Therefore, in this case, NiO acts as a matrix in the active semiconductor for the degradation of methylene blue. On the other hand, the NiO/Al2O3 compound is one of the least efficient photocatalysts for the degradation of methylene blue, which is probably due to an insulating effect caused by Al2O3 that cuts off the communication between p-NiO and methylene blue to start the degradation of the dye. This is also in agreement with the low photocatalyst efficiency of the insulating SiO2 matrix. An explanation for the results of the photocatalytic activity of the NiO/Na4.2Ca2.8(Si6O18) composite obtained from the Chitosan(NiCl2·6H2O)x//Na2O·CaO·SiO2 precursor is that Na4.2Ca2.8(Si6O18) has similar characteristics to those of SiO2 as a matrix.

3.4. Precious Metal Oxides Ir, Rh, Re: IrO2, Rh/RhO2, Rh2O3, ReO3

Among the metals in the periodic table, the so called precious—such as Ir, Rh, and Re, among others—are some of the most catalytically active [63]. Their activity is hugely enhanced at the nano-level [64,65]. These metals, as well as their metal oxides, exhibit high catalytic activity [63,65].

3.5. Ir

IrO2 is a promising conducting oxide used, for example, as an electrode material in ferroelectric capacitors for nonvolatile memory applications [66]. However, to the best of our knowledge, there are no reports on its photocatalytic activity. The only related studies are those about the catalytic reduction of 4-Nitrophenol [67] and its electrochemical catalytic activity toward oxygen evolution [68]. As with the previous metal oxides, the iridium oxide for the catalytic experiments was prepared by thermal treatment of the Chitosan(IrCl3)x and PSP-4-PVP(IrCl3)x macromolecular precursors [55]. The as-prepared IrO2has an adequate bandgap for the degradation of methylene blue under UV-visible irradiation (see Table 11). As previously mentioned, we have used this dye as a model dye to assay the photocatalytic properties of IrO2. Kinetic data for the photodegradation process of MB with IrO2 are displayed in Table 11. IrO2 from the PVP precursor exhibits better photocatalytic activity (57% photodegradation in 300 min) than the IrO2 from the chitosan precursor (38% photodegradation in 300 min). This effect was associated with the more porous morphology of IrO2 from the PVP precursor [55].

3.6. Rh

In this case, the Rhodium oxides Rh/RhO2 and Rh/Rh2O3 were obtained by pyrolysis of the PSP-4-PVP(RhCl3)x and Chitosan(RhCl3)x macromolecular precursors, respectively [56]. The bandgap values (see Table 12) are adequate for photocatalytic activity under UV-visible irradiation. The results (see Table 11) indicate that the Rh/RhO2 sample exhibits better photocatalytic activity (78% photodegradation in 300 min) than the Rh/Rh2O3 one (70% photodegradation in 300 min). As the porosity of both materials is similar, the higher photocatalytic activity of Rh/RhO2 could be due to its small particle size in the range 10–20 nm [56].

3.7. Re

ReO3 for the photocatalytic essays was obtained by solid-state thermal treatment of the Chitosan(ReCl3)x and PSP-4-PVP(ReCl3)x macromolecular precursors [54]. The bandgap for ReO3 indicates an adequate value for UV-visible photoactivation (see Table 12). ReO3 was also found to catalyze the photodegradation of MB with an efficiency of 53% and 64% in 300 min for the chitosan and PVP precursors, respectively (see Table 11). ReO3 is an unusual transition metal oxide as it presents a metallic behavior with conductivity close to that of copper [69]. Therefore, nanoparticles of rhenium trioxide produce a SERS (surface-enhanced Raman spectra) effect on some organic compounds, such as pyridine [70]. Despite this, their catalytic activity has only been proved in the catalytic degradation of methyl orange [71].

3.8. Th

Among the actinides oxides, thoria is an important and promising material used in ceramic catalyst sensor solid electrolytes, catalysis, optical materials, and the traditional nuclear industry [72,73,74,75]. The thorium oxide was prepared by thermal treatment of the Chitosan Th(NO3)4 and PS-co-4-PVPTh(NO3)4 macromolecular precursors at 800 °C. The ThO2/SiO2 and ThO2/TiO2 composites were also synthesized by pyrolysis of the Chitosan Th(NO3)4//MO2 andPS-co-4-PVP Th(NO3)4//MO2 macromolecular composites, where MO2 is SiO2 or TiO2.
To the best of our knowledge, no photocatalytic studies on ThO2, ThO2/SiO2, or ThO2/TiO2 have been reported. Regarding thoria, scarce literature data have been reported. As for ThO2 nanoparticles, Aller et al. [76] report values ranging from 5.13 eV to 4.5 eV, while for ThO2 thin films, Buono-Core et al. [77] report values from 4.5 eV to 4.61 eV. However, some of these values are theoretical calculations. Using the solid-state UV-visible absorption and the Tauc plot for ThO2, a value of 5.66 eV is estimated for the chitosan polymer precursor and 5.76 eV for the PS-4-co-PVP polymer precursor.
With this in mind, using UV-visible radiation, some catalytic activity for ThO2 could be expected. In fact, the values for the catalytic results are shown in Table 13. The values for the ThO2/SiO2 and ThO2/TiO2 nanocomposites are also shown for comparison. On the other hand, the bandgap values for thoria included in TiO2 as matrix exhibited values of 3.14 eV and 3.15 eV for the ThO2/TiO2 composites from chitosan and PS-4-co-PVP polymer precursors, respectively. These values are lower than those of ThO2 and ThO2/SiO2; however, as Buono-Core et al. point out [77], the optical bandgap energy is very sensitive to the preparation method and the experimental parameters applied in the synthesis [77]. In fact, Mahmoud reported a value of 3.82 eV for ThO2 prepared by a spray pyrolysis technique [78]. The estimated rate constant for the degradation of methylene blue in the presence of thorium prepared without the SiO2 and TiO2 matrices is greater than that of the pristine compounds, suggesting that the structural modification and synergy of the inorganic components play a fundamental role. This is related to the increase in the photocatalytic efficiency of the semiconductor as a result of the greater number of active sites provided by the (ThO2) PS-4-co-PVP and (ThO2) chitosan precursors.
In general and according to the examples shown here, the efficiency in photocatalysis using photocatalyst semiconductor oxides obtained by a solid method is higher than that exhibited by the same photocatalysts obtained by solution methods. This may be due to the high porosity of the catalysts obtained by the solid state method.

3.9. Heterojunction Structure

These structures are formed by coupling two semiconductors whose band structures are aligned in such a way that they allow for an enhanced charge separation compared to the single system. A maximum of two alignments are feasible with this binary combination, viz., type I and type II [79,80,81,82]. Figure 5 shows a schematic of the band alignment of the nanocomposites for both heterostructures. In the type I alignment, both the VB and CB edge potential of either semiconductor lies within the bandgap of the other semiconductor, as the latter possesses a wider bandgap. Since one has lower VB and CB potentials than the other, both bands act as hole and electron collecting sites, as is shown in Figure 5. In the type II heterostructure, the CB position of the first semiconductor lies above that of the second, whereas its VB lies within the bandgap of the second. Therefore, the photogenerated holes tend to migrate to the VB of the first one, whereas the excited electrons tend to migrate to the CB of the second. The mentioned band position and migration of the charge carriers are depicted in Figure 5b. In general, the type II heterostructure is widely preferred because it allows the migration of electrons and holes in the opposite direction (see Figure 5b) [79,80,81,82].
Some examples of both type I and type II band alignments are presented and discussed in reference [81].

3.10. Photocatalyst Mechanism of Dye Degradation by Nanostructures

In general, the system of heterogeneous photocatalysis using semiconductor materials consists of a light-harvesting antenna and several active species to facilitate pollutant degradation [83]. The series of chain oxidative-reductive reactions that occur at the photon-activated surface has been broadly proposed as:
Photocatalyst + hν → h+ + e
h+ + H2O → *OH + H+
h+ + OH → *OH
h+ + Pollutant → (Pollutant)+
e + O2 → *O2
*O2 + H+ → *OOH
2 *OOH → O2 + H2O2
H2O2 + *O2 → *OH + OH + O2
H2O2 + hν → 2 *OH
Hence, the final reaction is:
Pollutant + (OH, h+, OOH or O2) → degradation products
which produces the degradation of the pollutant.
When the semiconductor is irradiated by an input light possessing ultra-band-gap energy (hν > Eg), a valence band (VB) electron (e) is excited to the conduction band (CB), leaving behind a photogenerated hole (h+) at the VB. Accordingly, the produced e/h+ pairs are able to migrate to the surface of the semiconductor and participate in redox reactions.
The photocatalytic reaction usually involves three main active species: a hydroxyl radical (OH), h+, and a superoxide radical (O2), where OH is the primary oxidant in the photocatalytic degradation of the pollutant in the aqueous solution. OH radicals are normally generated via two routes: (i) H2O and OH in a water environment are readily oxidized by photogenerated h+ to form OH radicals; (ii) O2 present in an aqueous solution is reduced by photogenerated e to form O2 radicals, which subsequently react with h+ forming OOH radicals, whose further decomposition produces OH radicals.
Complementary to what was discussed above, two interesting reviews [84,85] analyze the photocatalytic properties of the TiO2 system incorporated in a carbon gel matrix and covalent organic frameworks (COFs). In Dongge et al., although still in its incipient stage in comparison with other traditional inorganic metal oxides, titania is proposed as having some advantages such as having a high surface due to the high porosity of the carbon gel matrix. In the second case, Yuhang et al. point out that even when COFs are highly porous and crystalline polymeric materials, they are photocatalytic systems different from nanostructured metal oxide semiconductors, and are a valid alternative in environmental remediation. Meanwhile, recent investigations [86,87] have proven that photocatalysis of nanostructured metal oxides is being increasingly adopted.

4. Conclusions

Nanostructured metal oxides play an important role in the environmental decontamination of organic dyes. An alternative route that often improves photocatalytic efficiency is the use of photocatalysts based on metal oxides obtained by a solid-state method. Specifically, the solid-state method based on the thermal treatment of Chitosan·MXn and PS-co-4-PVP MXn precursors affords pure nanostructured metal oxides MxOy, which are often more efficient photocatalysts than the respective MxOy obtained by other solution methods. It was found that the photocatalytic efficiency depends on the nature of the MXn precursor salts and the calcination temperature, as in the case of TiO2, where the most efficient photocatalyst was the one obtained from the Chitosan TiSO4 precursor pyrolyzed at 800 °C. The efficiency of most of the studied photocatalysts can be described by the following relationship:
PE = a band gap + b particle morphology + c particle size + d crystalline phase + e pyrolysis temperature
where PE is the photocatalyst efficiency. In particular, for a nanostructured metal oxide semiconductor photocatalyst obtained using our solid-state method, morphology is more important than the bandgap. In turn, when using the solid-state preparation method of the metal oxide photocatalyst, the temperature often modulates the morphology. This can, for example, generate a highly porous morphology, thus causing a very efficient photocatalytic activity.
However, research on this topic is a challenge since the parameters that induce a more efficient photocatalysis in solid-state synthesis are not fully known. An achievement in this matter could be a significant advance for environmental decontamination.

Author Contributions

C.D., M.L.V. and M.S. conceptualized, wrote, and agreed to the published version of the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

University of Chile Department of Chemistry, Faculty of Sciences.

Data Availability Statement

This study did not report any data.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hornyak, G.; Tibbals, H.F.; Dutta, J.; Moore, J. Introduction to Nanoscience and Nanotechnology; CRC Press: Boca Raton, FL, USA, 2008. [Google Scholar]
  2. Altavilla, C.; Ciliberto, E. Inorganic Nanoparticles; CRC Press: Boca Raton, FL, USA, 2011. [Google Scholar]
  3. Khin, M.A.; Sreekumaran Nair, A.; Jagadeesh Babu, V.; Murugan, R.; Ramakrishna, S. A review on nanomaterials for environmental remediation. Energy Environ. Sci. 2012, 5, 8075. [Google Scholar] [CrossRef]
  4. Edelstein, A.S.; Cammarata, R.C. Nanomaterials, Synthesis and Applications; Institute of Physics Publishing Bristol: Arrowsmith, UK, 2002. [Google Scholar]
  5. Cao, G. Nanostructures and Nanomaterial, Synthesis, Properties and Applications; Imperial College Press: London, UK, 2004. [Google Scholar]
  6. Ahmada, R.; Ahmadb, Z.; Ullah Khanb, A.; Riaz Mastoi, N.; Aslama, M.; Kim, J. Photocatalytic systems as an advanced environmental remediation: Recent developments, limitations and new avenues for applications. J. Environ. Chem. Eng. 2016, 4, 4143–4164. [Google Scholar] [CrossRef]
  7. Adeyemo, A.A.; Adeoye, I.O.; Bello, O.S. Metal organic frameworks as adsorbents for dye adsorption, overview, prospects and future challenges. Toxicol. Environ. Chem. 2012, 94, 1846–1863. [Google Scholar] [CrossRef]
  8. Chong, M.N.; Jin, B.; Chow, C.W.; Saint, C. Recent developments in photocatalytic water treatment technology: A review. Water Res. 2010, 44, 2997–3027. [Google Scholar] [CrossRef] [PubMed]
  9. Gunten, V. Oxidation Processes in Water Treatment: Are We on Track? Environ. Sci. Technol. 2018, 52, 5062–5075. [Google Scholar] [CrossRef]
  10. Nikam, A.V.; Prasad, B.L.V.; Kulkarnia, A.A. Wet Chemical Synthesis of Metal Oxide Nanoparticles, A Review. Cryst. Eng. Comm. 2018, 20, 5091–5107. [Google Scholar] [CrossRef]
  11. Oska, G. Metal oxide nanoparticles, synthesis, characterization and application. J. Sol. Gel. Sci. Techn. 2006, 37, 161–164. [Google Scholar] [CrossRef]
  12. Diaz, C.; Valenzuela, M.L.; Laguna-Bercero, M.A. Solid-State Preparation of Metal and Metal Oxides Nanostructures and Their Application in Environmental Remediation. Int. J. Mol. Sci. 2022, 23, 1093. [Google Scholar] [CrossRef]
  13. Wang, C.; Li, J.; Liang, X.; Zhang, Y.; Guo, G. Photocatalytic Organic Pollutants Degradation in Metal–Organic Frameworks. Energy Environ. Sci. 2014, 7, 2831–2867. [Google Scholar] [CrossRef]
  14. Liu, X.; Iocozzia, J.; Wang, Y.; Cui, X.; Chen, Y.; Zhao, S.; Li, Z.; Lin, Z. Noble metal–metal oxide nanohybrids with tailored nanostructures for efficient solar energy conversion, photocatalysis and environmental remediation. Energy Environ. Sci. 2017, 10, 402–434. [Google Scholar] [CrossRef]
  15. Gaya, U.I.; Abdullah, A.H. Heterogeneous photocatalytic degradation of organic contaminants over titanium dioxide: A review of fundamentals, progress and problems. J. Photochem. Photobiol. C 2008, 9, 1–12. [Google Scholar] [CrossRef]
  16. Khan, M.M.; Adil, S.F.; Al-Mayouf, A. Metal oxides as photocatalysts. J. Saudi Chem. Soc. 2015, 19, 462–464. [Google Scholar] [CrossRef] [Green Version]
  17. Preeti, S.; Abdullah, M.M.; Saiga, I. Role of Nanomaterials and Their Applications as Photo-catalst and Sensor, A review. Nano Res. Appl. 2016, 2, 1–10. [Google Scholar]
  18. Lany, S. Semiconducting transition metal oxides. J. Phys. Condens. Matter 2015, 27, 283203. [Google Scholar] [CrossRef] [PubMed]
  19. Wang, J.; Wang, Y.; Huang, Y.; Peijnenburg, J.G.; Chen, J.; Li, X. Development of a nano-QSPR model to predict band gaps of spherical metal oxide nanoparticles. RSC Adv. 2019, 9, 8426–8434. [Google Scholar] [CrossRef] [Green Version]
  20. Tran, F.; Blaha, P. Accurate Band Gaps of Semiconductors and Insulators with a Semilocal Exchange-Correlation Potential. Phys. Rev. Lett. 2009, 102, 226401. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  21. Díaz, C.; Valenzuela, M.L.; Laguna-Bercero, M.A.; Mendoza, K.; Cartes, P. Solventless preparation of thoria, their inclusion inside SiO2 and TiO2, their luminiscent properties and their photocataltytic behavior. ACS Omega 2021, 6, 9391–9400. [Google Scholar] [CrossRef]
  22. Guo, Y.; Ma, L.; Mao, K.; Ju, M.; Bai, Y.; Zhao, J.; Zeng, X. Eighteen functional monolayer metal oxides: Wide bandgap semiconductors with superior oxidation resistance and ultrahigh carrier mobility. Nanoscale Horiz. 2019, 4, 592–600. [Google Scholar] [CrossRef] [Green Version]
  23. Allende, P.; Laguna, M.A.; Barrientos, L.; Valenzuela, M.L.; Díaz, C. Solid state tuning Morphology, Crystal Phase and Size through Metal Macromolecular Complexes and Its Significance in the Photocatalytic Response. ACS Appl. Energy Mater. 2018, 1, 3159–3170. [Google Scholar] [CrossRef] [Green Version]
  24. Ikram, M.; Rashid, M.; Haider, A.; Naz, S.; Haider, J.; Raza, A.; Ansard, M.T.; Uddin, K.; Nageh, M.; Ali, M.; et al. A review of photocatalytic characterization, and environmental cleaning, of metal oxide nanostructured materials. Sustain. Mater. Technol. 2021, 30, e00343. [Google Scholar] [CrossRef]
  25. Ferreira, D.L.; Sousa, J.C.L.; Maronesi, R.N.; Bettini, J.; Schiavon, M.A.; Silva, A.G. Size-dependent bandgap and particle size distribution of colloidal semiconductor nanocrystals. J. Chem. Phys. 2017, 147, 154102–154111. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Ahmad, M.; Zhu, J. ZnO based advanced functional nanostructures: Synthesis, properties and applications. J. Mater. Chem. 2011, 21, 599–614. [Google Scholar] [CrossRef]
  27. Wang, Z.L. Zinc oxide nanostructures: Growth, properties and applications. J. Phys. Condens. Matter 2004, 16, R829–R858. [Google Scholar] [CrossRef]
  28. Chen, X.; Mao, S.S. Titanium Dioxide Nanomaterials, Synthesis, Properties, Modifications, and Applications. Chem Rev. 2007, 107, 2891–2959. [Google Scholar] [CrossRef] [PubMed]
  29. Ismail, A.A.; Bahnemannb, D.W. Mesoporous titania photocatalysts, preparation, characterization and reaction mechanisms. J. Mater. Chem. 2011, 21, 11686–11707. [Google Scholar] [CrossRef] [Green Version]
  30. He, H.; Liu, C.; Dubois, K.; Jin, T.; Louis, M.E.; Li, G. Enhanced Charge Separation in Nanostructured TiO2 Materials for Photocatalytic and Photovoltaic Applications. Ind. Eng. Chem. Res. 2012, 51, 11841–11849. [Google Scholar] [CrossRef]
  31. Ge, M.; Cao, C.; Huang, J.; Li, S.; Chen, Z.; Zhang, K.; Al-Deyab, S.S.; Lai, Y. A Review of One-dimensional TiO2 Nanostructured Materials for Environmental and Energy Applications. J. Mater. Chem. A 2016, 4, 6772–6801. [Google Scholar] [CrossRef]
  32. Díaz, C.; Valenzuela, M.L. Metallic Nanostructures Using Oligo and Polyphosphazenes as Template or Stabilizer in Solid State in Encyclopedia of Nanoscience and Nanotechnology; Nalwa, H.S., Ed.; American Scientific Publishers: Valencia, CA, USA, 2010; Volume 16, pp. 239–256. [Google Scholar]
  33. Díaz, C.; Valenzuela, M.L. Macromolecular Complexes MXn Polymer as solid state precursors of metal and metal oxides nanostructures. In Book of CRC Concise Encyclopedia of Nanotechnology; Kharisov, B.I., Kharisov, O.V., Mendez, U., Eds.; CRC Press: Boca Raton, FL, USA, 2016; Chapter 42; pp. 504–524. [Google Scholar]
  34. Ray, C.; Pai, T. Recent advances of metal–metal oxide nanocomposites and their tailored nanostructures in numerous catalytic applications. J. Mater. Chem. 2017, 5, 9465–9478. [Google Scholar] [CrossRef]
  35. Díaz, C.; Valenzuela, M.L. Small-Molecule and High-Polymeric Phosphazenes containing oxypyridine side groups and their organometallic derivatives; Useful precursors for metal nanostructured materials. Macromolecules 2006, 39, 103–111. [Google Scholar] [CrossRef]
  36. Díaz, C.; Valenzuela, M.L. Organometallic Derivatives of Polyphosphazenes as Precursors for Metallic Nanostructured Materials. J. Inorg. Organomet. Polym. Mater. 2006, 16, 419–435. [Google Scholar] [CrossRef]
  37. Díaz, C.; Valenzuela, M.L.; Zuñiga, L.; O’Dwyer, C. Organometallic derivatives of cyclotriphosphazene as precursors of Nanostructured metallic materials: A new solid state Method. J. Inorg. Organomet. Polym. Mater. 2009, 19, 507–520. [Google Scholar] [CrossRef] [Green Version]
  38. Díaz, C.; Valenzuela, M.L.; Lavayen, V.; O’Dwyer, C. Layered Graphitic Carbon Host Formation during Liquid-free Solid State Growth of Metal Pyrophosphates. Inorganic Chem. 2012, 51, 6228–6236. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  39. Díaz, C.; Valenzuela, M.L.; Cáceres, S.; O´Dwyer, C. Solution and surfactant-free growth of supported high index facet SERS active nanoparticles of rhenium by phase demixing. J. Mater. Chem. A 2013, 1, 1566–1572. [Google Scholar]
  40. Díaz, C.; Valenzuela, M.L.; Cáceres, S.; O´Dwyer, C.; Diaz, R. Solvent and stabilizer free growth of Ag and Pd nanoparticles using Metallic salts/cyclotriphosphazenes mixtures. Mater. Chem. Phys. 2013, 143, 124–132. [Google Scholar] [CrossRef]
  41. Díaz, C.; Valenzuela, M.L.; Zuñiga, L.; O’Dwyer, C. Solid State Pathways to Complex Shape Evolution and Tunable Porosity During Metallic Crystal Growth. Sci. Rep. 2013, 2642, 1–8. [Google Scholar]
  42. Díaz, C.; Valenzuela, M.L. Nanostructures: Properties: Production Methods and Application. In Organometallic-Metallic-Cyclotriphosphazene Mixtures: Solid-State Method for Metallic Nanoparticle Growth; Nova Science Publishers: New York, NY, USA, 2013; Chapter 5. [Google Scholar]
  43. Díaz, C.; Valenzuela, M.L. A General Solid-State Approach to Metallic, Metal Oxides and Phosphates Nanoparticles in Advances in Chemical Research; Nova Science Publishers: New York, NY, USA, 2011. [Google Scholar]
  44. Díaz, C.; Valenzuela, M.L. A general Solid-State approach to Metallic, Metal oxides and Phosphate nanoparticles Gold Nanoparticles: Properties Synthesis and Fabrication. In Solution and Solid State Methods to Prepare Au Nanoparticles: A comparison Chow PE; Nova Science Publishers: New York, NY, USA, 2010; Chapter 14. [Google Scholar]
  45. Díaz, C.; Valenzuela, M.L.; Lavayen, V.; Mendoza, K.; Peña, O.; O’Dwyer, C. Nanostructured copper oxides and phosphates from a new solid-state route. Inorg. Chim. Acta 2011, 377, 5–11. [Google Scholar] [CrossRef]
  46. Díaz, C.; Platoni, S.; Molina, A.; Valenzuela, M.L.; Geaney, H.; O’Dwyer, C. Novel Solid-State Route to Nanostructured Tin, Zinc and Cerium Oxides as Potential Materials for Sensors. J. Nanosci. Nanotechnol 2014, 14, 7648–7675. [Google Scholar] [CrossRef]
  47. Díaz, C.; Barrera, G.; Segovia, M.; Valenzuela, M.L.; Osiak, M.; O’Dwyer, C. Solvent-less method for efficient photocatalytic α-Fe2O3 nanoparticles for using macromolecular polymeric precursors. New J. Chem. 2016, 40, 6768–6776. [Google Scholar] [CrossRef]
  48. Díaz, C.; Valenzuela, M.L.; Laguna, M.A.; Orera, A.; Bobadilla, D.; Abarca, S.; Peña, O. Synthesis and Magnetic Properties of Nanostructured metallic Co, Mn and Ni oxide materials obtained from solid-state macromolecular complex precursors. RSC Adv. 2017, 7, 27729–27736. [Google Scholar] [CrossRef] [Green Version]
  49. Díaz, C.; Valenzuela, M.L.; Bobadilla, D.; Laguna-Bercero, M.A. Bimetallic Au//Ag Alloys Inside SiO2 using a solid-state method. J. Clust. Chem. 2017, 28, 2809–2815. [Google Scholar] [CrossRef] [Green Version]
  50. Díaz, C.; Valenzuela, M.L.; Garcia, C.; De la Campa, R.; Soto, A.-P. Solid-state synthesis of pure and doped lanthanides oxide nanomaterials by using polymer templates. Study of their luminescent properties. Mater. Lett. 2017, 209, 111–114. [Google Scholar] [CrossRef]
  51. Díaz, C.; Valenzuela, M.L.; Segovia, M.; De la Campa, R.; Soto, A.-P. Solution, Solid-State Two Step Synthesis and Optical Properties of ZnO and SnO Nanoparticles and Their Nanocomposites with SiO2. J. Clust. Sci. 2018, 29, 251–266. [Google Scholar] [CrossRef] [Green Version]
  52. Díaz, C.; Carrillo, D.; De la Campa, R.; Soto, A.-P.; Valenzuela, M.L. Solid-State synthesis of LnOCl/Ln2O3 (Ln = Eu, Nd) by using chitosan and PS-co-P4VP as polymeric supports. J. Rare Earth 2018, 36, 1326–1332. [Google Scholar] [CrossRef]
  53. Allende, P.; Barrientos, L.; Orera, A.; Laguna-Bercero, M.A.; Salazar, N.; Valenzuela, M.L.; Diaz, C. TiO2/SiO2 Composite for Efficient Protection of UVA and UVB Rays Through of a Solvent-Less Synthesis. J. Clust. Sci. 2019, 30, 1511–1517. [Google Scholar] [CrossRef]
  54. Díaz, C.; Valenzuela, M.L.; Cifuentes-Vaca, O.; Segovia, M.; Laguna-Bercero, M.A. Incorporation of Nanostructured ReO3 in Silica Matrix and Their Activity Toward Photodegradation of Blue Methylene. J. Inorg. Organomet. Polym. Mater. 2020, 30, 1726–1734. [Google Scholar] [CrossRef]
  55. Diaz, C.; Valenzuela, M.L.; Cifuentes-Vaca, O.; Segovia, M.; Laguna-Bercero, M.A. Iridium nanostructured metal oxide, its inclusion in Silica matrix and their activity toward Photodegradation of Methylene Blue. Mater. Chem. Phys. 2020, 252, 123276–123286. [Google Scholar] [CrossRef]
  56. Díaz, C.; Valenzuela, M.L.; Cifuentes-Vaca, O.; Segovia, M. Polymer precursors effect in the macromolecular metal-polymer on the Rh/RhO2/Rh2O3 phase using solvent-less synthesis and its photocatalytic activity. J. Inorg. Organomet. Polym. Mater. 2020, 30, 4702–4708. [Google Scholar] [CrossRef]
  57. Díaz, C.; Valenzuela, M.L.; Cifuentes-Vaca, O.; Segovia, M.; Laguna-Bercero, M.A. Incorporation of NiO into SiO2, TiO2, Al2O3, and Na4.2Ca2.8(Si6O18) Matrices, Medium Effect on the Optical Properties and Catalytic Degradation of Methylene Blue. Nanomaterials 2020, 10, 2470. [Google Scholar] [CrossRef]
  58. Sietsma, J.R.A.; Meeldijk, J.D.; Den Breejen, J.P.; Versluijs-Helder, M.; Jos van Dillen, A.; De Jongh, P.E.; De Jong, K.P. The Preparation of Supported NiO and Co3O4 Nanoparticles by the Nitric Oxide Controlled Thermal Decomposition of Nitrates. Ang. Chem Int. Ed. 2007, 46, 4547–4549. [Google Scholar] [CrossRef]
  59. Zhang, Z.; Shao, C.; Li, X.; Wang, C.; Zhang, M.; Liu, Y. Electrospun Nanofibers of p~Type NiO/n~Type ZnO Heterojunctions with Enhanced Photocatalytyc Activity. ACS. Appl. Mater. Interface Sci. 2010, 10, 2915–2923. [Google Scholar] [CrossRef]
  60. Yu, J.; Wang, W.; Cheng, B. Synthesis and Enhanced Photocatalytic Activity of a Hierarchical Porous Flowerlike p–n Junction NiO/TiO2 Photocatalyst. Chem. Asian 2010, 5, 2499–2506. [Google Scholar] [CrossRef] [PubMed]
  61. Bonomo, M.; Dini, D.; Decker, F. Electrochemical and Photoelectrochemical Properties of Nickel Oxide (NiO) With Nanostructured Morphology for Photoconversion Applications. Front. Chem. 2018, 6, 601. [Google Scholar] [CrossRef] [PubMed]
  62. Kelly, K.L.; Coronado, E.; Zhao, L.L.; Schatz, G. The optical properties of metal nanoparticles: The influence of size, shape, and dielectric environment. J. Phys. Chem. 2003, 107, 668–677. [Google Scholar] [CrossRef]
  63. Cotton, F.A.; Wilkinson, G. Advanced Inorganic Chemistry; John Wiley and Sons: New York, NY, USA, 1980; Chapters 22 and 30. [Google Scholar]
  64. Jin, R. The impacts of nanotechnology on catalysis by precious metal nanoparticles. Nanotechnol. Rev. 2012, 31–56. [Google Scholar] [CrossRef]
  65. Fernandez-Garcia, M.; Martinez-Arias, A.; Hanson, J.; Rodriguez, C. Nanostructured Oxides in Chemistry: Characterization and Properties. Chem. Rev. 2004, 104, 4063–4104. [Google Scholar] [CrossRef]
  66. Chen, R.S.; Korotcov, A.; Huiang, Y.S.; Tsai, D. One-dimensional conductive IrO2. Nanocrystals Nanotechnol. 2006, 17, 67–87. [Google Scholar] [CrossRef] [Green Version]
  67. Xu, D.; Diao, P.; Jin, T.; Wu, Q.; Liu, X.; Guo, X.; Gong, H.; Li, F.; Xiang, M.; Ronghai, Y. Iridium oxide nanoparticles and Iridium/Iridium Oxide Nanocompósites: Photochemical Fabrication and Application in Catalytic Rediction of Notrophenol. ACS Appl. Interfaces 2015, 7, 16738–16749. [Google Scholar] [CrossRef]
  68. Zhao, Y.; Hernandez, E.A.; Vargas-Barbosa, N.M.; Dysart, J.L.; Mallouk, T.E. A high yield Synthesis of ligand-free iridium oxide nanoparticles with high electrocatalytic activity. J. Phys. Chem. Lett. 2011, 2, 402–406. [Google Scholar] [CrossRef]
  69. Biswas, K.; Rao, C.N. Metallic ReO3 Nanoparticles. J. Phys. Chem. B 2006, 110, 842–845. [Google Scholar] [CrossRef]
  70. Biswas, K.; Bhat, S.V.; Rao, C.N. Surface-Enhanced Raman Spectra of Aza-aromatics on Nanocrystals of Metallic ReO3. J. Phys. Chem. C 2006, 111, 5689–5693. [Google Scholar] [CrossRef]
  71. Biswas, K.; Rao, C.N. Synthesis and Characterization of Nanocrystals of the Oxide Metals, RuO2, IrO2, and ReO3. J. Nanosci. Nanotechniol 2007, 7, 1969–1974. [Google Scholar] [CrossRef] [PubMed]
  72. Lin, Z.-W.; Kuang, Q.; Lian, W.; Jiang, Z.-Y.; Xie, Z.-X.; Huang, R.-B.; Zheng, L.-S. Preparation and Optical Properties of ThO2 and Eu-Doped ThO2 Nanotubes by the Sol−Gel Method Combined with Porous Anodic Aluminum Oxide Template. J. Phys. Chem. B 2006, 110, 23007–23011. [Google Scholar] [CrossRef] [PubMed]
  73. Hudry, D.; Apostolidis, C.; Walter, O.; Gouder, T.; Courtois, E.; Kubel, C.; Meyer, D. Non-aqueous Synthesis of Isotropic and Anisotropic Actinide Oxide Nanocrystals. Chem. Eur. J. 2012, 18, 8283–8287. [Google Scholar] [CrossRef] [PubMed]
  74. Hudry, D.; Apostolidis, C.; Walter, O.; Gouder, H.; Courtois, E.; Kubel, C.; Meyer, D. Controlled Synthesis of Thorium and Uranium Oxide Nanocrystals. Chem. Eur. J. 2013, 19, 5297–5305. [Google Scholar] [CrossRef]
  75. Tripathi, V.K.; Narajan, R. Sol–Gel Synthesis of High-Purity Actinide Oxide ThO2 and Its Solid Solutions with Technologically Important Tin and Zinc Ions. Inorg. Chem. 2016, 55, 12798–12906. [Google Scholar] [CrossRef]
  76. Pereira, F.J.; Castro, M.A.; Vazquez, M.D.; Deban, L.; Aller, A.J. Optical properties of ThO2–based nanoparticles. J. Lumin. 2017, 184, 169–178. [Google Scholar] [CrossRef]
  77. Huentupil, Y.; Cabello-Guzman, G.; Chornik, B.; Arancibia, R.; Buono-Cuore, G.E. Photochemical deposition, characterization and optical properties of thin films of ThO2. Polyhedron 2019, 157, 225–231. [Google Scholar] [CrossRef]
  78. Mahmoud, S.A. Characterization of thorium dioxide thin films prepared by the spray pyrolysis technique. Solid State Sci. 2002, 4, 221–228. [Google Scholar] [CrossRef]
  79. Li, H.; Zhou, Y.; Tu, W.; Ye, J.; Zou, Z. State-of-the-art progress in diverse heterostructured photocatalysts toward promoting photocatalytic performance. Adv. Funct. Mater. 2014, 25, 998–1013. [Google Scholar] [CrossRef]
  80. Wang, Y.; Wang, Q.; Zhan, X.; Wang, F.; Safdar, M.; He, J. Visible light driven type II heterostructures and their enhanced photocatalysis properties: A review. Nanoscale 2013, 5, 8326–8339. [Google Scholar] [CrossRef]
  81. Rani, A.; Reddy, R.; Sharma, U.; Mukherjee, P.; Mishra, P.; Kuila, A.; Sim, L.C.; Saravanan, P. A review on the progress of nanostructure materials for energy harnessing and environmental remediation. J. Nanostructure Chem. 2018, 8, 255–291. [Google Scholar] [CrossRef] [Green Version]
  82. Li, H.; Tu, W.; Zhou, Y.; Zou, Z. Z-Scheme Photocatalytic Systems for Promoting Photocatalytic Performance: Recent Progress and Future Challenge. Adv. Sci. 2016, 3, 1500389. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  83. Dong, S.; Feng, J.; Fan, M.; Pi, Y.; Hu, L.; Han, X.; Liu, M.; Sun, J.; Sun, J. Recent developments in heterogeneous photocatalytic water treatment using visible light responsive photocatalysts: A review. RSC Adv. 2015, 5, 14610–14630. [Google Scholar] [CrossRef]
  84. Dongge, M.; Jundan, L.; Anan, L.; Chuncheng, C. Carbon Gels-Modified TiO2: Promising Materials for Photocatalysis Applications. Materials 2020, 13, 1734. [Google Scholar]
  85. Yuhang, Q.; Dongge, M. Covalent Organic Frameworks: New Materials Platform for Photocatalytic Degradation of Aqueous Pollutants. Materials 2021, 14, 5600. [Google Scholar]
  86. Nunes, D.; Pimentel, A.; Branquinho, R.; Fortunato, E.; Martins, R. Metal Oxide-Based Photocatalytic Paper: A Green Alternative for Environmental Remediation. Catalysts 2021, 11, 504. [Google Scholar] [CrossRef]
  87. Shah, M.; Estrella, L.; Alemaida, I.; Lisin, A.; Moiseev, N.; Ahmadi, M.; Nazari, M.; Wali, M.; Zaheb, H.; Senjyu, T. Photocatalytic Applications of Metal Oxides for Sustainable Environmental Remediation. Metals 2021, 11, 80. [Google Scholar]
Figure 1. Photocatalytic activity of nanostructured metal oxides.
Figure 1. Photocatalytic activity of nanostructured metal oxides.
Photochem 02 00041 g001
Figure 2. Schematic representation of the solid-state method and its application in environmental remediation.
Figure 2. Schematic representation of the solid-state method and its application in environmental remediation.
Photochem 02 00041 g002
Figure 3. FE-SEM images of TiO2-Anatase-(III)p(800) (a) and an amplification of a portion of the structure (b) (Adapted from reference [23]).
Figure 3. FE-SEM images of TiO2-Anatase-(III)p(800) (a) and an amplification of a portion of the structure (b) (Adapted from reference [23]).
Photochem 02 00041 g003
Figure 4. Schematic diagrams for (a) energy bands of p-NiO and (b) TiO2 before contact, (c) formation of p-n junction and its energy diagram at equilibrium, and (d) transfer of holes from n-TiO2 to p-NiO under UV irradiation (Adapted from reference [57]).
Figure 4. Schematic diagrams for (a) energy bands of p-NiO and (b) TiO2 before contact, (c) formation of p-n junction and its energy diagram at equilibrium, and (d) transfer of holes from n-TiO2 to p-NiO under UV irradiation (Adapted from reference [57]).
Photochem 02 00041 g004
Figure 5. Schematic illustration of the type-I (a) and Type II (b) band alignments.
Figure 5. Schematic illustration of the type-I (a) and Type II (b) band alignments.
Photochem 02 00041 g005
Table 1. Bandgap values for the main metal oxides along the periodic table.
Table 1. Bandgap values for the main metal oxides along the periodic table.
Metal OxidesCompoundsBandgap (eV) a,bRef.
Metal transitionCr2O33.2[18]
V2O52.3[18]
Co3O41.5[18]
TiO23.3 (anatase)[18]
TiO23.0 (rutile)[18]
Mn2O33.27[19]
WO32.8[19]
MoO33.14[18]
NiO3.5[19]
Fe2O32.1[18]
Fe3O42.61[19]
Cu2O2.2[18]
CuO1.6[18]
Metal representativeZnO3.2,34[19,20]
SnO24.2[19]
Ga2O34.85[19]
Sb2O34.49[19]
LanthanidesCeO23.0–3.6[19]
La2O34.3[19]
ActinidesThO23.1[21]
a Values from references: [17,18,19,20,21,22]. b Other similar values are reported for some of the metal oxides.
Table 2. Photocatalytic Performances and Key Factors (Crystal Phase, Size and Morphology) of Different TiO2 Materials in the Degradation of Organic Water Pollutants [24].
Table 2. Photocatalytic Performances and Key Factors (Crystal Phase, Size and Morphology) of Different TiO2 Materials in the Degradation of Organic Water Pollutants [24].
Organic Water PollutantPhotocatalyst
(Concentration)
Irradiation LightReaction KineticRemoval %Irradiation Time
MB (1 × 10−5 M)TiO2 anatase. 1 g × Lt−1 with 340 nm cut-off filter, 330 nm > λ > 680 nmXe lamp (150 W)Pseudo first-order kinetic86.5%25 min
MB (10 ppm)TiO2 nanoparticles (2.5 g L−1)UV lamp (40 W)Langmuir–Hinshelwood Pseudo first-order kinetic71%60 min
MB (50 ppm)TiO2 nanofibersXenon lamp UV-vis (150 W) with AM
1.5 G filter λ > 400 nm
NM100%180 min
MB(10 mgL−1)TiO2 nanotubes (0.16 g L−1)xenon visible light (500 W)NM99.1%40 min
MB (0.01 mM)mesoporous TiO2 0.17 g L−1)UV light irradiation using eight tubes with a power source of 6 W, λ = 365 nmNM85%60 min
MB
(0.75 × 10−5 M)
Hollow titania micro-spheres (HTS)UV lamp 15 WPseudo first-order kinetic53%90 min
MB (1 mM)TiO2 nanoparticlesBlacklight lamp (1 mW)NM45%20 min
Other similar organic water pollutants malachite green, MG (10 ppm)TiO2nanoparticles (2.5 g L−1)UV lamp (40 W)Langmuir–Hinshelwood Pseudo first-order kinetic78%60 min
Rhodamine B, RhB, and methyl orange MO (0.01 mM)TiO2 hollow tetragonal nanocone (0.1 g L−1)full-arc Xe lamp (300 W) with a cutoff filter, λ > 420 nmNM95.0% for RhB, 90.7% for MO30 min
Bisphenol A BPA, (200 μM)TiO2 powder, Degussa P25 (0.5 g L−1)Xe arc lamp (300 W), IR water filter and cutoff filter, λ > 420 nmPseudo first-order kinetic75% for BPA4 h
acetaminophen, Ace (1.3 μM)TiO2 powder, Degussa P25 (NM)UVA/LED
λmax = 366 nm
NM100%8 min
RhB (NM)TiO2 nanopowder (0.3 g L−1)visible light
λ > 420 nm
NM90%5 h
Methyl orange 4 × 10−5 M TiO2 powder
(10 mg)
Xe lamp 300 WNM90%3 h
Table 3. Photocatalytic improvement parameter of various ZnO nano-photocatalysts (Adapted from reference [24]).
Table 3. Photocatalytic improvement parameter of various ZnO nano-photocatalysts (Adapted from reference [24]).
PhotocatalystBandgapSurface Area
Bi3+-ZnO3.15–2.6Increase
Co-ZnO3.34–3.06Increase
N-ZnO3.15–2.86Increase
F-ZnO3.35–2.51Increase
Fe-ZnO3.24–3.16Increase
Ag/ZnO3.30–3.21Increase
B/ZnO3.2–3.1Increase
Bi-TiO2
Ni-TiO2
2.99–3.08
3.02, 2.99–3.03
Decrease
Ag-TiO23.0–2.6Increase
Fe-TiO23.2–2.98Increase
N-TiO23.1–2.7Increase
Ti/WO33.4–3.31Decrease
Zn-WO33.2–3.12Slight decrease
Fe-SnO23.8–1.65Increase
Zn-SnO23.50–3.17Increase
Cu-SnO23.02–2.2Increase
Table 4. Photocatalytic improvement performance of various ZnO nano-photocatalysts (Adapted from reference [24]).
Table 4. Photocatalytic improvement performance of various ZnO nano-photocatalysts (Adapted from reference [24]).
CompositeSynthesis MethodPollutant for DegradationPC PerformanceIrradiation
1Co-ZnOSol-gelMB3 at.% Co-ZnO exhibited 92% degradation in 60 min.Visible light
2La-doped ZnOHydrothermal oxidationMO and MBBest PCA observed by S0.005 due to defects.UV light
3Pd/ZnOMicrowave hydrothermal, borohydride and photoreductionCRPd/ZnO synthesized by borohydride method has the highest PCA compared to other routes.UV light
4C, N co-doped ZnOTwo-step pyrolysisMB6C25 showed the best degradation due to its larger number of active sites.Solar stimulated light
5X-ZnO (X = Li, Al, N, P)Mass production technologyRhBPCA decreased as follows: N > Li > P > AlSunlight
6Fe-doped ZnOCombustionBPAFexZn1−xO (where x = 0.03) showed noticeable efficacy in the series.Sunlight
7Gd-ZnO filmsRF magnetron sputteringMB0.7 at.% Gd-ZnO exhibited higher PCA than ZnO.UV light
8Au-ZnOCombustionMBAg-ZnO demonstrated better activity than Au-ZnO.UV light
Ag-ZnO
9Eu3+-doped ZnOCoprecipitationRhBDoped ZnO (100%) degraded dye faster than ZnO.UV light
10Ce-ZnOHydrothermalPharmaceuticaNizatidine, levofloxacin and acetaminophen degraded around 95% within 4 h.UV light
11Cu-ZnOChemical growthMO and MBIncrease in PC efficiency was 57.5% for MO and 60% for MB in 180 min.UV light
12B-ZnOSol-gelCNDoped sample with 1.5 wt% exhibited 89% degradation whereas pure ZnO exhibited 75%.Solar stimulated light
13In-ZnOPlasma assisted chemical vaporsMB4 at.% In-ZnO showed improved PCA compared to ZnO and 8 at.% In-ZnO.Solar stimulated light
14Sm-ZnOChemical precipitationMBZn1−xSmxO x = 0.04 expressed maximum PC degradation (94.94%).Visible light
15WO3-doped ZnOhydrothermalDiazinon10 mg/L diazinon, 10 g/cm—2 2% O3-ZnO exhibited 99% degradation in 180 min.UV light
Table 5. Summary of various TiO2 nano-photocatalysts, pollutants, and irradiation sources along with their measured Photocatalysis (Adapted from reference [24]).
Table 5. Summary of various TiO2 nano-photocatalysts, pollutants, and irradiation sources along with their measured Photocatalysis (Adapted from reference [24]).
Doped Pollutants PCAIrradiation Source
1Ce-TiO2
La-TiO2
V-TiO2
RhB1.0%-Ce-TiO2 > 1.0%-V-TiO2 > undoped TiO2 > 1.0%-La-TiO2 showed degradation (%) 83.43 > 53.74 > 21.56 > 11.09, respectively.Solar light
2PF co-doped anatase TiO2MOIt demonstrated excellent PCA compared to undoped TiO2, F-TiO2, P-TiO2 and Degussa P25.Full-spectrum light
3Ga-doped TiO2nanopowderMO0.6 mol% Ga-TiO2 demonstrated up to 82% degradation.Visible light
4F, N co-doped TiO2MB97.31% degradation was achieved within 5 h.Visible light
5Nd-TiO2 filmMB0.1% Nd-doped TiO2 showed maximum degradation (92%).UV light
6Moroccan natural P-TiO2ICIts degradation increased at high values of pH, initial concentration and amount of catalyst.UV light
7Nb2O5-TiO2MB5 mol% expressed the highest PCA under visible light and equal efficiency as TiO2 under UV.UV and Visible light
8Mesoporous Ag-TiO2MOTiO2 with the lowest content of Ag exhibited higher PCA.UV and solar light
9Fe3+-TiO2CVDegradation kinetics rate increased with an increase in iron content.UV
10Ru/TiO22-CP0.4 wt% Ru/TiO2 showed high PCA using UV light and 0.2% Ru/TiO2 using visible light.UV and visible light
11Gd-TiO2RhB0.3% Gd-TiO2 demonstrated the best PCA.Visible light
12C-TiO2RhBMore decolorization compared to fluorescence spectroscopy.Visible light
13Pd-TiO2MB and MOMaximum degradation shown by 0.75 wt% Pd-TiO2 for mixture of dyes and 0.5 wt% Pd-TiO2 for single dye.Visible or solar light
14Graphene/TiO2BPAIt showed remarkable PCA compared to pure TiO2.Solar light
15Ni/TiO2Malathion94% degradation achieved.UV light
Table 6. Bandgap for TiO2 obtained using different precursors and temperatures (Adapted from reference [23]).
Table 6. Bandgap for TiO2 obtained using different precursors and temperatures (Adapted from reference [23]).
TiO2-(I)pTiO2-(II)p
Temperature (°C)PhaseBandgap (eV)Temperature (°C)PhaseBandgap (eV)
500Anatase3.66500Anatase3.43
600Mixture3.43600Anatase3.30
700Mixture3.63700Anatase3.22
800Rutile3.40800Mixture3.27
TiO2-(III)pTiO2-(IV)p
Temperature (°C)PhaseBandgap (eV)Temperature (°C)PhaseBandgap (eV)
500Anatase3.32500Anatase3.53
600Anatase3.43600Anatase3.45
700Anatase3.28700Mixture3.37
800Anatase3.65800Mixture3.38
TiO2-(V)pTiO2-(VI)p
Temperature (°C)PhaseBandgap (eV)Temperature (°C)PhaseBandgap (eV)
500Anatase3.47500Anatase3.72
600Anatase3.42600Mixture3.57
700Mixture3.21700Mixture3.36
800Mixture3.33800Rutile3.24
Table 7. Kinetic data for degradation of methylene blue with nanostructured TiO2 (Adapted from reference [23]).
Table 7. Kinetic data for degradation of methylene blue with nanostructured TiO2 (Adapted from reference [23]).
SampleApparent Photodegradation Rate Constant k (10−2 min−1)Degradation
η (%)
R2 (%)
TiO2-Anatase-(I)p 500 °C0.40 ± 0.041193.9
TiO2-Mixture-(I)p 600 °C0.80 ± 0.042098.4
TiO2-Mixture-(I)p 700 °C0.06 ± 0.031498.0
TiO2-Rutile-(I)p 800 °C1.30 ± 0.103093.8
TiO2-Anatase-(II)p 500 °C0.40 ± 0.041191.9
TiO2-Anatase-(II)p 600 °C0.33 ± 0.03788.8
TiO2-Anatase-(II)p 700 °C0.40 ± 0.021097.6
TiO2-Mixture-(II)p 800 °C1.00 ± 0.062397.2
TiO2-Anatase-(III)p 500 °C3.90 ± 0.206596.9
TiO2-Anatase-(III)p 600 °C4.10 ± 0.206597.4
TiO2-Anatase-(III)p 700 °C2.00 ± 0.063999.2
TiO2-Anatase-(III)p 800 °C7.13 ± 0.018799.8
TiO2-Anatase-(IV)p 500 °C2.40 ± 0.014597.8
TiO2-Anatase-(IV)p 600 °C3.10 ± 0.015598.0
TiO2-Mixture-(IV)p 700 °C0.20 ± 0.02591.1
TiO2-Mixture-(IV)p 800 °C4.00 ± 0.206397.1
TiO2-Anatase-(V)p 500 °C3.40 ± 0.105999.1
TiO2-Anatase-(V)p 600 °C1.30 ± 0.042799.2
TiO2-Mixture-(V)p 700 °C1.30 ± 0.092796.4
TiO2-Mixture-(V)p 800 °C0.60 ± 0.031498.0
TiO2-Anatase-(VI)p 500 °C1.70 ± 0.033499.7
TiO2-Mixture-(VI)p 600 °C2.40 ± 0.024895.9
TiO2-Mixture-(VI)p 700 °C2.60 ± 0.025194.6
TiO2-Rutile-(VI)p 800 °C0.50 ± 0.021498.5
Table 8. Kinetic data for the degradation of MB with α-Fe2O3 obtained from PS-co-4-PVP and chitosan macromolecular precursors (Adapted from reference [47]).
Table 8. Kinetic data for the degradation of MB with α-Fe2O3 obtained from PS-co-4-PVP and chitosan macromolecular precursors (Adapted from reference [47]).
PrecursorEg (eV)
Chitosan(FeCl3) 1:11.83
Chitosan(FeCl3) 1:52.15
PS-co-4-PVP(FeCl3) 1:12.12
Fe+3-PS-co-4-PVP(FeCl3) 1:52.12
Chitosan(FeCl2) 1:12.15
Chitosan(FeCl2) 1:52.15
PS-co-4-PVP(FeCl2) 1:11.90
PS-co-4-PVP(FeCl2) 1:52.09
Table 9. Kinetic data for the degradation of MB with α-Fe2O3 obtained from PS-co-4-PVP (1:1) and chitosan (1:1) macromolecular precursors (Adapted from reference [47]).
Table 9. Kinetic data for the degradation of MB with α-Fe2O3 obtained from PS-co-4-PVP (1:1) and chitosan (1:1) macromolecular precursors (Adapted from reference [47]).
PhotocatalystApparent Photodegradation Rate Constant k (10−2 min−1)Discoloration Rate η(%) at 60 minDiscoloration Rate η(%) at 150 min
α-Fe2O3 (PS-co-4-PVP)1.2 ± 0.0462.686.9
α-Fe2O3 (Chitosan)2.1 ± 0.173.494.6
Table 10. Kinetic data for the photodegradation process of MB with NiO and with the NiO/SiO2, NiO/TiO2, NiO/Al2O3, and NiO/glasses composites (Adapted from reference [57]).
Table 10. Kinetic data for the photodegradation process of MB with NiO and with the NiO/SiO2, NiO/TiO2, NiO/Al2O3, and NiO/glasses composites (Adapted from reference [57]).
PhotocatalystApparent Photodegradation Rate Constant k (10−2 M·min−1)Discoloration Rate (%)R2 Linear Fit (%)
NiO-chitosan a2.471%0.998
NiO-PS-4-PVP2.268%0.991
NiO/SiO2-chitosan2.369%0.999
NiO/SiO2-PS-4-PVP1.648%0.996
NiO/TiO2-chitosan2.991%0.992
NiO/TiO2-PS-4-PVP2.681%0.980
NiO/Al2O3-chitosan1.545%0.990
NiO/Na4.2Ca2.8(Si6O18)2.675%0.990
a Nomenclature: Composite, hyphen (-), name of the precursor polymer.
Table 11. Kinetic data for the photodegradation process of MB with IrO2, Rh/RhO2, Rh/Rh2O3, and ReO3.
Table 11. Kinetic data for the photodegradation process of MB with IrO2, Rh/RhO2, Rh/Rh2O3, and ReO3.
PhotocatalystPhotodegradation Rate Constant k (10−3 M·min−1)Discoloration Rate (%)R2 Linear Fit (%)Ref.
IrO2-PS-4-PVP1.753%0.995[55]
IrO2-chitosan2.438%0.991[55]
Rh/RhO2a78%b[56]]
Rh2O3a70%b[56]
ReO3-PS-4-PVP2.864%0.977[54]
ReO3-chitosan2.853%0.997[54]
a Photodegradation rate constant k not informed. b Linear fit not informed.
Table 12. Bandgap for the Ir, Rh, and Re oxides.
Table 12. Bandgap for the Ir, Rh, and Re oxides.
PhotocatalystBandgap (eV)Ref.
ReO34.36[54]
IrO22.4–2.6[55]
Rh/RhO23.7[56]
Rh2O33.0[56]
Table 13. Kinetic data for the photodegradation process of MB with ThO2 and with the ThO2/SiO2, ThO2/TiO2 composites (Adapted from reference [21]).
Table 13. Kinetic data for the photodegradation process of MB with ThO2 and with the ThO2/SiO2, ThO2/TiO2 composites (Adapted from reference [21]).
PhotocatalystEg (eV)Apparent PhotodegradationDiscoloration Rate (%)R2 Linear Fit (%)
ThO2(chitosan precursor)5.663.7 × 10−3670.992
ThO2 (PS-4-PVP precursor)5.752.2 × 10−3660.967
ThO2/SiO2 (chitosan precursor)5.507.7 × 10−4240.979
ThO2/SiO2 (PS-4-PVP precursor)5.68.5 × 10−4250.923
ThO2/TiO2 (chitosan precursor)3.141.4 × 10−3390.815
ThO2/TiO2 (PS-4-PVP precursor)3.148.7 × 10−4270.941
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Díaz, C.; Segovia, M.; Valenzuela, M.L. Solid State Nanostructured Metal Oxides as Photocatalysts and Their Application in Pollutant Degradation: A Review. Photochem 2022, 2, 609-627. https://doi.org/10.3390/photochem2030041

AMA Style

Díaz C, Segovia M, Valenzuela ML. Solid State Nanostructured Metal Oxides as Photocatalysts and Their Application in Pollutant Degradation: A Review. Photochem. 2022; 2(3):609-627. https://doi.org/10.3390/photochem2030041

Chicago/Turabian Style

Díaz, Carlos, Marjorie Segovia, and Maria Luisa Valenzuela. 2022. "Solid State Nanostructured Metal Oxides as Photocatalysts and Their Application in Pollutant Degradation: A Review" Photochem 2, no. 3: 609-627. https://doi.org/10.3390/photochem2030041

Article Metrics

Back to TopTop