Next Article in Journal
Effect of the Second-Shell Coordination Environment on the Performance of P-Block Metal Single-Atom Catalysts for the Electrosynthesis of Hydrogen Peroxide
Previous Article in Journal
Photocatalytic, Antimicrobial, and Cytotoxic Efficacy of Biogenic Silver Nanoparticles Fabricated by Bacillus amyloliquefaciens
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Photocatalytic Activity of Metal- and Non-Metal-Anchored ZnO and TiO2 Nanocatalysts for Advanced Photocatalysis: Comparative Study

by
Hamad AlMohamadi
1,2,
Sameer A. Awad
3,
Ashwani Kumar Sharma
4,*,
Normurot Fayzullaev
5,
Arístides Távara-Aponte
6,
Lincoln Chiguala-Contreras
7,
Abdelfattah Amari
8,
Carlos Rodriguez-Benites
9,
Mohamed A. Tahoon
10,11 and
Hossein Esmaeili
12,*
1
Department of Chemical Engineering, Faculty of Engineering, Islamic University of Madinah, Madinah 42351, Saudi Arabia
2
Sustainability Research Center, Islamic University of Madinah, Madinah 42351, Saudi Arabia
3
Department of Medical Laboratories Techniques, Al-Maarif University College, Ramadi 31001, Iraq
4
Department of Applied Sciences, Chandigarh Engineering College, Chandigarh Group of Colleges, Jhanjeri, Mohali 140307, Punjab, India
5
Department of Polymer Chemistry and Chemical Technology, Samarkand State University, Samarkand 140101, Uzbekistan
6
Departamento Académico de Física, Universidad Nacional de Trujillo, Trujillo 13011, Peru
7
Facultad de Ingeniería Mecánica Eléctrica, Universidad Nacional Agraria de la Selva, Tingo María 10131, Peru
8
Department of Chemical Engineering, College of Engineering, King Khalid University, Abha 61411, Saudi Arabia
9
Dirección de Investigación, Centro de Investigación de la Creatividad, Universidad de Ciencias y Artes de América Latina, Av. La Molina 3755, Lima 15026, Peru
10
Department of Chemistry, College of Science, King Khalid University, P.O. Box 9004, Abha 61413, Saudi Arabia
11
Chemistry Department, Faculty of Science, Mansoura University, Mansoura 35516, Egypt
12
Department of Chemical Engineering, Bushehr Branch, Islamic Azad University, Bushehr 7515895496, Iran
*
Authors to whom correspondence should be addressed.
Catalysts 2024, 14(7), 420; https://doi.org/10.3390/catal14070420
Submission received: 6 June 2024 / Revised: 20 June 2024 / Accepted: 25 June 2024 / Published: 30 June 2024
(This article belongs to the Section Photocatalysis)

Abstract

:
This review article provides useful information on TiO2 and ZnO photocatalysts and their derivatives in removing organic contaminants such as dyes, hydrocarbons, pesticides, etc. Also, the reaction mechanisms of TiO2 and ZnO photocatalysts and their derivatives were investigated. In addition, the impact of adding metallic (e.g., Ag, Co, Pt, Pd, Cu, Au, and Ni) and non-metallic (e.g., C, N, O, and S) cocatalysts to their structure on the photodegradation efficiency of organic compounds was thoroughly studied. Moreover, the advantages and disadvantages of various synthesis procedures of ZnO and TiO2 nanocatalysts were discussed and compared. Furthermore, the impact of photocatalyst dosage, photocatalyst structure, contaminant concentration, pH, light intensity and wavelength, temperature, and reaction time on the photodegradation efficiency were studied. According to previous studies, adding metallic and non-metallic cocatalysts to the TiO2 and ZnO structure led to a remarkable enhancement in their stability and reusability. In addition, metallic and non-metallic cocatalysts attached to TiO2 and ZnO demonstrated remarkable photocatalytic efficiency in removing organic contaminants.

1. Introduction

Water scarcity has recently become a major problem owing to global warming, rapid industrial growth, depletion of water resources, environmental pollution, and uncontrolled development of groundwater [1,2]. The rapid industrial growth after the Industrial Revolution has remarkably increased the standard of living, but it has resulted in threats to human health [3]. Because of the development and growth of industries, industrial effluents are becoming more polluted and difficult to treat. The release of organic compounds and chemical fertilizers from various industries has polluted river water and has increasingly become a global pollution problem [4]. Moreover, it is difficult to completely remove non-degradable organic contaminants using biological treatment technologies [5]. Biological processes are safe, economical, and reliable, but the removal percentage of suspended solids is low, so they need better operation management [6]. The coagulation and precipitation techniques can suspend solid particles with the formation of flocs after adding polymeric coagulants and/or inorganic coagulants (such as Al, Fe, etc.). Coagulants bind solids together to form larger particles (flocs), then the flocs settle and separate from the effluent. This method has a high purification performance, but utilizing chemicals and generating biological sludge causes blockage of pipes and water deterioration, thus limiting the use of this process [7,8]. In the Fenton oxidation process, organic compounds are broken down using a strong oxidizer reagent. OH radicals can be produced by oxidizer reagents through a reaction between H2O2 and iron salts. This process is easy to apply, and additional equipment is not utilized excessively in comparison to other oxidation processes or photo-oxidation processes. However, this process has some drawbacks such as sludge production and high operating costs for secondary processes [9,10].
An advanced water treatment process can purify water at a high rate and effectively eliminate pollutants [11]. Advanced oxidation technology utilizes various techniques for the enhancement of oxidation power. For efficient water refinement, various contaminants must be removed economically [12]. Generally, various active species can play a key role in the photocatalytic reaction, including superoxide radicals (O2•−), holes (h+), and hydroxyl radicals (OH). These compounds play a crucial role in the efficient degradation of organic pollutants [13]. Advanced technology using TiO2 and ZnO photocatalysts has attracted the most attention for producing OH using optical energy without additional chemicals. Also, the operating cost of this technique may be significantly decreased by solar energy. The low photocatalytic performance of ZnO and TiO2 may limit the practical application of this advanced technology to treat wastewater [14]. Numerous studies have been performed to modify their surface by metal and non-metal cocatalysts for their possible practical utilization [15].
Various analyses are used to characterize the structure of ZnO and TiO2 nanocatalysts, including XRD for determining crystal phases and crystal size, AFM for determining refractive index profiles, FTIR for specifying the functional groups, and EDX for determining the constituents. All these features are critical in the photocatalytic reaction. Also, SEM and TEM analyses are used to characterize the morphology and appearance structure of a catalyst [16,17]. Figure 1 demonstrates SEM and TEM photos of ZnO and TiO2 nanocatalysts. As shown, ZnO and TiO2 have various morphologies, shapes, and sizes.
TiO2 and ZnO photocatalysts are extensively utilized for photodegradation of contaminants under UV light or sunlight. However, owing to their band gap range, they have limitations for the photodegradation of pollutants. For doing so, TiO2 and ZnO photocatalysts can be combined with other compounds to synthesize efficient photocatalysts, some of which include La/TiO2 [20], CuO/WO3/TiO2 [21], Fe3O4/TiO2 [22], carbon-anchored TiO2 [23], potassium titanate (K2Ti6O13) [24], rGO/Fe3O4/ZnO [25], tungsten/silver/ZnO [26], Pb/ZnO, Cd/ZnO, Ag/ZnO [27], and Bi(12)ZnO(20) [28]. There are two common strategies for the synthesis of catalysts, including doping and heterojunction. Belousov and coworkers compared the photocatalytic activity of Bi2WO6 for water treatment. To prepare the catalyst, they used two methods including doping and heterojunction. Using the hydrothermal process, they synthesized Bi2WxMo1–xO6 through the doping strategy and Bi2WO6/g-C3N4 through the heterojunction method, and then employed them for the photodegradation of methylene blue. According to their results, the highest photodegradation efficiency of methylene blue was obtained using the catalyst prepared by the doping strategy. Overall, the attractiveness of the doping technique was seen in terms of photocatalytic efficiency, turnover number, and turnover frequency compared to the heterojunction method [29]. In this review article, the performance of the doping strategy for the development of ZnO and TiO2 catalysts in the removal of organic pollutants was investigated.
This paper focuses on comparing the photocatalytic capabilities of TiO2 and ZnO and their composites in the degradation of organic compounds in wastewater. Also, the advantages, disadvantages, and structural features of these photocatalysts were thoroughly investigated. Moreover, the photodegradation mechanism of contaminants using TiO2 and ZnO as well as metal and non-metal cocatalysts anchored to TiO2 and ZnO were completely discussed. Furthermore, the photodegradation efficiency of different contaminants using these photocatalysts and their derivatives (metallic and non-metallic cocatalysts in their structure) were compared. One of the main features of catalysts is their stability in the photocatalytic reaction, so the recyclability of TiO2 and ZnO and their composites were fully studied. Therefore, this article makes a comprehensive comparison between these two catalysts and their composites in the photodegradation of organic pollutants, their photocatalytic efficiency, mechanisms, and reusability, which have not been scrutinized in previous review articles.

2. Organic Pollutants

Municipal sewage pollutants mainly include organic substances, a variety of pathogenic microorganisms, and so on, with complex structures such as drugs, antioxidants, polycyclic aromatic hydrocarbons, pharmaceutical intermediates, food additives, PPCPs, steroids, and phenolic compounds [30]. Also, pollutants in industrial wastewaters include gas condensates [5], petroleum products, non-chlorinated compounds, petroleum hydrocarbons, trinitrotoluene, and polycyclic aromatic hydrocarbon [31]. Hydrocarbon contaminants can enter wastewater through various industries such as oil and gas refineries, urban waste, dairy, the food industry, and slaughterhouses, and cause problems for human health and the environment. These pollutants, in addition to disrupting biodiversity, can directly or indirectly enter human food chains, reduce the quality of life, and cause many diseases. This type of wastewater contains aromatic hydrocarbons, dissolved gases, dispersed oil, inorganic salts and acids, suspended particles, ketones, and phenols. This type of wastewater is available in the form of oil/water emulsion [3,32]. The discharge of these wastewaters to sea must be performed under certain conditions. The highest permissible concentration of hydrocarbon compounds for discharge into the sea should be 40 ppm, and TDS less than 32,000 ppm [32].
Another kind of contaminant in wastewater is volatile organic compounds (VOCs), which have destructive effects on human health. Volatile organic pollutants are categorized into three groups, including semi-volatile organic compounds (SVOCs), volatile organic compounds (VOCs), and very volatile organic compounds (VVOCs). Propane and butane are some VVOCs which are poisonous compounds in very low concentrations. VOCs are also toxic compounds found in household products and the environment, including isopropyl alcohol, acetone, formaldehyde, toluene, vinyl chloride, and hexanal. The molecular weight and boiling point of SVOCs are higher than those of VOCs and can evaporate at room temperature. Pesticides such as chlordane and plasticizers such as phthalates are some of the SVOCs [33]. Pollution caused by pesticides disrupts the functioning of aquatic ecosystems, decreases biodiversity, reduces food sources, interferes with the concentration of dissolved oxygen, destroys the habitats of aquatic organisms, and increases the amount of biochemical oxygen demand (BOD) [34]. The utmost allowable limit of organic compounds for drainage into the river for utilization in agricultural irrigation should be less than 200 ppm [35].
The main source of VOCs includes incomplete burning of fossil fuels, solvents used in inks and paints (e.g., acetone, ethyl acetate, and glycol ether), utilizing biofuels (e.g., cooking oil and bioethanol), biomass combustion, especially from agricultural residuals, and VOCs released from metal working fluids [33]. The presence of these contaminants in sewages results in various problems for the environment. The important symptoms of exposure to these compounds include dizziness, headache, and in acute cases, it can lead to anesthesia. In more severe cases, respiratory and circulatory disorders can lead to death [33].
Also, tetracycline, sulfamethoxazole, cefixime, amoxicillin, gentamicin, erythromycin, and ciprofloxacin are kinds of antibiotics that are present in hospital effluents. The presence of antibiotics in effluents in low and high concentrations is dangerous and causes serious damage to human health. The presence of antibiotics in wastewater has detrimental impacts on human health and the aquatic environment owing to their toxicity, genotoxicity, and mutagenicity. According to reports, the concentration of antibiotics in surface and groundwater reaches 100 micrograms per liter. Also, the concentration of amoxicillin in raw effluent and urban wastewater treatment is 171 mg/L and 13 ng/L, respectively [36]. The acceptable standard of WHO for antibiotics in wastewater is 1 mg/L [37]. Therefore, their concentrations in wastewater should be attenuated.
Dyes are another kind of organic compound which are utilized in many industries like food, plastics, paper, cosmetics, and textiles. Among them, the most abundant dyes are related to the textile industry, and these dyes can enter the wastewater [38]. More than 7 × 105 tons of dyes are produced annually, 10–15% of which enters the sewage during the dyeing process. Water pollution by dyes results in many environmental problems. The highest allowable limit of dyes in industrial discharge is in the range of 0.01–0.05 mg/L. Different groups of dyes are considered carcinogens or mutants. Dyes can prevent sunlight from entering the water and disrupt biological process in the water. Most dyes are poisonous to organisms and have adverse impacts on the aquatic life and the entire ecosystem. The presence of dye in water causes reproductive dysfunction, kidney damage, central nervous system damage, liver damage, brain damage, carcinogenesis, and mutagenicity. Basic dyes are cationic dyes and most of them are crystalline compounds, which can be derived from positively charged sulfur or nitrogen atoms. Under visible and UV light irradiation, dyes have high stability and insensitivity and do not degrade without photocatalysts [39,40,41]. Cationic, anionic, and non-ionic are three important categories of dyes, among which the toxicity of cationic dyes is more than the other two categories. There are various types of dyes in sewage, including crystal violet, Congo Red (CR), methyl orange (MO), reactive orange, reactive black, reactive red, methylene blue (MB), methyl violet, brilliant cresyl blue, and Safranin-T that may be generated by the plastic, leather, food dye, printing, paper, and textile industries [42,43]. Figure 2 illustrates different categories of organic compounds with some examples.

3. Various Processes for Wastewater Treatment

There are different processes for wastewater treatment, including the photocatalytic process, adsorption, coagulation, ion exchange, electrochemical, electrocoagulation, chemical deposition, and membrane technology [1,3]. Figure 3 demonstrates various processes to eliminate pollutants. Also, the advantages and disadvantages of each process can be seen in this figure [1,11]. As shown, the sorption process has many advantages such as low cost, high efficiency, simplicity, and wide adaptability [44,45]. However, this method has some drawbacks. For instance, the removal efficiency of microorganisms by the sorption process is low. Also, fine particles cannot be effectively removed from wastewater. Another method is ion exchange, which is not efficient for removing hydrocarbon and organic compounds. In addition, the advantage of membrane technology is to produce less waste. However, energy consumption in the membrane is high. Although the electrochemical process has high selectivity for most ions, this technique has high capital cost and high energy consumption. Among these approaches, the photocatalytic process is effective for removing different kinds of organic pollutants such as hydrocarbons, gas condensates, and oil from wastewater [1,11,46]. There are two important catalysts, namely TiO2 and ZnO, which have been widely used by previous researchers. In the following sections, their advantages and disadvantages, reaction mechanisms, reusability, and performance in wastewater treatment are discussed.

4. TiO2 Photocatalyst

TiO2 is a non-toxic and insoluble substance in water. TiO2 has high stability with a high catalytic performance. Other properties of TiO2 include high specific surface area, high crystallinity, and high concentration of surface hydroxyls. The specific surface area is an important parameter in photocatalytic efficiency and usually a high value of the specific surface area is associated with a low crystallinity of the catalyst [47,48]. Another name for TiO2 is Titania, which can be available in three forms such as anatase, brookite, and rutile [48]. The rutile (tetragonal) crystalline structure of TiO2 is quite stable in particles greater than 35 nm [49]. TiO2 in anatase and rutile structures have oxidation strengths of 3–3.2 eV, which are very strong oxidation strengths [14]. One of the most important applications of TiO2 is its utilization in the photocatalytic process. To this end, UV light, halogen lamps, or solar irradiation can help accelerate the photocatalytic reaction in the presence of TiO2 nanocatalysts [50]. The photocatalyst has two important properties. For instance, the photocatalyst should not be consumed or participate directly in the reaction. Also, the photocatalyst provides other mechanism pathways. The photocatalytic reaction is based on the absorption of solar energy in the semiconductor gap. Few semiconductors can be utilized as the catalyst in the photocatalysis process, of which, TiO2 is the most extensively utilized [48]. Various morphologies of TiO2 consist of zero-, one-, two-, and three-dimensional structures. By optimization of the shape and size, the photocatalytic activity of TiO2 particles in the wastewater treatment process can be maximized [51].
TiO2 can be synthesized easily by various physical, chemical, and thermal procedures such as chemical or physical vapor deposition, sol-gel, inverse micelle, solvothermal, hydrothermal, sonochemical, microwave, and electrodeposition processes [52]. Synthesizing metallic semiconductor oxides by conventional physical mixing or chemical deposition usually produces insoluble materials that are inherently difficult to control in terms of their morphology, size, and dispersing metal components. These methods need a long time and multi-stage processes. The sonochemistry procedure is an efficient process for preparing mesoporous materials. Ultrasound is useful to synthesize an extensive range of nanostructured materials, which include high specific surface area oxides, alloys, carbides, and transition metals [15]. Neppolian et al. synthesized TiO2 nanophotocatalysts by a combination of ultrasonic and sol-gel processes. They investigated the impact of various factors on the synthesis method, including magnetic stirring, ultrasonic sources (e.g., bath and horn), ultrasonic time, power density, temperature, and reactor size [53]. Table 1 indicates the advantages and disadvantages of various processes for TiO2 synthesis. As shown, the production of high-quality crystals as well as easy control of crystals in TiO2 can be achieved by the hydrothermal process. Also, the fabrication of TiO2 with the sol-gel process has several benefits such as high purity products, good size distribution, remarkable specific surface area, economical, uniform size of particles, fine particle size, and simple synthesis. Among these processes, utilizing the microwave method has significant benefits such as a short reaction time, high reaction rate, and high efficiency [54].

4.1. Features and Reaction Mechanism of TiO2 Photocatalyst

Utilization of TiO2 as a photocatalyst for the elimination of contaminants from wastewater has several benefits, which include the following: (1) the process is conducted at room temperature and 1 atm, (2) complete decontamination without secondary contamination, (3) producing high surface area and high catalytic activity, and (4) using the photocatalyst in multiple cycles and decreasing the costs [56]. The decomposition process of contaminants is mainly an oxidative reaction, which depends on the features of the photocatalyst. TiO2 is also low-cost, has good band gap energy, long-term stability against light, and is safe. Generally, the band distance and wavelength of TiO2 are between 3 and 3.2 eV and 400 nm, respectively, indicating that UV light at a wavelength less than 400 nm leads to an inverse reaction. Also, UV irradiation at a wavelength less than 400 nm initiates a photoreaction. Unfavorable recombination of holes (h+) and electrons (e), and low yield in the visible region under irradiation are two important disadvantages of TiO2 [15]. The catalytic activity of the TiO2 photocatalyst can be enhanced under UV irradiation because TiO2 uses only 5% of the solar energy [57].
At UV light < 400 nm, the surface of TiO2 may reach temperatures above 30,000 °C; this temperature can oxidize all substances. Hence, all organic compounds are completely decomposed into CO2 and water. Figure 4a illustrates a schematic of the degradation of contaminants by forming photo-induced charge carrier electrons/holes (e/h+) on the TiO2 surface. Also, Figure 4b shows that the catalyst surface is surrounded by the contaminant molecules. On the TiO2 surface, oxygen molecules react with electrons to produce oxygen radicals. Also, water molecules react with h+ to generate hydroxide radicals and H+ ions. The final products will be CO2 and H2O [14]. In other words, electrons introduced into the conduction band from the valence band can be easily transferred to the catalyst surface and are trapped by binding to noble metals, intermediate metals, and rare earth elements. Metallic cations increase the ability of radicals produced in the photocatalysis reaction and then decrease the recombination life of e/h+ [58].
By irradiating UV light to the TiO2 surface, the photo-induced electrons react with dissolved oxygen for producing O2. The photo-induced h+ in the valence band penetrates the surface of the TiO2 photocatalyst and reacts with the absorbed water molecules to form OH. OH• is a highly active species in the photocatalytic process. The reaction mechanism of TiO2 and the electron–hole pair is defined as follows. As shown, TiO2 reacts with hv to produce e and h+. Next, e and h+ can produce two important radicals such as OH and O2, which have important roles in the photocatalytic reaction [14,58].
TiO 2 + hv TiO 2   ( e C B + h V B + )
Oxidation : e C B   +   H 2 O   +   OH + H +
Reduction : O 2   +   e C B O 2
e C B + HO 2 HO 2
h V B + + HO 2 H 2 O 2
T i O 2 h + + H 2 O a d s T i O 2 + H O a d s . + H +
T i O 2 h + + H O T i O 2 + H O a d s .
T i O 2 e + O 2 T i O 2 + O 2 .
O 2 . + H 2 O H O 2 . + H O
O 2 + H + H O 2
2 H O 2 . H 2 O 2 + O 2
T i O 2 e + H 2 O 2 T i O 2 + H O + H O .
T i O 2 h + + R a d s T i O 2 + R a d s . +
where, R is the adsorbed pollutant [59].
Also, TiO2 is photochemically stable and non-toxic. Under irradiation, the charge-pair on TiO2 reacts directly with solid lattice ions. Also, TiO2 is resistant to various acidic and alkaline pHs. TiO2 can be utilized in the photocatalytic treatment of the environment. Hence, TiO2 can be used as a promising photocatalyst to degrade organic contaminants. Nevertheless, TiO2 particles are inactive in visible light. To solve this problem, numerous studies have been performed on the synthesis of various composites of TiO2 with metals and non-metals (e.g., ceramics, zeolites, carbon materials, fibers, and glasses) as supports in order to have photocatalytic activity in a wide range of visible light [14].
The recombination behavior of e/h+ decreases with the decreasing size of TiO2 NPs, which can be because of the increase in interfacial charge transfer at TiO2 surfaces. Also, the photocatalytic activity of TiO2 of less than a few nanometers decreases because of the predominant recombination of the electron/hole at the TiO2 surface. According to the reports, TiO2 nanotubes are more effective in degrading contaminants than TiO2 particles, indicating the rapid mass transfer of contaminants on the surface of nanotubes [49].

4.2. Improved Photocatalytic Activity of TiO2 Using Metallic and Non-Metallic Cocatalysts

As mentioned earlier, TiO2 photocatalysts can decompose organic contaminants under UV irradiation. Researchers have tried to improve the optical sorption range of TiO2 catalysts from ultraviolet irradiation to visible light to boost its photocatalytic activity. Therefore, for improving the photocatalytic activity of TiO2 in visible light, its surface should be modified [60]. The catalytic activity of TiO2 can be improved by placing suitable materials on the surface of TiO2 for its application under visible irradiation [57,61,62]. Metallic particles such as Pd, Pt, Fe, Ru, Au, and Ag can be utilized on the TiO2 surface to improve photocatalytic activity by suppressing recombination behaviors of e/h+. The induced electrons migrate to the surface of the metal particles, stabilizing the photo-induced h+ on TiO2 as the life time of the charge carrier increases. To this end, more superoxide (O2) and OH radicals can be produced. As the crystallographic aspects of TiO2 increase, the photocatalytic activity of TiO2 also produces more OH•, which can contribute to the photodegradation of organic pollutants. Also, the element O in the lattice of TiO2 may be replaced by various heteroatoms such as B, S, F, N, P, and binary elements B-C and N-S to perform photocatalytic reactions in the presence of visible light. In addition, metals and non-metals introduced to the TiO2 structure can improve its catalytic activity for the decomposition of organic pollutants [14,63]. Also, modifying the surface of TiO2 improves the uptake capacity of contaminants, which can be useful in advanced oxidation technology. The modified TiO2 with nanotubes, foams, and mesoporous phases has shown better photocatalytic behavior compared to the unmodified TiO2 [64].
The TiO2 photocatalyst on a nanoscale has a great surface area/volume proportion, leading to increased charge separation and ion trapping at the TiO2 surface. TiO2 nanoparticles show enhanced oxidative power compared to TiO2 microparticles. However, the TiO2 nanocatalyst cannot be employed directly in wastewater treatment owing to the aggregation of particles during the degradation process as well as their physical and chemical features. To this end, the catalytic activity of TiO2 NPs can be improved by incorporating TiO2 onto carbon materials for composite synthesis such as graphene, carbon nanotubes, and activated carbon nanofibers. Carbon materials can be excellent supports for TiO2 NPs owing to their unique features such as excellent mechanical, thermal, chemical, electrical, and optical properties, which lead to a rapid charge transfer on the surface of TiO2/carbon nanocomposites [65]. TiO2/carbon nanocomposites are suitable catalysts for the elimination of organic contaminants from effluents. Graphene NPs act as a bridge for excellent electron transfer and electron sinks. TiO2/graphene nanocomposites have an extensive range of band gaps (2.66–3.18 eV), suggesting that long-wavelength lights in the visible region can be absorbed by the TiO2/carbon nanocomposite. One of the greatest benefits of carbon nanomaterials is their high specific surface area, in which TiO2 NPs may be distributed in their structure, leading to improved selectivity of organic contaminants. The TiO2/carbon nanocomposite has a good impact on self-purification in comparison to UV irradiation and ozonation when used for wastewater treatment. These nanocomposites can be effective industrially for wastewater purification because carbon NPs are employed as a support to stabilize the composite structure [66].
The photocatalytic activity of TiO2 can be enhanced by enhancing the interfacial charge transfer and reducing the recombination of e/h+. To this end, various metals can be used to improve the photocatalytic activity of TiO2, including Ag, boron (B), Si, Ni, and so on under UV light or sunlight irradiation. Generally, there are several ways to improve the photocatalytic activity of TiO2, including enhancing the ratio of surface/volume, integrating TiO2 with other semiconductors, optimizing its particle size, and anchoring metals or non-metals. The presence of metal cocatalysts in the TiO2 structure remarkably affects its photo-reactivity because by shifting the catalyst band gap to the visible region, it changes the interfacial electron transfer rate and the charge-carrying recombination rate [67]. A cocatalyst ion can trap e or h+, leading to a longer lifespan of the produced charge carriers and increased photocatalytic activity. Figure 5 indicates various cocatalysts used in the TiO2 structure. As shown, noble metals, transition metals, anionic compounds, metal oxides, and transition metal ceramics are the main cocatalysts for improving the TiO2 surface. Au, Pt, Pd, Rh, and Ag are some noble metals. Also, Fe, Al, Zr, Co, Cu, and Ni are several types of transition metals. Moreover, Fe2O3, Cr2O3, and SiO2 are the most important metal oxides. Furthermore, WO3, SnO2, and MoO3 are the most important transition metal ceramics. Eventually, C, N, F, O, and S are non-metallic cocatalysts. The addition of transition metals to the TiO2 surface can enhance its catalytic activity and reduce the recombination of photogenerated e and h+ [15,68]. Previous studies have shown that the addition of metals and metal oxides to the catalyst surface improves the photocatalytic activity significantly compared to semi-conductors [68]. The photodegradation mechanism of contaminants using metal- and non-metal-anchored TiO2 is also illustrated in Figure 6 [69,70]. As shown, the presence of metals and non-metals on the TiO2 surface can help improve its photocatalytic activity by producing more oxygen radicals. These oxygen radicals play a key role in the photodegradation of organic contaminants.
Numerous studies have been performed on the expansion of the solar absorption band by combining TiO2 with new metals, intermediates, and anions, which maximizes the performance of the photocatalyst [49]. Table 2 presents the influence of various metal cocatalysts on the photocatalytic activity of TiO2. Accordingly, Wang et al. (2019) synthesized a 5% Pd-TiO2 photocatalyst to degrade 2,2′,4,4′-tetrabromodiphenyl ether from water under a UV lamp and showed 100% photodegradation efficiency, which is a remarkable amount [71]. In another work, Yadav et al. (2020) utilized a B-TiO2 nanophotocatalyst for 4-nitrophenol removal from water. The utmost photodegradation efficiencies using TiO2 and B-TiO2 were 79 and 90%, respectively, indicating that the photodegradation efficiency of TiO2 was enhanced with its doping by boron [72]. Also, Sescu et al. (2020) synthesized a Au/TiO2 photocatalyst using two different procedures, incipient wet impregnation (IWI) and ultrasound impregnation (UI). Then, Au/TiO2 (IWI) and Au/TiO2 (UI) were used for the degradation of 2,4 dinitrophenol under UV irradiation. According to their outcomes, the photodegradation efficiency of 2,4 dinitrophenol using Au/TiO2 (IWI) and Au/TiO2 (UI) were 50 and 37%, respectively, which were less than the photodegradation efficiency of pure TiO2 (60%). Comparing these two methods displays that the IWI procedure is more efficient than the UI procedure. Therefore, they used the IWI procedure to synthesize Pd/TiO2 for the degradation of 2,4 dinitrophenol and Rhodamine 6G contaminants under UV irradiation. The photodegradation efficiencies of 2,4 dinitrophenol and Rhodamine 6G using the Pd/TiO2 photocatalyst were 67% and 96%, respectively, indicating that the Pd/TiO2 photocatalyst can remove Rhodamine 6G dye with a significant efficiency [73]. In general, previous studies show that most photocatalytic processes are performed under UV light. Also, Ag and Pd cocatalysts showed greater potential for photodegradation of organic contaminants than other cocatalysts.

5. ZnO Nanophotocatalyst

The ZnO nanocatalyst is one of the most efficient catalysts in the degradation of organic contaminants due to its unique features like excellent oxidation ability, direct and extensive band gap in the spectral region close to UV, high photocatalytic activity, and high binding energy. UV light can be absorbed on ZnO at a wavelength lower than 385 nm [78]. The rock salt, wurtzite, and cubic structures are important crystalline structures of ZnO. Among the three structures, the wurtzite structure of ZnO is the most common structure and has the utmost stability. Also, ZnO in the form of rock salt is wholly rare. At room temperature and pressure, the crystalline structure of ZnO is a hexagonal wurtzite. Moreover, ZnO can crystallize in the structure of wurtzite. The ZnO wurtzite structure is the most common form of ZnO due to its stability in environmental conditions [79]. Different structures of ZnO are exhibited in Figure 7. In the ZnO wurtzite structure, each Zn atom is surrounded by four O atoms. ZnO has many benefits compared to TiO2, including low cost, chemical stability, no toxicity, and abundance [80]. Other properties of ZnO include its insolubility in water, that it is odorless, and has a bitter taste. ZnO is used in catalysis processes, fertilizers, the rubber industry, the paint industry, and cosmetics. The development of nano-ZnO with precisely controllable properties has recently gained considerable scientific attention [81].
Different zinc salts can be utilized to fabricate ZnO, including Zn(C2H3O2)2.2H2O, Zn(SO4)2.7H2O, Zn(NO3)2·6H2O, and ZnCl2 [82]. ZnO nanoparticles and doped ZnO can be synthesized by various techniques, including pulsed-laser deposition, chemical coprecipitation, hydrothermal, thermal decomposition, sol-gel, liquid–solid solution, vapor condensation, microwave, and spray pyrolysis processes [78,83]. The hydrothermal procedure is a well-known and efficient process for synthesizing ZnO nanoparticles, which is performed at high pressure and temperature [83]. Chemical coprecipitation is one of the most successful processes to synthesize ZnO nanoparticles with a fine particle size distribution. Chemical coprecipitation can prevent complex steps like alkoxide reflux and therefore takes less time than other methods [78]. Although the chemical coprecipitation process has many benefits, this process will leave a large amount of solution, which causes waste and high costs for this method [84]. Compared to the coprecipitation process, the yield of the product obtained from the hydrothermal process is lower [85]. Also, the synthesis of ZnO nanoparticles by the sol-gel process has several important advantages such as a uniform distribution of particles and synthesis at low temperatures. However, microwave-based synthesis has attracted a lot of attention due to its advantages such as being simpler, faster, and more energy efficient. The precursor solution is irradiated by the microwave source. Energy transfer via relaxation and/or resonance can lead to a relatively fast heating process. Also, the heating process by the microwave source leads to uniform heating in a short time, which results in a uniform distribution of particles [86]. In addition, pulsed laser deposition processes can be easily understood, but their theoretical aspect is still not fully understood because of the complexities of the interaction between the laser beam and the target substance [87]. Table 3 indicates the advantages and disadvantages of various processes for the synthesis of ZnO. According to the characteristics of the different processes listed in Table 3, it seems that sol-gel and microwave processes can be suitable methods for ZnO synthesis compared to other techniques.

5.1. Characterization of ZnO Nanocatalyst

The ZnO nanocatalyst is a semiconductor with a broad band gap, as well as high binding energy in environmental conditions. ZnO NPs have better optical, electrical, and magnetic features than ZnO microparticles. ZnO NPs also have unique physical, thermal, and chemical properties like biocompatibility, low cost, and non-toxicity. ZnO is an environmentally friendly substance, which is well suited for a wide range of daily applications that pose no risk to human health. Also, ZnO NPs have a longer shelf life compared to other metal oxides like Fe2O3, SiO2, WO3, and TiO2. Because of its exceptional features, ZnO nanocatalysts can be utilized as valuable catalysts to treat wastewater [82]. Considering that ZnO has the same band gap energy (3.2 eV) as TiO2, their photocatalytic capabilities are similar. Also, ZnO nanocatalysts are relatively cheaper than TiO2 nanocatalysts, and the use of ZnO is more economical in large-scale water purification processes. ZnO can adsorb a broad range of solar spectra and more optical quantum than many semiconductor metal oxides. However, the main disadvantage of ZnO is its wide band gap energy, therefore, the absorption of light by ZnO is limited to the visible light range. This leads to rapid recombination of light-generated charges, resulting in low photocatalytic performance [92]. The addition of cocatalysts affects the optical features of ZnO and possibly shifts the amplitude of optical absorption to the visible region. Therefore, the photocatalytic activity of ZnO can be enhanced by doping it with other materials, which is described in detail in Section 4.2.

5.2. Photocatalysis Mechanism of ZnO

In a desirable photocatalytic reaction in the presence of zinc oxide particles and active oxidizing species like oxygen, organic contaminants can be converted to H2O, CO2, and inorganic acids. The photocatalytic reaction begins when ZnO NPs absorb photons from light with energies greater than their band gap energy. Thus, the photo-induced electron rises from the valence band to the conduction band, creating h+ and e on the ZnO surface [93]. The presence of the O atom as an electron absorber lengthens the pair of recombinant electron cavities and the formation of superoxide radicals. The reaction between h+ and OH results in the production of hydroxyl radicals. Figure 8a indicates the photocatalytic mechanism of anionic- and cationic-anchored ZnO catalysts and the photodegradation mechanism of the contaminant on the ZnO surface under sunlight is illustrated in Figure 8b. The reaction mechanism of ZnO in the photocatalytic process has several steps, which are presented below [94,95]:
Z n O + h v e + h +    
Absorbed   oxygen :   O 2 + e O 2 .
Water   ionization :   H 2 O O H + H +
Superoxide   protonation :   O 2 . + H + H O O . H O O . + e H O O H O O + H + H 2 O 2 H 2 O 2 + e O H + O H . H 2 O + h + H + + O H

5.3. Improving the Photodegradation Efficiency of ZnO

The recombination of photo-produced h+ and e is one of the main disadvantages of semiconductor photocatalysis. This recombination stage reduces quantum efficiency and wastes energy. Thus, the recombination process of e/h+ must be inhibited to have an impressive photocatalysis process. Metallic cocatalysts can counteract the recombination problem by increasing the isolation of charge between h+ and e. Also, cocatalysts can trap e and reduce the chance of electron/hole recombination, which inactivates the photocatalytic process. Moreover, the production of active oxygen species and hydroxyl radicals significantly increases the charge separation efficiency. In general, the cocatalyst concentration, operating conditions, and synthesis procedure have significant impacts on the metal-anchored semiconductor photocatalysis process [96].
Anion-anchored zinc oxide photocatalysts have shown higher photocatalytic degradation efficiency than pure zinc oxide. The presence of isolated N2p modes above the utmost ZnO valence band in the Ne ZnO sample increases its ability to absorb visible light. Under visible light irradiation, narrow-band gaps in N-ZnO require less energy to induce the charge carriers (e and h+). The increased photocatalytic activity of C–zinc oxide can have several reasons, including increased adsorption of contaminants on the surface of the catalyst, higher UV absorption than pure ZnO, and the creation of new energy levels below the conduction band of ZnO, where e photoexcited are eliminated by these new energy levels to prevent electron-hole recombination. All these factors increase the number of charge carriers and improve the photodegradation performance. Also, the number of oxygen vacancies in ZnO has a key role in influencing the photo-activity of ZnO attached to S. In the photocatalytic reaction, oxygen vacancies become centers for trapping e. Therefore, the more oxygen vacancies, the greater the catalytic activity [97,98]. Figure 9 shows the mechanism of pollutant degradation by metal- and non-metal-doped ZnO [98]. As shown in Figure 9a, the metal anchored on the ZnO surface contributes to the generation of e and subsequently O2 radicals, resulting in better photodegradation of organic contaminants. The same role is played by non-metals attached to ZnO to generate O2 radicals followed by the photodegradation of pollutants (Figure 9b) [99].
Despite the high photodegradation efficiency of ZnO, studies show that the anchored ZnO has significant photocatalytic efficiency compared to the undoped ZnO. ZnO surface modification using cationic cocatalysts has attracted much attention. The chemical, electrical, magnetic, and structural features of ZnO can be adjusted by adding cationic cocatalysts such as Ni, Mn, Co, Al, and Sb. The anchored elements are usually iso-morphic to the Zn ion, like Cu(II), Co(II), Ni(II), and Mn(II). The doped ZnO photocatalyst shows a faster response to the degradation of organic contaminants than pure ZnO [100]. Table 4 reports the impact of different metallic cocatalysts on the photocatalytic efficiency of ZnO. As shown, Adeel and coworkers (2021) studied the degradation process of MO using Co/ZnO. The photodegradation efficiency of ZnO and Co/ZnO were 46 and 93%, showing that the photodegradation efficiency of ZnO doubled after its doping with Co [101]. In another work, Vallejo et al. (2020a) compared the photodegradation efficiency of MB using ZnO and Ag/ZnO. Their results displayed photodegradation efficiencies of 2.7 and 45.1% for ZnO and Ag/ZnO, respectively, indicating a remarkable increment in photodegradation efficiency after ZnO doping with Ag. These photodegradation efficiencies were attained after 120 min of visible irradiation [102]. Also, Vallejo et al. (2020b) synthesized Co/ZnO and Cu/ZnO photocatalysts and compared their photodegradation efficiencies with ZnO in MB removal under visible light after 140 min. Their findings demonstrated that the photodegradation efficiencies of ZnO, Co/ZnO, and Cu/ZnO were 2.7, 62.6, and 42.5%, respectively [103]. Moreover, Cr/ZnO was able to decompose MO dye with an excellent photodegradation efficiency of 99.8% under UV-vis light illumination after 100 min [104]. Among these works, Cr/ZnO showed the highest photodegradation efficiency (99.8%) for removing methyl orange after 100 min under UV light.
Table 5 also reports the impact of various non-metallic cocatalysts on the ZnO photocatalytic activity for some organic contaminants. As shown, Wu and coworkers (2014) studied the photodegradation performance of MB using N-anchored ZnO under visible light irradiation and compared its results with pure ZnO. Their outcomes indicated that the photodegradation efficiency of MN using pure ZnO and N-anchored ZnO photocatalysts are 76.2 and 81.6%, respectively [108]. Also, Li and coworkers (2012) used pure ZnO and C-anchored ZnO for photodegradation of MB under visible light. After 60 min, C-anchored ZnO was able to decompose MB with a photodegradation efficiency of 98.1%, which is a significant amount compared to pure ZnO (54.3%) [109]. Moreover, the photodegradation efficiency of Bisphenol A using C-anchored ZnO under UV irradiation was 100%, indicating high degradation efficiency [110]. Furthermore, Fu et al. (2012) studied the photocatalytic reaction of MB dye using C-anchored ZnO under visible light and compared the photodegradation efficiency of ZnO and C-anchored ZnO. The highest photodegradation efficiency of ZnO and C-anchored ZnO were 26 and 80% under optimal conditions such as 2.5% catalyst and 200 °C, which showed a significant increase in the photodegradation efficiency after ZnO doping with C [111]. In general, most studies have been performed by N- and C-anchored ZnO photocatalysts, and both cocatalysts showed significant photodegradation activity. Therefore, these studies demonstrate that utilizing C- and N-anchored ZnO photocatalysts is more attractive to researchers than other cocatalysts.

6. Recyclability of ZnO and TiO2 Photocatalysts

The most important factor for the practical application of catalysts is their recyclability and stability. Catalyst recyclability shows how many times a catalyst can be utilized in the catalytic process. After repeated reuse, the degradation efficiency of the catalyst is reduced, which may be due to the saturation of active sites, deactivation of active sites, and chemical decomposition of the catalyst structure. If the degradation efficiency changes slightly after repeated use of the catalyst, it indicates that the catalyst is highly reusable. Thus, determining the reusability of a catalyst is critical [18,115]. Another important parameter in the stability of catalysts is the leaching of elements from the surface of the catalyst into water [116]. Usually, the leaching of cocatalysts takes place during the photocatalytic reaction, which reduces the reusability of catalysts [117]. The transition metal on the catalyst surface is intentionally eliminated by acid treatment, cyclic voltammetry, or thermal dealloying to form a core-shell structure. Because of differences in surface energy or lattice mismatch, transition metal atoms can be simply separated. Furthermore, when oxygen species, which are oxygen reduction reaction intermediates, are placed on the metal atom on the catalyst surface, the separation of metal atoms is accelerated. Choi et al. (2019) stated that the use of the gas phase reduction method increases the durability and activity of the catalyst and minimizes the leaching of cocatalysts from the catalyst surface [118]. Also, the generation of peroxides during the photocatalytic reaction can lead to deactivation and performance limitation, thus limiting its industrial applications [116]. Table 6 reports the reusability of various composites of ZnO and TiO2 in different cycles.
Nikoofar et al. (2015) synthesized ZnO nanorods for imidazole removal. They studied the reusability of ZnO nanorods in four runs and their results showed that the elimination efficiencies of imidazole after the first, second, third, and fourth cycles were 83%, 83%, 81%, and 80%, respectively, which indicates that the elimination efficiency has declined slightly (less than 3% after four reuse cycles), showing remarkable reusability of ZnO nanorods [119]. In another study, the reusability of the Ag/ZnO nanocomposite was studied in eliminating MB, MO, and CR after five successive steps. A slight decrease in dye removal efficiency was observed after five cycles, indicating significant recyclability of Ag/ZnO for eliminating organic dyes [129]. Also, Shelar et al. (2020) investigated the stability and recyclability of the Ag-ZnO nanocatalyst for eliminating MB dye. The photodegradation efficiency of MB was 95%. Then, the nanocatalyst was eluted three times with deionized water and reused in the photocatalytic process. After four cycles, the photodegradation efficiency reduced to 89.5%, indicating a high stability of the Ag-ZnO nanocatalyst [120]. Moreover, Fe3O4@S-ZnO was utilized for degrading ofloxacin. The recyclability study illustrated that the photodegradation efficiency of ofloxacin was above 90% after six cycles, indicating remarkable stability and reusability of Fe3O4@S-ZnO [121]. Furthermore, Khan and coworkers (2021) studied the reusability of ZnO NPs for photodegradation of MB and Rhodamine B dyes after five consecutive cycles. The photodegradation efficiencies of MB and Rhodamine B dyes decreased from 93.25 to 86.63% and 91.06 to 83.61% after five cycles, respectively, indicating that the efficiency reduction percentage is less than 10%. ZnO indicates excellent photocatalytic activity and high stability after five cycles [18]. In addition, Ikram et al. (2021) studied the reusability of Mg-ZnO nanorods for eliminating a mixture of MB and ciprofloxacin. After four cycles, the photodegradation efficiency slightly reduced from 82 to 75% after 40 min [122].
Also, Padikkaparambil and coworkers (2013) studied the reusability of a 2%Au-TiO2 nanocatalyst for MO photodegradation under UV irradiation in 11 reuse cycles. Under operating conditions (3 g/L catalyst, 10 mg/L MO, and 1 h irradiation time), the photocatalytic activity of Au/TiO2 did not change (about 100% photodegradation efficiency) even after 10 consecutive tests, which indicates the high reusability of 2%Au-TiO2 [123]. Also, Wu and Zhang (2019) studied the reusability and stability of a Samarium/Nitrogen co-TiO2/diatomite for eliminating tetracycline under visible light. After five cycles, the photodegradation efficiency reduced from 87.1% to 83.2% (less than 5% reduction), indicating high stability and significant reusability of the Samarium/Nitrogen co-doped TiO2/diatomite catalyst [124]. Moreover, Song et al. studied the reusability of C-TiO2/carbon nanofibers in removing Rhodamine B from aqueous solutions. The nanocatalyst showed a photodegradation efficiency of 94.2% under UV light. After six reuse cycles, the photodegradation efficiency reached 92%, showing significant reusability and stability of C-TiO2/carbon nanofibers [125]. In another work, Bahrudin and coworkers (2019) studied the reusability of TiO2 and TiO2/polyaniline for MO dye photodegradation. After ten reuse cycles, the photodegradation efficiency of MO using TiO2 and TiO2/polyaniline decreased from 58.3 to 36.1% and 86 to 46.2%, respectively [126]. Moreover, Moosavi et al. (2020) studied the recyclability and stability of Fe3O4/AC/TiO2 for seven cycles of MB photodegradation. After seven reuse cycles, the photodegradation efficiency reduced from 98 to 93% (about 5% reduction). They attributed the decrease in the photodegradation efficiency after seven cycles to the following reasons: (i) Particle losses may take place in washing and drying steps, which result in lower doses in the next cycle, leading to the decrease in the surface catalytic activity as well as the efficiency. (ii) Characteristics of the catalyst like aggregation may change during different cycles because the aggregation of particles reduces the specific surface area and the number of active sites. (iii) The catalytic activity of the catalyst dwindles after each cycle due to the blockage of active sites and pores [127]. Furthermore, the reusability of the 2-(methacryloyloxy) ethyltrimethylammonium chloride/TiO2 photocatalyst was studied for MB removal under UV irradiation. The photodegradation efficiency of the aforementioned photocatalyst after 270 min was 99.66% and after 20 reuse cycles, the photodegradation efficiency reached 98.7%, i.e., a reduction of less than 1% in the photodegradation efficiency, indicating remarkable stability of the photocatalyst [128]. In addition, a TiO2/2NiO photocatalyst was used to decompose MB in five reuse cycles and the outcomes showed that the photodegradation efficiency of MB using TiO2/2NiO reduced from 100% to 72.6% after five cycles, indicating its high stability [115].
In general, examining the reusability of ZnO and TiO2 nanocatalysts demonstrates that they have high reusability and their utilization in industrial applications is cost-effective. Also, adding metallic and non-metallic cocatalysts to the structure of TiO2 and ZnO catalysts increased their stability.

7. Literature Review

So far, many works have been conducted on the removal of organic compounds by catalysts. Most of these catalysts are based on TiO2 or ZnO because these catalysts have shown high performance compared to other catalysts. These catalysts can degrade organic contaminants in the presence of light or UV. Table 7 reports the performance of TiO2 and ZnO nanocatalysts in the photodegradation of organic contaminants.
According to Table 7, Mansoori and coworkers studied the photocatalytic efficiency of TiO2 in the removal of toluene, phenol, nitrobenzene, and MO. According to their results, the photodegradation efficiency of phenol, nitrobenzene, and MO was 100%, which is a remarkable amount. In addition, the degradation efficiency of toluene by the TiO2 photocatalyst was 71% [130]. Also, the Bi12TiO20 nanophotocatalyst could remove cefixime with a degradation efficiency of 94.93% after 3 h, while Bi12ZnO20 removed 80% of cefuroxime from water after 4 h [131]. In another study, Zhang and coworkers studied the performance of TiO2 in the removal of parathion. Under optimal conditions (i.e., contaminant concentration of 50 ppm and catalyst dosage of 1 g/L), the maximum photocatalytic efficiency was 70% [132]. Also, Razip and coworkers (2019) synthesized an Fe3O4/TiO2 nanocatalyst and used it in the removal of MO from wastewater. After 1h of UV irradiation, the photodegradation efficiency of MO dye was obtained at 90.3% [22]. In addition, Hou and coworkers synthesized a PVP/TiO2/polydopamine nanocatalyst, and then used it to remove malachite green, MB, and MO dyes from wastewater. Based on their results, the photodegradation efficiency of malachite green, MB, and MO dyes using the nanocatalyst were 45, 25, and 24%, respectively, which are low values. Their results reveal that the PVP/TiO2/polydopamine nanocatalyst is not suitable for the photocatalytic process [133]. Also, carbon-TiO2 nanocatalysts can eliminate methyl ethyl ketone under UV light with a photodegradation efficiency of 94% [23]. Moreover, Akhlaghian and Najafi (2018) synthesized a CuO/WO3/TiO2 nanocatalyst to remove 4-chlorophen. According to their outcomes, the aforementioned nanocatalyst was able to remove the pollutant with a degradation efficiency of 94.8%, which was obtained at a catalyst dosage of 0.75 g/L and H2O2 amount of 563.16 mmol/L, after 3 h [21]. In another study, Nagaraju et al. (2020) investigated the photocatalytic degradation of chlorobenzene from water using several catalysts such as ZnO, Pb/ZnO, Cd/ZnO, and Ag/ZnO in the presence of LED light and tungsten light. Among these photocatalysts, Pb/ZnO could decompose chlorobenzene with 100% degradation efficiency in the presence of both lights. Also, pure ZnO showed the minimum degradation efficiency (71%) in the presence of LED light [28]. Furthermore, Elmolla and Chaudhuri (2010) utilized a ZnO nanophotocatalyst for removing amoxicillin, ampicillin, and cloxacillin and their outcomes demonstrated that ZnO could eliminate all contaminants with a significant photodegradation percentage of 100% under ultraviolet after 180 min [134]. Moreover, Boytsova and coworkers synthesized TiO2 at different annealing temperatures (400–1200 °C), and then studied their photodegradation rates in the removal of crystal violet dye. According to their results, the highest photodegradation rate was obtained for the synthesized TiO2 nanocatalyst at 800 °C [135]. In addition, Chiang and Lin studied the photocatalytic performance of ZnO/SnO2 to remove methylene blue from aqueous media. Under optimal conditions, i.e., pH 12, a photocatalyst dosage of 0.5 g/L, and a reaction time of 60 min, the highest photodegradation efficiency was 96%, which is a considerable amount [136]. In another study, Ce/ZnO showed significant photocatalytic efficiency (99.5%) in the removal of direct red-23 dye after 70 min [137]. Also, Al-Namshah and coworkers compared the photocatalytic performance of ZnO, Ce/ZnO, and CeO2/ZnO in the removal of methyl green dye. Based on their outcomes, the photodegradation efficiencies of methyl green using ZnO, Ce/ZnO, and CeO2/ZnO were 68, 98, and 100%, which demonstrates that CeO2/ZnO has a remarkable photocatalytic efficiency [138].
Overall, TiO2 in the removal of nitrobenzene, phenol, and MO, Bi12TiO20 in the removal of cefixime, Fe3O4/TiO2 (P25) in the removal of MO dye, CuO/WO3/TiO2 in the removal of 4-Chlorophen, carbon/TiO2 in the removal of methyl ethyl ketone, as well as, ZnO in the removal of amoxicillin, ampicillin, and cloxacillin, Fe3O4/CuO/ZnO/graphene and ZnO/SnO2 in the removal of MB, Ce/ZnO in the removal of direct red-23 dye, Pb/ZnO and Ag/ZnO in the removal of chlorobenzene, and Ce/ZnO and CeO2/ZnO in the removal of methyl green, due to significant photodegradation efficiencies (more than 90%), are strongly suggested as powerful catalysts.
Table 7. The performance of TiO2 and ZnO nanocatalysts and their derivatives in degrading different organic pollutants.
Table 7. The performance of TiO2 and ZnO nanocatalysts and their derivatives in degrading different organic pollutants.
NanocatalystsContaminantsOptimal ConditionsDE (%)References
TiO2NitrobenzeneCD = 0.1 M, CC = 50 ppm100[130]
Bi12TiO20Cefixime3 h94.93[131]
TiO2ParathionCD = 1 g/L, CC = 50 ppm70[132]
TiO2TolueneCD = 5 g, CC = 45 ppm71[130]
TiO2Phenol1.8 g/L catalyst dose100[130]
TiO2BenzeneCD = 5 g, CC = 45 ppm72[133]
TiO2MOCD = 3 g/L, CC = 30 ppm100[130]
Fe3O4/TiO2 (P25)MO1 h under UV light irradiation90.3[22]
Fe3O4/TiO2 (UV100)MO1 h under UV light irradiation51.6[22]
CuO/WO3/TiO24-ChlorophenolCD = 0.75 g/L, H2O2 amount = 563.16 mmol/L, 3 h94.8[21]
CuO/WO3/TiO23-Phenyl-1-propanolCD = 0.75 g/L, H2O2 amount = 563.16 mmol/L, 3 h85.13[21]
Carbon- TiO2Methyl ethyl ketoneUnder UV light94[23]
La/TiO2Ramazol
Brilliant blue
-72[22]
ZnOAmoxicillinUltraviolet, pH = 11, catalyst dose = 0.5 g/L, time 180 min100[134]
ZnOAmpicillinUltraviolet, pH = 11, catalyst dose = 0.5 g/L, time 180 min100[134]
ZnOCloxacillinUltraviolet, pH = 11, catalyst dose = 0.5 g/L, time 180 min100[134]
ZnO/SnO2MBpH 12, CD = 0.5 g/L, time 60 min96[136]
Ce/ZnODirect red-23Reaction time = 70 min99.5[137]
ZnOMethyl greenCC = 20 ppm, CD = 2 g/L, time = 60 min68[138]
Ce/ZnOMethyl greenCC = 20 ppm, CD = 2 g/L, time = 60 min98[138]
CeO2/ZnOMethyl greenCC = 20 ppm, CD = 2 g/L, time = 60 min100[138]
PVP/TiO2/polydopamineMalachite green60 min, 10 mg/L of dye45[139]
PVP/TiO2/polydopamineMB60 min, 10 mg/L of dye25[139]
PVP/TiO2/polydopamineMO60 min, 10 mg/L of dye24[139]
rGO/Fe3O4/ZnOMV120 min, CD = 0.04 g/L83.5[25]
Tungsten/silver/ZnOPonceau 4RpH 5.64, CD = 0.08 g/L, 25 °C78.8[26]
Bi12ZnO20Cefuroxime4 h80[28]
ZnBi2O4CefiximeSolar light (98 mW/cm2), 30 min89[140]
ZnBi2O4CefiximeUV irradiation (20 mW/cm2), 2 h88[140]
Ag/TiO2PhenolpH 7, CD = 1.5 g/L, CC = 5 ppm, power light = 18 W82.65[141]
Ce/TiO2Sulfur blackpH = 9.5, CC = 200 ppm92[142]
CeO2/ZnO/TiO2Rhodamine BpH 12, CC = 5 ppm, CD = 0.2 g/L, time = 180 min80[143]
Pure ZnOChlorobenzeneLED light71[27]
Pb/ZnOChlorobenzeneLED light100[27]
Ag/ZnOChlorobenzeneLED light95[27]
Cd/ZnOChlorobenzeneLED light90[27]
Pure ZnOChlorobenzeneTungsten light90[27]
Pb/ZnOChlorobenzeneTungsten light100[27]
Ag/ZnOChlorobenzeneTungsten light83[27]
Cd/ZnOChlorobenzeneTungsten light73[27]
CC = contaminant concentration; CD = catalyst dose.

8. Factors Affecting Photodegradation Efficiency

8.1. Photocatalyst Dosage

Catalyst dosage or catalyst loading is a critical factor in the photocatalytic efficiency of organic contaminants. Increasing the catalyst dose leads to an enhancement in the specific surface area of the catalyst, the creation of more active sites, and ultimately the formation of more hydroxyl and superoxide radicals. Thus, the degradation of organic contaminants will be facilitated and a higher photodegradation efficiency will be attained. The photodegradation efficiency initially enhances with enhancing the photocatalyst dosage until it reaches an optimal value. At the photocatalyst concentration beyond the optimal concentration, the photodegradation efficiency decreases because of light scattering. Increasing the photocatalyst dosage beyond the optimal value leads to the agglomeration of catalyst particles and then decreases the specific surface area of the catalyst to absorb light, which ultimately leads to a reduction in the degradation efficiency. On the other hand, it prevents the penetration of light into the sewage. As a critical result, an optimal value of the catalyst must be determined to prevent overuse of the catalyst and to achieve maximum degradation efficiency [40].
In one study, Zhang and co-workers (2015) studied the photodegradation efficiency of MB using TiO2 in the range of photocatalyst dosage from 1 to 10%. According to their outcomes, the maximum photodegradation efficiency of MB (93.78%) was attained at an optimal value of 7% [144]. Also, Notodarmojo et al. (2017) studied the degradation efficiency of RB-5 dye from water and found that the degradation efficiency increases with increasing the TiO2 photocatalyst concentration from 0.5 to 2.5 g/L. Therefore, 2.5 g/L was considered the optimum photocatalyst dosage [145]. Moreover, Saiful Amran and co-workers (2019) investigated the maximum degradation efficiency of MB using carbon-TiO2 in the photocatalyst dosage range of 1-3%. According to their outcomes, the highest photodegradation efficiency (82.67%) was obtained at 2 wt.% photocatalyst dosage [146]. Furthermore, Erdemoğlu et al. studied the influence of nano-TiO2 dosage (0.1–1 wt.%) on photodegradation of CR under visible light. The nanocatalyst could remove 94% of CR at a catalyst dosage of 0.25 wt.%. By enhancing catalyst dosage from 0.25 to 1 wt.%, the photodegradation efficiency decreased [147].

8.2. Photocatalyst Structure

Photocatalytic efficiency is effectively influenced by the photocatalytic structure. The synthesis of nanostructured photocatalysts has recently received much attention because of their unique structures and features. There are various morphologies for nanostructured photocatalysts, including nanoflowers, nanosheets, nanowires, nanorods, nanodumbbells, nanobelts, and nanospiral disks. Some structures of these photocatalysts are illustrated in Figure 10. Nanostructured photocatalysts have a high specific active area with great catalytic strength. A high surface/volume proportion presents better physical and chemical features. Each of these structures has shown various photocatalytic features [148,149].

8.3. Contaminant Concentration on Photodegradation Efficiency

Contaminant concentration has an important influence on the photocatalytic process. Previous studies show that the concentration of organic pollutants in aqueous solution has an undesirable impact on photodegradation efficiency. The higher the concentration of contaminants in the effluent, the lower the photodegradation efficiency, which is because the concentration of the target contaminant becomes more and more. Organic contaminants can be adsorbed on the photocatalyst surface. The number of contaminant molecules increases, while the number of active sites remains constant [154]. Hence, the generation of hydroxyl radicals is not enough and there will be only a few active sites on the photocatalyst surface to absorb hydroxyl ions. Consumption of hydroxyl radicals (OH.) by the produced intermediates decreases the photodegradation efficiency in solutions with high contaminant concentrations. Therefore, the lower the pollutant concentration, the less competition there is for consumption [155,156]. Parida and Parija (2006) investigated the impact of phenol concentration on the degradation efficiency using ZnO in various irradiation strengths. Under solar irradiation, the photodegradation yield decreased from 100% to 60% with an enhancement in the phenol dosage, while under ultraviolet light, the photodegradation efficiency reduced from 94% to 52% with an enhancement in the phenol concentration [157]. Also, Benhabiles et al. (2016) studied the influence of MB dye concentration (10–30 mg/L) on the photocatalytic process by commercial TiO2. After 5 h, 70% of MB dye was degraded by TiO2 at MB concentration of 10 mg/L, while at MB concentration of 30 mg/L, the catalyst could remove only 30% of MB [158]. Moreover, Shelar et al. (2020) surveyed the impact of MB dye concentration (10–40 mg/L) on photodegradation using Ag/ZnO. The photodegradation efficiency reduced from 95% to 65% with enhancing MB dosage from 10 to 40 mg/L, indicating that the highest photodegradation efficiency occurs at the lowest dosage of MB dye [120].
Generally, previous studies demonstrate that the highest photodegradation efficiency of contaminants occurs at the lowest contaminant concentration.

8.4. pH

Solution pH plays a vital role in the photocatalytic reaction of water purification. pH can change the surface charge of the photocatalyst [159]. Determination of optimal pH in the photodegradation process depends on the zero-point charge (pHzpc) of the photocatalyst. pH and the photodegradation rate do not have a specific relation. The TiO2 photocatalyst will be negatively charged if pH > pHZPC. In this case, the TiO anion will be formed. Also, if pH < pHZPC, the TiO2 photocatalyst will be positively charged and form the TiOH+2 cation. Depending on the catalyst used, the surface charge of the photocatalyst at 4.5 < pHZPC < 7 is neutral. When operating at pHzpc, the surface charge of the TiO2 photocatalyst is positively charged and absorbs negatively charged molecules electrostatically over time. Industrial effluents may be discharged at different pHs, which complicates the photocatalytic reaction. Also, hydroxyl radicals are rapidly eliminated at high pHs, inhibiting their reaction with the contaminant [40,156]. There are many effective factors on pH in the photocatalytic reaction, including the electrostatic charge of catalyst particles, band structure, and crystal size. Each of these factors can alter the photocatalyst surface charge and generally alter the photodegradation efficiency [156]. Generally, the optimal pH in photodegradation depends on the contaminant and the photocatalyst. For example, Hameed et al. (2009) studied the impact of pH on the photodegradation efficiency of carbofuran using TiO2/Ultraviolet and pH 7 was obtained as the optimal pH [160]. Also, Lopez-Alvarez et al. (2011) surveyed the impact of pH on the photodegradation efficiency of carbofuran using TiO2/solar light and their outcomes showed that pH 7.6 is the optimal value [161]. Moreover, in the work conducted by Saien and Khezrianjoo (2008) for the degradation of carbendazim using TiO2/ultraviolet, the optimal pH was 4 [162]. Furthermore, Elmolla and Chaudhuri (2010) found an optimal pH of 11 for the photodegradation of amoxicillin, ampicillin, and cloxacillin using ZnO/ultraviolet [135]. In another work, the highest photodegradation efficiency of TiO2 for MB removal was attained at a pH of 11 [158]. Also, Shelar et al. studied the impact of pH on photodegrading MB dye in the range of 2–12 and found that the utmost photodegradation efficiency is obtained at pH 8 [120].

8.5. Light Intensity and Wavelength

Photodegradation efficiency depends on the absorption of light by a catalyst [163]. Light intensity specifies how much light can be adsorbed by a photocatalyst at a given wavelength. The higher the light intensity, the more radiation is deposited on the photocatalyst surface, leading to the production of more hydroxyl radicals and an enhancement of the reaction rate. At higher light intensities, the photodegradation reaction depends on the mass transfer between the reactants, because the photocatalyst surface is completely covered by saturated solids, limiting the mass transfer for sorption and desorption. Therefore, the photocatalytic reaction speed remains constant in spite of an increase in the intensity of light. Previous studies have shown that there is a relationship between the photodegradation efficiency of the contaminant and light intensity for different organic compounds. The generation of hydroxyl radicals increases with increasing light intensity, which leads to an improvement in the degradation rate [156,164]. Elaziouti et al. investigated the impact of light intensity in the range of 50–90 j/cm2 on the photodegradation efficiency of CR and benzopurpurine 4B. The photodegradation efficiency of CR and benzopurpurine 4B increased with enhancing the light intensity from 50 to 80 j/cm2 and 50 to 90 j/cm2, respectively. Because more photons at higher light intensities will be available for excitation on the catalyst surface, more electron–hole pairs can be produced [165].
Based on the wavelength of UV irradiation, there are three electromagnetic spectra, including UV-A, UV-B, and UV-C. These are categorized based on the wavelength range. For example, the light wavelength ranges for UV-C, UV-B, and UV-A are between 100 and 280, 280 and 315, and 314 and 400 nm, respectively [166]. Because of the shorter penetration of higher energy photons, the photodegradation rate is greater at 254 nm, which increases the number of pairs of electron holes produced to decompose the target contaminant [167].

8.6. Temperature

It is better to perform the photodegradation reaction at 25 °C and 1 bar because of photonic activation, which is useful to purify water, by which the heating stage can be eliminated to save energy. However, the optimum temperature for photocatalytic reactions can occur at 20–80 °C. The results show that the photodegradation efficiency of organic pollutants increases with enhancing the reaction temperature; however, it may lead to a decline in the adsorption capacity of reactive species and dissolved oxygen, leading to lower photodegradation efficiency [168]. Chen and coworkers studied the impact of temperature (0–50 °C) on the photocatalytic activity of TiO2 and Pd/TiO2 photocatalysts for MB elimination under UV light, and their outcomes demonstrated that the photodegradation efficiency increases with increasing temperature [169]. Also, Hu and coworkers investigated the impact of temperature on the photodegradation of MO dye using TiO2. According to their results, under UV-vis irradiation, the photocatalytic activity of TiO2 was increased by enhancing temperature from 38 to 100 °C. Hence, the reaction was endothermic. Also, the reaction rate constant increased from 0.00031 to 0.00217 min−1 with enhancing temperature from 34 to 100 °C [170]. Moreover, Barakat and coworkers studied the influence of temperature (5–55 °C) on the degradation of Rhodamine B using Ag-TiO2. The highest photodegradation efficiency under irradiation occurred at 55 °C [171]. Generally, the photodegradation efficiency of organic contaminants increases with enhancing temperature and the highest removal performance occurs at high temperatures.

8.7. Reaction Time

Irradiation time or reaction time is a critical factor in the photodegradation process and is when the contaminant is exposed to light or photon energy. Many researchers have investigated the impact of reaction time on the photodegradation efficiency of organic contaminants using TiO2 and ZnO photocatalysts. The longer the reaction time, the higher the photodegradation efficiency of the contaminant [40]. According to the work conducted by Shi et al. (2019), eliminating MO using TiO2/chitosan increased with enhancing the reaction time and finally reached equilibrium. At the beginning of the photocatalytic reaction, there is a large number of active sites at the photocatalyst surface to attach with MO molecules, resulting in a high degradation efficiency. After the equilibrium reaction time, no change in the photodegradation efficiency was observed [172]. Elaziouti et al. studied the impact of reaction time (0–80 min) on the photodegradation efficiency of CR and benzopurpurine 4B. A total of 95% of CR dye and 97.2% of benzopurpurine 4B were degraded using 1 g/L ZnO after 60 and 80 min, respectively [165]. Also, Erdemoğlu et al. studied the influence of irradiation time on the photodegradation of CR dye using TiO2. Under visible irradiation, CR was completely decomposed after 30 min [147].

9. Conclusions and Future Perspectives

TiO2 and ZnO photocatalysts under either UV light or solar irradiation were studied for the degradation of organic compounds from wastewater. These catalysts have outstanding features like low cost, excellent degradation efficiency, high photocatalytic activity, etc., and are known as important catalysts in the water and wastewater industry. Both catalysts have been shown to be usable in the presence of sunlight and UV light, but the results generally indicate that using UV light is more efficient. Previous studies have shown that a mix of these catalysts with other materials can enhance the degradation efficiency of organic compounds. In this review study, the impact of various factors like temperature, time, photocatalyst loading, photocatalyst shape, pH, light wavelength, and light intensity was investigated on degrading organic compounds and the outcomes demonstrated that pH, photocatalyst dosage, temperature, and light intensity have significant impacts on the photodegradation efficiency. Also, the impact of these photocatalysts on the degradation rate of various organic compounds from wastewater was fully surveyed. Moreover, various cocatalysts can be utilized in the structure of TiO2 and ZnO nanocatalysts, so the impact of adding metal and non-metal cocatalysts on their structures was fully studied. Studies show that the addition of some metal cocatalysts such as Ag, Pd, and Co and non-metal cocatalysts such as C and N in the structure of ZnO and TiO2 significantly increases the photodegradation performance of organic contaminants. Furthermore, the reusability of ZnO and TiO2 catalysts showed that they have high stability and significant reusability. Therefore, they can be used in several cycles without a significant decrease in their photocatalytic activity. The 2%Au-TiO2 nanocatalyst with a photodegradation efficiency of 100% after 11 reuse cycles showed outstanding stability for the elimination of methyl orange dye. Also, the photodegradation efficiency of 2-(methacryloyloxy) ethyltrimethylammonium chloride/TiO2 photocatalyst reduced from 99.66% to 98.7% after 20 reuse cycles for the elimination of MB dye, indicating remarkable stability.
Generally, previous studies have shown that they can remove more than 90% of most organic contaminants. Therefore, both metals and non-metals anchored to TiO2 and ZnO are strongly recommended for the elimination of organic compounds from municipal and industrial wastewaters.

Author Contributions

Conceptualization, H.A. and N.F.; methodology, S.A.A. and M.A.T.; software, S.A.A. and A.A.; validation, N.F., A.K.S. and L.C.-C.; formal analysis, A.K.S. and C.R.-B.; investigation, A.T.-A.; resources, H.A.; data curation, A.T.-A. and H.E.; writing—original draft preparation, A.A.; writing—review and editing, H.E.; visualization, H.E.; supervision, A.K.S.; project administration, L.C.-C.; funding acquisition, A.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Deanship of Scientific Research at King Khalid University under grant number RGP.2/36/45.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The authors extend their appreciation to the Deanship of Scientific Research at King Khalid University for funding this work through the Large Group Project under grant number RGP.2/36/45.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Tamjidi, S.; Moghadas, B.K.; Esmaeili, H.; Khoo, F.S.; Gholami, G.; Ghasemi, M. Improving the surface properties of adsorbents by surfactants and their role in the removal of toxic metals from wastewater: A review study. Process Saf. Environ. Prot. 2021, 148, 775–795. [Google Scholar] [CrossRef]
  2. Yang, Y. Dynamic Relationship of Urban and Rural Water Shortage Risks Based on the Economy–Society–Environment Perspective. Agriculture 2022, 12, 148. [Google Scholar] [CrossRef]
  3. Liang, H.; Esmaeili, H. Application of nanomaterials for demulsification of oily wastewater: A review study. Environ. Technol. Innov. 2021, 22, 101498. [Google Scholar] [CrossRef]
  4. Khaleghi, H.; Esmaeili, H.; Jaafarzadeh, N.B. Ramavandi, Date seed activated carbon decorated with CaO and Fe3O4 nanoparticles as a reusable sorbent for removal of formaldehyde. Korean J. Chem. Eng. 2022, 39, 146–160. [Google Scholar] [CrossRef]
  5. Esmaeili, H.; Esmaeilzadeh, F.; Mowla, D. Effect of surfactant on stability and size distribution of gas condensate droplets in water. J. Chem. Eng. Data 2014, 59, 1461–1467. [Google Scholar] [CrossRef]
  6. Kuhad, R.C.; Gupta, R. Biological remediation of petroleum contaminants. In Advances in Applied Bioremediation; Springer: Berlin/Heidelberg, Germany, 2009; pp. 173–187. [Google Scholar]
  7. Lürling, M.; Kang, L.; Mucci, M.; van Oosterhout, F.; Noyma, N.P.; Miranda, M.; Huszar, V.L.; Waajen, G.; Marinho, M.M. Coagulation and precipitation of cyanobacterial blooms. Ecol. Eng. 2020, 158, 106032. [Google Scholar] [CrossRef]
  8. Kumar, V.; Othman, N.; Asharuddin, S. Applications of natural coagulants to treat wastewater—A review. MATEC Web Conf. EDP Sci. 2017, 103, 06016. [Google Scholar] [CrossRef]
  9. Sugita, T. Study of Water Treatment Using Photocatalytic Material. Ph.D. Thesis, Gunma University, Maebashi, Japan, 2015. [Google Scholar]
  10. Karale, R.S.; Wadkar, D.V.; Wagh, M.P. Effect of hydrogen peroxide and ferrous ion on the degradation of 2-Aminopyridine. J. Appl. Water Eng. Res. 2023, 11, 54–65. [Google Scholar] [CrossRef]
  11. Xu, M.; Yu, F.; Liu, Y.; Li, W.; Li, C.; Song, F.; Wu, G.; Xu, Z.; Qiu, J. Integrating Uranyl-Affinity “Hooks” into Conjugated Polymers Achieving Giant Built-in Electric Field for Boosting Photocatalytic Uranium Extraction from Seawater. Macromolecules 2024, 57, 5679–5690. [Google Scholar] [CrossRef]
  12. Cuerda-Correa, E.M.; Alexandre-Franco, M.F.; Fernández-González, C. Advanced oxidation processes for the removal of antibiotics from water. An overview. Water 2020, 12, 102. [Google Scholar] [CrossRef]
  13. Ren, X.; Philo, D.; Li, Y.; Shi, L.; Chang, K.; Ye, J. Recent advances of low-dimensional phosphorus-based nanomaterials for solar-driven photocatalytic reactions. Coord. Chem. Rev. 2020, 424, 213516. [Google Scholar] [CrossRef]
  14. Xie, Y.; Hu, J.; Esmaeili, H.; Wang, D.; Zhou, Y. A review study on wastewater decontamination using nanotechnology: Performance, mechanism and environmental impacts. Powder Technol. 2022, 412, 118023. [Google Scholar] [CrossRef]
  15. Shirsath, S.R.; Pinjari, D.V.; Gogate, P.R.; Sonawane, S.H.; Pandit, A.B. Ultrasound assisted synthesis of doped TiO2 nano-particles: Characterization and comparison of effectiveness for photocatalytic oxidation of dyestuff effluent. Ultrason. Sonochem. 2013, 20, 277–286. [Google Scholar] [CrossRef] [PubMed]
  16. Zhang, T.; Ren, X.; Mo, S.; Cao, W.; Zhou, C.; Ma, F.; Chen, R.; Zeng, C.; Shi, L.; Liu, T.; et al. Modulating Fe/P ratios in Fe-P alloy through smelting reduction for long-term electrocatalytic overall water splitting. J. Mater. Sci. Technol. 2024, 199, 66–74. [Google Scholar] [CrossRef]
  17. Hu, M.; Yao, Z.; Wang, X. Characterization techniques for graphene-based materials in catalysis. AIMS Mater. Sci. 2017, 4, 755–788. [Google Scholar] [CrossRef]
  18. Khan, M.; Ware, P.; Shimpi, N. Synthesis of ZnO nanoparticles using peels of Passiflora foetida and study of its activity as an efficient catalyst for the degradation of hazardous organic dye. SN Appl. Sci. 2021, 3, 528. [Google Scholar] [CrossRef]
  19. Bruno, M.E.; Tasat, D.R.; Ramos, E.; Paparella, M.L.; Evelson, P.; Rebagliati, R.J.; Cabrini, R.L.; Guglielmotti, M.B.; Olmedo, D.G. Impact through time of different sized titanium dioxide particles on biochemical and histopathological parameters. J. Biomed. Mater. Res. A 2014, 102, 1439–1448. [Google Scholar] [CrossRef] [PubMed]
  20. Yu, C.; Cai, D.; Yang, K.; Yu, J.C.; Zhou, Y.; Fan, C. Sol–gel derived S,I-codoped mesoporous TiO2 photocatalyst with high visible-Light photocatalytic activity. J. Phys. Chem. Solids 2010, 71, 1337–1343. [Google Scholar] [CrossRef]
  21. Akhlaghian, F.; Najafi, A. CuO/WO3/TiO2 photocatalyst for degradation of phenol wastewater. Sci. Iran. 2018, 25, 3345–3353. [Google Scholar] [CrossRef]
  22. Razip, N.I.M.; Lee, K.M.; Lai, C.W.; Ong, B.H. Recoverability of Fe3O4/TiO2 nanocatalyst in methyl orange degradation. Mater. Res. Express 2019, 6, 075517. [Google Scholar] [CrossRef]
  23. Shayegan, Z.; Haghighat, F.; Lee, C.S. Carbon-doped TiO2 film to enhance visible and UV light photocatalytic degradation of indoor environment volatile organic compounds. J. Environ. Chem. Eng. 2020, 8, 104162. [Google Scholar] [CrossRef]
  24. Sadovnikov, A.A.; Garshev, A.V.; Eliseev, A.A.; Beltiukov, A.N.; Naranov, E.R.; Li, W.; Sutherland, A.J.; Boytsova, O.V. Nanowhiskers of K2Ti6O13 as a promoter of photocatalysis in anatase mesocrystals. Catal. Today 2021, 378, 133–139. [Google Scholar] [CrossRef]
  25. Thangavel, S.; Thangavel, S.; Raghavan, N.; Krishnamoorthy, K.; Venugopal, G. Visible-light driven photocatalytic degradation of methylene-violet by rGO/Fe3O4/ZnO ternary nanohybrid structures. J. Alloys Compd. 2016, 665, 107–112. [Google Scholar] [CrossRef]
  26. Vafaee, M.; Olya, M.E.; Drean, J.Y.; Hekmati, A.H. Synthesize, characterization and application of ZnO/W/Ag as a new nanophotocatalyst for dye removal of textile wastewater; kinetic and economic studies. J. Taiwan Inst. Chem. Eng. 2017, 80, 379–390. [Google Scholar] [CrossRef]
  27. Nagaraju, P.; Puttaiah, S.H.; Wantala, K.; Shahmoradi, B. Preparation of modified ZnO nanoparticles for photocatalytic degradation of chlorobenzene. Appl. Water Sci. 2020, 10, 137. [Google Scholar] [CrossRef]
  28. Baaloudj, O.; Nasrallah, N.; Assadi, A.A. Facile synthesis, structural and optical characterizations of Bi12ZnO20 sillenite crystals: Application for Cefuroxime removal from wastewater. Mater. Lett. 2021, 304, 130658. [Google Scholar] [CrossRef]
  29. Belousov, A.S.; Parkhacheva, A.A.; Suleimanov, E.V.; Fukina, D.G.; Koryagin, A.V.; Shafiq, I.; Krasheninnikova, O.V.; Kuzmichev, V.V. Doping vs. heterojunction: A comparative study of the approaches for improving the photocatalytic activity of flower-like Bi2WO6 for water treatment with domestic LED light. Catal. Commun. 2023, 180, 106705. [Google Scholar] [CrossRef]
  30. Liu, Q.; Zhao, Z.; Li, H.; Su, M.; Liang, S.X. Occurrence and removal of organic pollutants by a combined analysis using GC-MS with spectral analysis and acute toxicity. Ecotoxicol. Environ. Saf. 2021, 207, 111237. [Google Scholar] [CrossRef]
  31. Jain, M.; Khan, S.A.; Sharma, K.; Jadhao, P.R.; Pant, K.K.; Ziora, Z.M.; Blaskovich, M.A. Current perspective of innovative strategies for bioremediation of organic pollutants from wastewater. Bioresour. Technol. 2022, 344, 126305. [Google Scholar] [CrossRef]
  32. Aguirre Hernandez, E.A. Management of Produced Water from the Oil and Gas Industry: Characterization, Treatment, Disposal and Beneficial Reuse. Doctoral Dissertation, Politecnico di Torino, Turin, Italy, 2021. [Google Scholar]
  33. David, E.; Niculescu, V.C. Volatile Organic Compounds (VOCs) as Environmental Pollutants: Occurrence and Mitigation Using Nanomaterials. Int. J. Environ. Res. Public Health 2021, 18, 13147. [Google Scholar] [CrossRef]
  34. Amaral, E.T.; Bender, L.B.Y.C.; Rizzetti, T.M.; de Souza Schneider, R.D.C. Removal of organic contaminants in water bodies or wastewater by microalgae of the genus chlorella: A review. Case Stud. Chem. Environ. Eng. 2023, 8, 100476. [Google Scholar] [CrossRef]
  35. Ramirez Castillo, F.Y.; Avelar González, F.J.; Garneau, P.; Marquez Diaz, F.; Guerrero Barrera, A.L.; Harel, J. Presence of multi-drug resistant pathogenic Escherichia coli in the San Pedro River located in the State of Aguascalientes, Mexico. Front. Microbiol. 2013, 4, 147. [Google Scholar] [CrossRef] [PubMed]
  36. Ramírez-Franco, J.H.; Galeano, L.-A.; Vicente, M.-A. Fly ash as photo-Fenton catalyst for the degradation of amoxicillin. J. Environ. Chem. Eng. 2019, 7, 103274. [Google Scholar] [CrossRef]
  37. Tavasol, F.; Tabatabaie, T.; Ramavandi, B.; Amiri, F. Design a new photocatalyst of sea sediment/titanate to remove cephalexin antibiotic from aqueous media in the presence of sonication/ultraviolet/hydrogen peroxide: Pathway and mechanism for degradation. Ultrason. Sonochem. 2020, 65, 105062. [Google Scholar] [CrossRef] [PubMed]
  38. Shojaei, M.; Esmaeili, H. Ultrasonic-assisted synthesis of zeolite/activated carbon@ MnO2 composite as a novel adsorbent for treatment of wastewater containing methylene blue and brilliant blue. Environ. Monit. Assess. 2022, 194, 279. [Google Scholar] [CrossRef] [PubMed]
  39. Tan, I.A.W.; Ahmad, A.L.; Hameed, B.H. Enhancement of basic dye adsorption uptake from aqueous solutions using chemically modified oil palm shell activated carbon. Colloids Surf. A Physicochem. Eng. Asp. 2008, 318, 88–96. [Google Scholar] [CrossRef]
  40. Yang, L.; Li, Z.; Wang, X.; Li, L.; Chen, Z. Facile electrospinning synthesis of S-scheme heterojunction CoTiO3/g-C3N4 nanofiber with enhanced visible light photocatalytic activity. Chin. J. Catal. 2024, 59, 237–249. [Google Scholar] [CrossRef]
  41. Hui, K.C.; Suhaimi, H.; Sambudi, N.S. Electrospun-based TiO2 nanofibers for organic pollutant photodegradation: A comprehensive review. Rev. Chem. Eng. 2022, 38, 641–668. [Google Scholar] [CrossRef]
  42. Laskar, N.; Kumar, U. Adsorption of crystal violet from wastewater by modified bambusa tulda. KSCE J. Civ. Eng. 2018, 22, 2755–2763. [Google Scholar] [CrossRef]
  43. Wanyonyi, W.C.; Onyari, J.M.; Shiundu, P.M. Adsorption of Congo red dye from aqueous solutions using roots of Eichhornia crassipes: Kinetic and equilibrium studies. Energy Procedia 2014, 50, 862–869. [Google Scholar] [CrossRef]
  44. Wang, Z.; Yang, P.; He, X.; Yu, Q. Preparation of intercalated MXene by TPAOH and its adsorption characteristics towards U (VI). J. Radioanal. Nucl. Chem. 2024, 333, 1999–2014. [Google Scholar] [CrossRef]
  45. Bai, B.; Bai, F.; Li, X.; Nie, Q.; Jia, X.; Wu, H. The remediation efficiency of heavy metal pollutants in water by industrial red mud particle waste. Environ. Technol. Innov. 2022, 28, 102944. [Google Scholar] [CrossRef]
  46. Shen, Y.; Sun, P.; Ye, L.; Xu, D. Progress of anaerobic membrane bioreactor in municipal wastewater treatment. Sci. Adv. Mater. 2023, 15, 1277–1298. [Google Scholar] [CrossRef]
  47. Lebedev, V.A.; Kozlov, D.A.; Kolesnik, I.V.; Poluboyarinov, A.S.; Becerikli, A.E.; Grünert, W.; Garshev, A.V. The amorphous phase in titania and its influence on photocatalytic properties. Appl. Catal. B Environ. 2016, 195, 39–47. [Google Scholar] [CrossRef]
  48. Bakhtiari, R.; Kamkari, B.; Afrand, M.; Abdollahi, A. Preparation of stable TiO2-Graphene/Water hybrid nanofluids and development of a new correlation for thermal conductivity. Powder Technol. 2021, 385, 466–477. [Google Scholar] [CrossRef]
  49. Gatto, S. Photocatalytic Activity Assessment of Micro-Sized TiO2 Used as Powders and as Starting Material for Porcelain Gres Tiles Production. Doctoral Thesis, Università degli Studi di Milano, Milan, Italy, 2014. [Google Scholar]
  50. Sarafraz, M.; Sadeghi, M.; Yazdanbakhsh, A.; Amini, M.M.; Sadani, M.; Eslami, A. Enhanced photocatalytic degradation of ciprofloxacin by black Ti3+/N-TiO2 under visible LED light irradiation: Kinetic, energy consumption, degradation pathway, and toxicity assessment. Process Saf. Environ. Prot. 2020, 137, 261–272. [Google Scholar] [CrossRef]
  51. Inamuddin, I.; Asiri, A.M.; Lichtfouse, E. Nanophotocatalysis and Environmental Applications. Detoxification and Disinfection; Springer: Berlin/Heidelberg, Germany, 2020; Volume 30. [Google Scholar]
  52. Lin, C.H.; Chen, W.H. Graphene Family Nanomaterials (GFN)-TiO2 for the Photocatalytic Removal of Water and Air Pollutants: Synthesis, Characterization, and Applications. Nanomaterials 2021, 11, 3195. [Google Scholar] [CrossRef]
  53. Neppolian, B.; Wang, Q.; Jung, H.; Choi, H. Ultrasonic-assisted sol-gel method of preparation of TiO2 nano-particles: Characterization, properties and 4-chlorophenol removal application. Ultrason. Sonochem. 2008, 15, 649–658. [Google Scholar] [CrossRef]
  54. Nyamukamba, P.; Okoh, O.; Mungondori, H.; Taziwa, R.; Zinya, S. Synthetic methods for titanium dioxide nanoparticles: A review. Titan. Dioxide-Mater. A Sustain. Environ. 2018, 8, 151–1755. [Google Scholar]
  55. MODAN, E.M.; PLĂIAȘU, A.G. Advantages and disadvantages of chemical methods in the elaboration of nanomaterials. Ann. “Dunarea De Jos” Univ. Galati. Fascicle IX Metall. Mater. Sci. 2020, 43, 53–60. [Google Scholar] [CrossRef]
  56. Madjene, F.; Aoudjit, L.; Igoud, S.; Lebik, H.; Boutra, B. A review: Titanium dioxide photocatalysis for water treatment. Transnatl. J. Sci. Technol. 2013, 3, 1857–8047. [Google Scholar]
  57. Otieno, S.; Lanterna, A.E.; Mack, J.; Derese, S.; Amuhaya, E.K.; Nyokong, T.; Scaiano, J.C. Solar Driven Photocatalytic Activity of Porphyrin Sensitized TiO2: Experimental and Computational Studies. Molecules 2021, 26, 3131. [Google Scholar] [CrossRef]
  58. Lee, S.Y.; Park, S.J. TiO2 photocatalyst for water treatment applications. J. Ind. Eng. Chem. 2013, 19, 1761–1769. [Google Scholar] [CrossRef]
  59. Stamate, M.; Lazar, G. Application of titanium dioxide photocatalysis to create self-cleaning materials. Rom. Tech. Sci. Acad. MOCM 2007, 13, 280–285. [Google Scholar]
  60. Fernandes, A.; Gągol, M.; Makoś, P.; Khan, J.A.; Boczkaj, G. Integrated photocatalytic advanced oxidation system (TiO2/UV/O3/H2O2) for degradation of volatile organic compounds. Sep. Purif. Technol. 2019, 224, 1–14. [Google Scholar] [CrossRef]
  61. Xu, X.; Feng, X.; Wang, W.; Song, K.; Ma, D.; Zhou, Y.; Shi, J.W. Construction of II-type and Z-scheme binding structure in P-doped graphitic carbon nitride loaded with ZnO and ZnTCPP boosting photocatalytic hydrogen evolution. J. Colloid Interface Sci. 2023, 651, 669–677. [Google Scholar] [CrossRef] [PubMed]
  62. Thomas, J.; Yoon, M. Facile synthesis of pure TiO2 (B) nanofibers doped with gold nanoparticles and solar photocatalytic activities. Appl. Catal. B Environ. 2012, 111, 502–508. [Google Scholar] [CrossRef]
  63. Nguyen, V.H.; Tran, Q.B.; Nguyen, X.C.; Ho, T.T.T.; Shokouhimehr, M.; Vo, D.V.N.; Lam, S.S.; Nguyen, H.P.; Hoang, C.T.; Ly, Q.V.; et al. Submerged photocatalytic membrane reactor with suspended and immobilized N-doped TiO2 under visible irradiation for diclofenac removal from wastewater. Process Saf. Environ. Prot. 2020, 142, 229–237. [Google Scholar] [CrossRef]
  64. Nakayama, N.; Hayashi, T. Preparation of TiO2 nanoparticles surface-modified by both carboxylic acid and amine: Dispersibility and stabilization in organic solvents. Colloids Surf. A Physicochem. Eng. Asp. 2008, 317, 543–550. [Google Scholar] [CrossRef]
  65. Nuraje, N.; Asmatulu, R.; Mul, G. Green Photo-Active Nanomaterials: Sustainable Energy and Environmental Remediation; Royal Society of Chemistry: London, UK, 2015. [Google Scholar] [CrossRef]
  66. Amir, M.N.I.; Julkapli, N.M.; Bagheri, S.; Yousefi, A.T. TiO2 hybrid photocatalytic systems: Impact of adsorption and photocatalytic performance. Rev. Inorg. Chem. 2015, 35, 151–178. [Google Scholar] [CrossRef]
  67. Arul, A.R.; Manjulavalli, T.E.; Venckatesh, R. Visible light proven Si doped TiO2 nanocatalyst for the photodegradation of Organic dye. Mater. Today Proc. 2019, 18, 1760–1769. [Google Scholar] [CrossRef]
  68. Ren, X.; Wei, S.; Wang, Q.; Shi, L.; Wang, X.S.; Wei, Y.; Yang, G.; Philo, D.; Ichihara, F.; Ye, J. Rational construction of dual cobalt active species encapsulated by ultrathin carbon matrix from MOF for boosting photocatalytic H2 generation. Appl. Catal. B Environ. 2021, 286, 119924. [Google Scholar] [CrossRef]
  69. Janczarek, M.; Kowalska, E. On the origin of enhanced photocatalytic activity of copper-modified titania in the oxidative reaction systems. Catalysts 2017, 7, 317. [Google Scholar] [CrossRef]
  70. Jyothi, M.S.; Nayak, V.; Reddy, K.R.; Naveen, S.; Raghu, A.V. Non-metal (Oxygen, Sulphur, Nitrogen, Boron and Phosphorus)-Doped Metal Oxide Hybrid Nanostructures as Highly Efficient Photocatalysts for Water Treatment and Hydrogen Generation. In Nanophotocatalysis and Environmental Applications: Materials and Technology; Springer: Berlin/Heidelberg, Germany, 2019; pp. 83–105. [Google Scholar]
  71. Wang, R.; Tang, T.; Wei, Y.; Dang, D.; Huang, K.; Chen, X.; Yin, H.; Tao, X.; Lin, Z.; Dang, Z.; et al. Photocatalytic debromination of polybrominated diphenyl ethers (PBDEs) on metal doped TiO2 nanocomposites: Mechanisms and pathways. Environ. Int. 2019, 127, 5–12. [Google Scholar] [CrossRef]
  72. Yadav, V.; Verma, P.; Sharma, H.; Tripathy, S.; Saini, V.K. Photodegradation of 4-nitrophenol over B-doped TiO2 nanostructure: Effect of dopant concentration, kinetics, and mechanism. Environ. Sci. Pollut. Res. 2020, 27, 10966–10980. [Google Scholar] [CrossRef]
  73. Sescu, A.M.; Favier, L.; Lutic, D.; Soto-Donoso, N.; Ciobanu, G.; Harja, M. TiO2 doped with noble metals as an efficient solution for the photodegradation of hazardous organic water pollutants at ambient conditions. Water 2020, 13, 19. [Google Scholar] [CrossRef]
  74. Rengaraj, S.; Li, X.Z. Enhanced photocatalytic activity of TiO2 by doping with Ag for degradation of 2, 4, 6-trichlorophenol in aqueous suspension. J. Mol. Catal. A Chem. 2006, 243, 60–67. [Google Scholar] [CrossRef]
  75. Whang, T.J.; Huang, H.Y.; Hsieh, M.T.; Chen, J.J. Laser-induced silver nanoparticles on titanium oxide for photocatalytic degradation of methylene blue. Int. J. Mol. Sci. 2009, 10, 4707–4718. [Google Scholar] [CrossRef] [PubMed]
  76. Fang, T.; Yang, C.; Liao, L. Photoelectrocatalytic degradation of high COD dipterex pesticide by using TiO2/Ni photo electrode. J. Environ. Sci. 2012, 24, 1149–1156. [Google Scholar] [CrossRef]
  77. Khaoulani, S.; Chaker, H.; Cadet, C.; Bychkov, E.; Cherif, L.; Bengueddach, A.; Fourmentin, S. Wastewater treatment by cyclodextrin polymers and noble metal/mesoporous TiO2 photocatalysts. Comptes Rendus Chim. 2015, 18, 23–31. [Google Scholar] [CrossRef]
  78. Rekha, K.; Nirmala, M.; Nair, M.G.; Anukaliani, A. Structural, optical, photocatalytic and antibacterial activity of zinc oxide and manganese doped zinc oxide nanoparticles. Phys. B Condens. Matter. 2010, 405, 3180–3185. [Google Scholar] [CrossRef]
  79. Espitia, P.J.P.; Soares, N.D.F.F.; Coimbra, J.S.D.R.; de Andrade, N.J.; Cruz, R.S.; Medeiros, E.A.A. Zinc oxide nanoparticles: Synthesis, antimicrobial activity and food packaging applications. Food Bioproc. Technol. 2012, 5, 1447–1464. [Google Scholar] [CrossRef]
  80. Muruganandam, G.; Mala, N.; Pandiarajan, S.; Srinivasan, N.; Ramya, R.; Sindhuja, E.; Ravichandran, K. Synergistic effects of Mg and F doping on the photocatalytic efficiency of ZnO nanoparticles towards MB and MG dye degradation. J. Mater. Sci. Mater. 2017, 28, 18228–18235. [Google Scholar] [CrossRef]
  81. Coleman, V.A.; Jagadish, C. Basic properties and applications of ZnO. In Zinc Oxide Bulk, Thin Films and Nanostructures; Elsevier Science Ltd.: Amsterdam, The Netherlands, 2006; pp. 1–20. [Google Scholar]
  82. Shaba, E.Y.; Jacob, J.O.; Tijani, J.O.; Suleiman, M.A.T. A critical review of synthesis parameters affecting the properties of zinc oxide nanoparticle and its application in wastewater treatment. Appl. Water Sci. 2021, 11, 48. [Google Scholar] [CrossRef]
  83. Saedy, S.; Haghighi, M.; Amirkhosrow, M. Hydrothermal synthesis and physicochemical characterization of CuO/ZnO/Al2O3 nanopowder. Part I: Effect of crystallization time. Particuology 2012, 10, 729–736. [Google Scholar] [CrossRef]
  84. Strano, V.; Greco, M.G.; Ciliberto, E.; Mirabella, S. Zno microflowers grown by chemical bath deposition: A low-cost approach for massive production of functional nanostructures. Chemosensors 2019, 7, 62. [Google Scholar] [CrossRef]
  85. Manabeng, M.; Mwankemwa, B.S.; Ocaya, R.O.; Motaung, T.E.; Malevu, T.D. A review of the impact of zinc oxide nanostructure morphology on perovskite solar cell performance. Processes 2022, 10, 1803. [Google Scholar] [CrossRef]
  86. Lee, Y.C.; Yang, C.S.; Huang, H.J.; Hu, S.Y.; Lee, J.W.; Cheng, C.F.; Huang, C.C.; Tsai, M.K.; Kuang, H.C. Structural and optical properties of ZnO nanopowder prepared by microwave-assisted synthesis. J. Lumin 2010, 130, 1756–1759. [Google Scholar] [CrossRef]
  87. Villanueva, Y.Y.; Liu, D.R.; Cheng, P.T. Pulsed laser deposition of zinc oxide. Thin Solid Films 2006, 501, 366–369. [Google Scholar] [CrossRef]
  88. Xu, D.; Yu, Q.; Chen, T.; Zhong, S.; Ma, J.; Bao, L.; Zhang, L.; Zhao, F.; Du, S. Effects of PEG1000 and sol concentration on the structural and optical properties of sol–gel ZnO porous thin films. Materials 2018, 11, 1840. [Google Scholar] [CrossRef]
  89. Mikrovalov, V. Microwave-assisted non-aqueous synthesis of ZnO nanoparticles. Mater. Tehnol. 2011, 45, 173–177. [Google Scholar]
  90. Tecaru, A.; Danciu, A.I.; Muşat, V.; Fortunato, E.; Elangovan, E. Zinc oxide thin films prepared by spray pyrolysis. J. Optoelectron. Adv. Mater. 2010, 12, 1889–1893. [Google Scholar]
  91. Ciolan, M.A.; Motrescu, I. Pulsed laser ablation: A facile and low-temperature fabrication of highly oriented n-type zinc oxide thin films. Appl. Sci. 2022, 12, 917. [Google Scholar] [CrossRef]
  92. Pasang, T.; Namratha, K.; Parvin, T.; Ranganathaiah, C.; Byrappa, K. Tuning of band gap in TiO2 and ZnO nanoparticles by selective doping for photocatalytic applications. Mater. Res. Innov. 2015, 19, 73–80. [Google Scholar] [CrossRef]
  93. Asgari, E.; Esrafili, A.; Jafari, A.J.; Kalantary, R.R.; Nourmoradi, H.; Farzadkia, M. The comparison of ZnO/polyaniline nanocomposite under UV and visible radiations for decomposition of metronidazole: Degradation rate, mechanism and mineralization. Process Saf. Environ. Prot. 2019, 128, 65–76. [Google Scholar] [CrossRef]
  94. Wang, Y.; Yang, C.; Liu, Y.; Fan, Y.; Dang, F.; Qiu, Y.; Zhou, H.; Wang, W.; Liu, Y. Solvothermal Synthesis of ZnO Nanoparticles for Photocatalytic Degradation of Methyl Orange and p-Nitrophenol. Water 2021, 13, 3224. [Google Scholar] [CrossRef]
  95. Abdullah, F.H.; Bakar, N.A.; Bakar, M.A. Current advancements on the fabrication, modification, and industrial application of zinc oxide as photocatalyst in the removal of organic and inorganic contaminants in aquatic systems. J. Hazard. Mater. 2022, 424, 127416. [Google Scholar] [CrossRef]
  96. Lofrano, G.; Libralato, G.; Brown, J. Nanotechnologies for Environmental Remediation; Springer International Publishing: New York, NY, USA, 2017. [Google Scholar]
  97. Kim, S.; Jeong, J.; Hoang, Q.V.; Han, J.W.; Prasetio, A.; Jahandar, M.; Kim, Y.H.; Cho, S.; Lim, D.C. The role of cation and anion dopant incorporated into a ZnO electron transporting layer for polymer bulk heterojunction solar cells. RSC Adv. 2019, 9, 37714–37723. [Google Scholar] [CrossRef]
  98. Güy, N.; Çakar, S.; Özacar, M. Comparison of palladium/zinc oxide photocatalysts prepared by different palladium doping methods for congo red degradation. J. Colloid Interface Sci. 2016, 466, 128–137. [Google Scholar] [CrossRef]
  99. Raj, K.P.; Sadaiyandi, K.; Kennedy, A.; Thamizselvi, R. Structural, optical, photoluminescence and photocatalytic assessment of Sr-doped ZnO nanoparticles. Mater. Chem. Phys. 2016, 183, 24–36. [Google Scholar]
  100. Kuzhalosai, V.; Subash, B.; Senthilraja, A.; Dhatshanamurthi, P.; Shanthi, M. Synthesis, characterization and photocatalytic properties of SnO2–ZnO composite under UV-A light, Spectrochim. Acta A Mol. Biomol. Spectrosc. 2013, 115, 876–882. [Google Scholar] [CrossRef]
  101. Adeel, M.; Saeed, M.; Khan, I.; Muneer, M.; Akram, N. Synthesis and characterization of Co–ZnO and evaluation of its photocatalytic activity for photodegradation of methyl orange. ACS Omega 2021, 6, 1426–1435. [Google Scholar] [CrossRef]
  102. Vallejo, W.; Cantillo, A.; Díaz-Uribe, C. Methylene blue photodegradation under visible irradiation on Ag-doped ZnO thin films. Int. J. Photoenergy 2020, 2020, 1627498. [Google Scholar] [CrossRef]
  103. Vallejo, W.; Cantillo, A.; Salazar, B.; Diaz-Uribe, C.; Ramos, W.; Romero, E.; Hurtado, M. Comparative study of ZnO thin films doped with transition metals (Cu and Co) for methylene blue photodegradation under visible irradiation. Catalysts 2020, 10, 528. [Google Scholar] [CrossRef]
  104. Alkallas, F.H.; Ben Gouider Trabelsi, A.; Nasser, R.; Fernandez, S.; Song, J.M.; Elhouichet, H. Promising Cr-Doped ZnO Nanorods for Photocatalytic Degradation Facing Pollution. Appl. Sci. 2021, 12, 34. [Google Scholar] [CrossRef]
  105. Ragupathy, S.; Priyadharsan, A.; AlSalhi, M.S.; Devanesan, S.; Guganathan, L.; Santhamoorthy, M.; Kim, S.C. Effect of doping and loading Parameters on photocatalytic degradation of brilliant green using Sn doped ZnO loaded CSAC. Environ. Res. 2022, 210, 112833. [Google Scholar] [CrossRef]
  106. Tzewei, T.; Ibrahim, A.H.; Abidin, C.Z.A.; Ridwan, F.M. The effect of iron doping on ZnO catalyst on dye removal efficiency. IOP Conf. Ser. Earth Environ. Sci. 2020, 476, 012108. [Google Scholar] [CrossRef]
  107. Kong, J.Z.; Li, A.D.; Li, X.Y.; Zhai, H.F.; Zhang, W.Q.; Gong, Y.P.; Li, H.; Wu, D. Photo-degradation of methylene blue using Ta-doped ZnO nanoparticle. J. Solid State Chem. 2010, 183, 1359–1364. [Google Scholar] [CrossRef]
  108. Wu, C.; Zhang, Y.C.; Huang, Q. Solvothermal synthesis of N-doped ZnO microcrystals from commercial ZnO powder with visible light-driven photocatalytic activity. Mater. Lett. 2014, 119, 104–106. [Google Scholar] [CrossRef]
  109. Li, B.; Liu, T.; Wang, Y.; Wang, Z. ZnO/graphene-oxide nanocomposite with remarkably enhanced visible-light-driven photocatalytic performance. J. Colloid Interface Sci. 2012, 377, 114–121. [Google Scholar] [CrossRef]
  110. Yi-hao, T.; Hang, Z.; Yin, W.; Ming-hui, D.; Guo, J.; Bin, Z. Facile fabrication of nitrogen-doped zinc oxide nanoparticles with enhanced photocatalytic performance. Micro. Nano. Lett. 2015, 10, 432–434. [Google Scholar] [CrossRef]
  111. Fu, D.; Han, G.; Chang, Y.; Dong, J. The synthesis and properties of ZnO–graphene nano hybrid for photodegradation of organic pollutant in water. Mater. Chem. Phys. 2012, 132, 673–681. [Google Scholar] [CrossRef]
  112. Prabakaran, E.; Pillay, K. Synthesis of N-doped ZnO nanoparticles with cabbage morphology as a catalyst for the efficient photocatalytic degradation of methylene blue under UV and visible light. RSC Adv. 2019, 9, 7509–7535. [Google Scholar] [CrossRef]
  113. Bechambi, O.; Sayadi, S.; Najjar, W. Photocatalytic degradation of bisphenol A in the presence of C-doped ZnO: Effect of operational parameters and photodegradation mechanism. J. Ind. Eng. Chem. 2015, 32, 201–210. [Google Scholar] [CrossRef]
  114. Li, X.; Wang, Q.; Zhao, Y.; Wu, W.; Chen, J.; Meng, H. Green synthesis and photo-catalytic performances for ZnO-reduced graphene oxide nanocomposites. J. Colloid Interface Sci. 2013, 411, 69–75. [Google Scholar] [CrossRef]
  115. Zayed, M.; Samy, S.; Shaban, M.; Altowyan, A.S.; Hamdy, H.; Ahmed, A.M. Fabrication of TiO2/NiO pn nanocomposite for enhancement dye photodegradation under solar radiation. Nanomaterials 2022, 12, 989. [Google Scholar] [CrossRef]
  116. John, K.I.; Ho, G.; Li, D. Recent progresses in synthesis and modification of g-C3N4 for improving visible-light-driven photocatalytic degradation of antibiotics. Water Sci. Technol. 2024, 89, 3047–3078. [Google Scholar] [CrossRef]
  117. Sivaraman, C.; Vijayalakshmi, S.; Leonard, E.; Sagadevan, S.; Jambulingam, R. Current developments in the effective removal of environmental pollutants through photocatalytic degradation using nanomaterials. Catalysts 2022, 12, 544. [Google Scholar] [CrossRef]
  118. Choi, J.; Cho, J.; Roh, C.W.; Kim, B.S.; Choi, M.S.; Jeong, H.; Ham, H.C.; Lee, H. Au-doped PtCo/C catalyst preventing Co leaching for proton exchange membrane fuel cells. Appl. Catal. B Environ. 2019, 247, 142–149. [Google Scholar] [CrossRef]
  119. Nikoofar, K.; Haghighi, M.; Lashanizadegan, M.; Ahmadvand, Z. ZnO nanorods: Efficient and reusable catalysts for the synthesis of substituted imidazoles in water. J. Taibah. Univ. Sci. 2015, 9, 570–578. [Google Scholar] [CrossRef]
  120. Shelar, S.G.; Mahajan, V.K.; Patil, S.P.; Sonawane, G.H. Effect of doping parameters on photocatalytic degradation of methylene blue using Ag doped ZnO nanocatalyst. SN Appl. Sci. 2020, 2, 820. [Google Scholar] [CrossRef]
  121. Wang, X.; Jin, H.; Wu, D.; Nie, Y.; Tian, X.; Yang, C.; Zhou, Z.; Li, Y. Fe3O4@ S-doped ZnO: A magnetic, recoverable, and reusable Fenton-like catalyst for efficient degradation of ofloxacin under alkaline conditions. Environ. Res. 2020, 186, 109626. [Google Scholar] [CrossRef]
  122. Ikram, M.; Aslam, S.; Haider, A.; Naz, S.; Ul-Hamid, A.; Shahzadi, A.; Ikram, M.; Haider, J.; Ahmad, S.O.A.; Butt, A.R. Doping of Mg on ZnO nanorods demonstrated improved photocatalytic degradation and antimicrobial potential with molecular docking analysis. Nanoscale Res. Lett. 2021, 16, 78. [Google Scholar] [CrossRef]
  123. Padikkaparambil, S.; Narayanan, B.; Yaakob, Z.; Viswanathan, S.; Tasirin, S.M. Au/TiO2 reusable photocatalysts for dye degradation. Int. J. Photoenergy 2013, 2013, 752605. [Google Scholar] [CrossRef]
  124. Wu, Q.; Zhang, Z. The preparation of self-floating Sm/N co-doped TiO2/diatomite hybrid pellet with enhanced visible-light-responsive photoactivity and reusability. Adv. Powder Technol. 2019, 30, 415–422. [Google Scholar] [CrossRef]
  125. Song, L.; Jing, W.; Chen, J.; Zhang, S.; Zhu, Y.; Xiong, J. High reusability and durability of carbon-doped TiO2/carbon nanofibrous film as visible-light-driven photocatalyst. J. Mater. Sci. 2019, 54, 3795–3804. [Google Scholar] [CrossRef]
  126. Bahrudin, N.N.; Nawi, M.A.; Nawawi, W.I. Enhanced photocatalytic decolorization of methyl orange dye and its mineralization pathway by immobilized TiO2/polyaniline. Res. Chem. Intermed. 2019, 45, 2771–2795. [Google Scholar] [CrossRef]
  127. Moosavi, S.; Li, R.Y.M.; Lai, C.W.; Yusof, Y.; Gan, S.; Akbarzadeh, O.; Chowhury, Z.Z.; Yue, X.G.; Johan, M.R. Methylene blue dye photocatalytic degradation over synthesised Fe3O4/AC/TiO2 nano-catalyst: Degradation and reusability studies. Nanomaterials 2020, 10, 2360. [Google Scholar] [CrossRef]
  128. Ma, L.; Chen, Y.; Zheng, J. An efficient, stable and reusable polymer/TiO2 photocatalytic membrane for aqueous pollution treatment. J. Mater. Sci. 2021, 56, 11335–11351. [Google Scholar] [CrossRef]
  129. Yeganeh-Faal, A.; Bordbar, M.; Negahdar, N.; Nasrollahzadeh, M. Green synthesis of the Ag/ZnO nanocomposite using Valeriana officinalis L. root extract: Application as a reusable catalyst for the reduction of organic dyes in a very short time. IET Nanobiotechnol. 2017, 11, 669–676. [Google Scholar] [CrossRef]
  130. Mansoori, G.A.; Bastami, T.R.; Ahmadpour, A.; Eshaghi, Z. Environmental application of nanotechnology. Annu. Rev. Nano. Res. 2008, 2, 439–493. [Google Scholar]
  131. Baaloudj, O.; Nasrallah, N.; Bouallouche, R.; Kenfoud, H.; Khezami, L.; Assadi, A.A. High efficient Cefixime removal from water by the sillenite Bi12TiO20: Photocatalytic mechanism and degradation pathway. J. Clean. Prod. 2022, 330, 129934. [Google Scholar] [CrossRef]
  132. Zhang, Y.Z.; Wang, X.; Feng, Y.; Li, J.; Lim, C.T.; Ramakrishna, S. Coaxial electrospinning of (fluorescein isothiocyanate-Conjugated bovine serum albumin)-Encapsulated poly ("-caprolactone) nanofibers for sustained release. Biomacromolecule 2006, 7, 1049–1057. [Google Scholar] [CrossRef]
  133. Chuang, C.S.; Wang, M.K.; Ko, C.H.; Ou, C.C.; Wu, C.H. Removal of benzene and toluene by carbonized bamboo materials modified with TiO2. Bioresour. Technol. 2008, 99, 954–958. [Google Scholar] [CrossRef]
  134. Elmolla, E.S.; Chaudhuri, M. Degradation of amoxicillin, ampicillin and cloxacillin antibiotics in aqueous solution by the UV/ZnO photocatalytic process. J. Hazard. Mater. 2010, 173, 445–449. [Google Scholar] [CrossRef]
  135. Boytsova, O.; Zhukova, I.; Tatarenko, A.; Shatalova, T.; Beiltiukov, A.; Eliseev, A.; Sadovnikov, A. The Anatase-to-Rutile Phase Transition in Highly Oriented Nanoparticles Array of Titania with Photocatalytic Response Changes. Nanomaterials 2022, 12, 4418. [Google Scholar] [CrossRef]
  136. Chiang, Y.J.; Lin, C.C. Photocatalytic decolorization of methylene blue in aqueous solutions using coupled ZnO/SnO2 photocatalysts. Powder Technol. 2013, 246, 137–143. [Google Scholar] [CrossRef]
  137. Kumar, R.; Umar, A.; Kumar, G.; Akhtar, M.S.; Wang, Y.; Kim, S.H. Ce-doped ZnO nanoparticles for efficient photocatalytic degradation of direct red-23 dye. Ceram. Int. 2015, 41, 7773–7782. [Google Scholar] [CrossRef]
  138. Al-Namshah, K.S.; Mariappan, S.M.; Shkir, M.; Hamdy, M.S. Photocatalytic degradation mechanism of Ce-loaded ZnO catalysts toward methyl green dye pollutant. Appl. Phys. A 2021, 127, 452. [Google Scholar] [CrossRef]
  139. Hou, X.; Cai, Y.; Song, X.; Wu, Y.; Zhang, J.; Wei, Q. Electrospun TiO2 nanofibers coated with polydopamine for enhanced sunlight-driven photocatalytic degradation of cationic dyes. Surf. Interface Anal. 2019, 51, 169–176. [Google Scholar] [CrossRef]
  140. Baaloudj, O.; Assadi, A.A.; Azizi, M.; Kenfoud, H.; Trari, M.; Amrane, A.; Assadi, A.A.; Nasrallah, N. Synthesis and characterization of ZnBi2O4 nanoparticles: Photocatalytic performance for antibiotic removal under different light sources. Appl. Sci. 2021, 11, 3975. [Google Scholar] [CrossRef]
  141. Norouzi, M.; Fazeli, A.; Tavakoli, O. Photocatalytic degradation of phenol under visible light using electrospun Ag/TiO2 as a 2D nano-powder: Optimizing calcination temperature and promoter content. Adv. Powder Technol. 2022, 33, 103792. [Google Scholar] [CrossRef]
  142. Touati, A.; Jlaiel, L.; Najjar, W.; Sayadi, S. Photocatalytic degradation of sulfur black dye over Ce-TiO2 under UV irradiation: Removal efficiency and identification of degraded species. Euro-Mediterr. J. Environ. Integr. 2019, 4, 4. [Google Scholar] [CrossRef]
  143. 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, 2014, 536123. [Google Scholar] [CrossRef]
  144. Zhang, Y.; Park, M.; Kim, H.-Y.; El-Newehy, M.; Rhee, K.Y.; Park, S.-J. Effect of TiO2 on photocatalytic activity of polyvinylpyrrolidone fabricated via electrospinning. Compos. B Eng. 2015, 80, 355–360. [Google Scholar] [CrossRef]
  145. Notodarmojo, S.; Sugiyana, D.; Handajani, M.; Kardena, E.; Larasati, A. Synthesis of TiO2 nanofiber-nanoparticle composite catalyst and its photocatalytic decolorization performance of reactive black 5 dye from aqueous solution. J. Eng. Technol. Sci. 2017, 49, 3. [Google Scholar] [CrossRef]
  146. Saiful Amran, S.N.B.; Wongso, V.; Abdul Halim, N.S.; Husni, M.K.; Sambudi, N.S.; Wirzal, M.D.H. Immobilized carbondoped TiO2 in polyamide fibers for the degradation of methylene blue. J. Asian Ceram. Soc. 2019, 7, 321–330. [Google Scholar] [CrossRef]
  147. Erdemoğlu, S.; Aksu, S.K.; Sayılkan, F.; Izgi, B.; Asiltürk, M.; Sayılkan, H.; Frimmel, F.; Güçer, Ş. Photocatalytic degradation of Congo Red by hydrothermally synthesized nanocrystalline TiO2 and identification of degradation products by LC–MS. J. Hazard. Mater. 2008, 155, 469–476. [Google Scholar] [CrossRef] [PubMed]
  148. Xiang, Q.; Yu, J.; Jaroniec, M. Enhanced photocatalytic H2-production activity of graphene-modified titania nanosheets. Nanoscale 2011, 3, 3670–3678. [Google Scholar] [CrossRef]
  149. Fan, H.; Wu, R.; Liu, H.; Yang, X.; Sun, Y.; Chen, C. Synthesis of metal-phase-assisted 1T@2H-MoS2 nanosheet-coated black TiO2 spheres with visible light photocatalytic activities. J. Mater. Sci. 2018, 53, 10302–10312. [Google Scholar] [CrossRef]
  150. Xia, T.; Wang, Y.; Mai, C.; Pan, G.; Zhang, L.; Zhao, W.; Zhang, J. Facile in situ growth of ZnO nanosheets standing on Ni foam as binder-free anodes for lithium ion batteries. RSC Adv. 2019, 9, 19253–19260. [Google Scholar] [CrossRef]
  151. Guo, L.; Ji, Y.L.; Xu, H.; Simon, P.; Wu, Z. Regularly shaped, single-crystalline ZnO nanorods with wurtzite structure. J. Am. Chem. Soc. 2002, 124, 14864–14865. [Google Scholar] [CrossRef]
  152. Zhou, Q.; Xie, B.; Jin, L.; Chen, W.; Li, J. Hydrothermal Synthesis and Responsive Characteristics of Hierarchical Zinc Oxide Nanoflowers to Sulfur Dioxide. J. Nanotechnol. 2016, 2016, 742104. [Google Scholar] [CrossRef]
  153. Li, Q.; Wang, Q.; Chen, Z.; Ma, Q.; An, M. A facile and flexible approach for large-scale fabrication of ZnO nanowire film and its photocatalytic applications. Nanomaterials 2019, 9, 846. [Google Scholar] [CrossRef]
  154. Srivastava, R.R.; Kumar Vishwakarma, P.; Yadav, U.; Rai, S.; Umrao, S.; Giri, R.; Saxena, P.S.; Srivastava, A. 2D SnS2 Nanostructure-Derived Photocatalytic Degradation of Organic Pollutants Under Visible Light. Front. Nanotechnol. 2021, 3, 711368. [Google Scholar] [CrossRef]
  155. Tayeb, A.M.; Hussein, D.S. Synthesis of TiO2 nanoparticles and their photocatalytic activity for methylene blue. Am. J. Nanomater. 2015, 3, 57–63. [Google Scholar]
  156. Tawfik, A.; Alalm, M.G.; Awad, H.M.; Islam, M.; Qyyum, M.A.; Al-Muhtaseb, A.A.H.; Osman, A.I.; Lee, M. Solar photo-oxidation of recalcitrant industrial wastewater: A review. Environ. Chem. Lett. 2022, 20, 1839–1862. [Google Scholar] [CrossRef]
  157. Parida, K.M.; Parija, S. Photocatalytic degradation of phenol under solar radiation using microwave irradiated zinc oxide. Sol. Energy 2006, 80, 1048–1054. [Google Scholar] [CrossRef]
  158. Benhabiles, O.; Mahmoudi, H.; Lounici, H.; Goosen, M.F. Effectiveness of a photocatalytic organic membrane for solar degradation of methylene blue pollutant. Desalin. Water Treat. 2016, 57, 14067–14076. [Google Scholar] [CrossRef]
  159. Mohsenzadeh, M.; Mirbagheri, S.A.; Sabbaghi, S. Degradation of 1,2-dichloroethane by photocatalysis using immobilized PAni-TiO2 nano-photocatalyst. Environ. Sci. Pollut. Res. 2019, 26, 31328–31343. [Google Scholar] [CrossRef]
  160. Hameed, B.H.; Salman, J.M.; Ahmad, A.L. Adsorption isotherm and kinetic modeling of 2,4-D pesticide on activated carbon derived from date stones. J. Hazard. Mater. 2009, 163, 121–126. [Google Scholar] [CrossRef]
  161. Lopez-Alvarez, B.; Torres-Palma, R.A.; Penuela, G. Solar photocatalitycal treatment of carbofuran at lab and pilot scale: Effect of classical parameters, evaluation of the toxicity and analysis of organic by-products. J. Hazard. Mater. 2011, 191, 196–203. [Google Scholar] [CrossRef]
  162. Saien, J.; Khezrianjoo, S. Degradation of the fungicide carbendazim in aqueous solutions with UV/TiO2 process: Optimization, kinetics and toxicity studies. J. Hazard. Mater. 2008, 157, 269–276. [Google Scholar] [CrossRef]
  163. Whyte, H.E. Evaluation of the Performance of Photocatalytic Systems for the Treatment of Indoor Air in Medical Environments. Doctoral Dissertation, Ecole Nationale Supérieure Mines-Télécom Atlantique, Nantes, France, 2018. [Google Scholar]
  164. Lee, K.M.; Lai, C.W.; Ngai, K.S.; Juan, J.C. Recent developments of zinc oxide based photocatalyst in water treatment technology: A review. Water Res. 2016, 88, 428–448. [Google Scholar] [CrossRef]
  165. Elaziouti, A.; Ahmed, B. ZnO-assisted photocatalytic degradation of congo Red and benzopurpurine 4B in aqueous solution. J. Chem. Eng. Process. Technol. 2011, 2, 1000106. [Google Scholar]
  166. Gómez-López, V.M.; Koutchma, T.; Linden, K. Ultraviolet and pulsed light processing of fluid foods. In Novel Thermal and non-Thermal Technologies for Fluid Foods; Academic Press: Cambridge, MA, USA, 2012; pp. 185–223. [Google Scholar]
  167. Bayarri, B.; Abellán, M.N.; Giménez, J.; Esplugas, S. Study of the wavelength effect in the photolysis and heterogeneous photocatalysis. Catal. Today 2007, 129, 231–239. [Google Scholar] [CrossRef]
  168. Nilsson, A. Removal of Formaldehyde from Industrial Waste Water. Master’s Thesis, Chalmers University of Technology, Göteborg, Sweden, 2020. [Google Scholar]
  169. Chen, Y.W.; Hsu, Y.H. Effects of reaction temperature on the photocatalytic activity of TiO2 with Pd and Cu cocatalysts. Catalysts 2021, 11, 966. [Google Scholar] [CrossRef]
  170. Hu, Q.; Liu, B.; Song, M.; Zhao, X. Temperature effect on the photocatalytic degradation of methyl orange under UV-vis light irradiation. J. Wuhan Univ. Technol.-Mater. Sci. Ed. 2010, 25, 210–213. [Google Scholar] [CrossRef]
  171. Barakat, N.A.; Kanjwal, M.A.; Chronakis, I.S.; Kim, H.Y. Influence of temperature on the photodegradation process using Ag-doped TiO2 nanostructures: Negative impact with the nanofibers. J. Mol. Catal. A Chem. 2013, 366, 333–340. [Google Scholar] [CrossRef]
  172. Shi, X.; Zhang, X.; Ma, L.; Xiang, C.; Li, L. TiO2-Doped chitosan microspheres supported on cellulose acetate fibers for adsorption and photocatalytic degradation of methyl orange. Polymers 2019, 11, 1293. [Google Scholar] [CrossRef]
Figure 1. SEM and TEM images of ZnO (a,b) [18] and TiO2 (c,d) Reprinted/adapted with permission from Ref. [19].
Figure 1. SEM and TEM images of ZnO (a,b) [18] and TiO2 (c,d) Reprinted/adapted with permission from Ref. [19].
Catalysts 14 00420 g001
Figure 2. Different categories of organic compounds.
Figure 2. Different categories of organic compounds.
Catalysts 14 00420 g002
Figure 3. Various processes for wastewater treatment along with their advantages and disadvantages.
Figure 3. Various processes for wastewater treatment along with their advantages and disadvantages.
Catalysts 14 00420 g003
Figure 4. Photo-induced charge carrier (e/h+) generation on the TiO2 surface for eliminating contaminants (a) and photodegradation of contaminants on the TiO2 surface (b) Reprinted/adapted with permission from Ref. [58].
Figure 4. Photo-induced charge carrier (e/h+) generation on the TiO2 surface for eliminating contaminants (a) and photodegradation of contaminants on the TiO2 surface (b) Reprinted/adapted with permission from Ref. [58].
Catalysts 14 00420 g004aCatalysts 14 00420 g004b
Figure 5. Types of cocatalysts used in the TiO2 structure.
Figure 5. Types of cocatalysts used in the TiO2 structure.
Catalysts 14 00420 g005
Figure 6. Photodegradation mechanism of contaminants by metal- (a) and non-metal-anchored TiO2 (b) Reprinted/adapted with permission from Refs. [69,70].
Figure 6. Photodegradation mechanism of contaminants by metal- (a) and non-metal-anchored TiO2 (b) Reprinted/adapted with permission from Refs. [69,70].
Catalysts 14 00420 g006
Figure 7. Different structures of ZnO, including salt rock (cubical structure) (a), zinc blend (cubical structure) (b), and wurtzite (hexagonal structure) (c) (Zn and O atoms are represented by grey and black spheres, respectively) Reprinted/adapted with permission from Ref. [81].
Figure 7. Different structures of ZnO, including salt rock (cubical structure) (a), zinc blend (cubical structure) (b), and wurtzite (hexagonal structure) (c) (Zn and O atoms are represented by grey and black spheres, respectively) Reprinted/adapted with permission from Ref. [81].
Catalysts 14 00420 g007
Figure 8. Catalytic mechanism by cationic and anionic cocatalysts (a) and photodegradation mechanism under sunlight (b) on the surface of ZnO NPs Reprinted/adapted with permission from Ref. [95].
Figure 8. Catalytic mechanism by cationic and anionic cocatalysts (a) and photodegradation mechanism under sunlight (b) on the surface of ZnO NPs Reprinted/adapted with permission from Ref. [95].
Catalysts 14 00420 g008
Figure 9. Photodegradation mechanism of contaminants using metal-anchored ZnO (Reprinted/adapted with permission from Ref. [98]) (a) and non-metal-anchored ZnO (Reprinted/adapted with permission from Ref. [99]) (b).
Figure 9. Photodegradation mechanism of contaminants using metal-anchored ZnO (Reprinted/adapted with permission from Ref. [98]) (a) and non-metal-anchored ZnO (Reprinted/adapted with permission from Ref. [99]) (b).
Catalysts 14 00420 g009aCatalysts 14 00420 g009b
Figure 10. Different forms of ZnO nanostructured photocatalysts, including (a) nanosheets [150], (b) nanorods (Reprinted/adapted with permission from Ref. [151]), (c) nanoflowers (Reprinted/adapted with permission from Ref. [152]), and (d) nanowires (Reprinted/adapted with permission from Ref. [153]).
Figure 10. Different forms of ZnO nanostructured photocatalysts, including (a) nanosheets [150], (b) nanorods (Reprinted/adapted with permission from Ref. [151]), (c) nanoflowers (Reprinted/adapted with permission from Ref. [152]), and (d) nanowires (Reprinted/adapted with permission from Ref. [153]).
Catalysts 14 00420 g010
Table 1. Advantages and disadvantages of various techniques for TiO2 fabrication.
Table 1. Advantages and disadvantages of various techniques for TiO2 fabrication.
ProcessAdvantages and DisadvantagesRef.
HydrothermalAdvantages: good size distribution, crystal shape control, low defects, synthesizing large crystals with high quality, fine particle size
Disadvantages: high equipment cost, high temperature and pressure needed, long synthesis time
[55]
Sol-gelAdvantages: high purity products, good size distribution, remarkable specific surface area, economical, uniform size of particles, fine particle size, ease of synthesis
Disadvantages: agglomeration of particles, long processing time, using organic solvents which may be toxic
[55]
Flame pyrolysisAdvantages: rapid and mass production
Disadvantages: requires high energy, ease of rutile formation
[52]
SolvothermalAdvantages: high crystallinity, suitability for materials, low defects, better control of features of TiO2 compared to hydrothermal process
Disadvantages: requires organic solvents, unstable at high temperatures
[52,54]
Inverse micelleAdvantages: fine particle sizes, high crystallinity, low defects
Disadvantages: high cost, high crystallization temperature
[52]
SonochemicalAdvantages: high specific surface area, simple control of particles and morphology, efficient for mesoporous materials, improved reaction rate, short time, no additives
Disadvantages: low yield, inefficient energy
[15,55]
Microwave heatingAdvantages: fast heating, short reaction time, high reaction rate and efficiency[54]
Table 2. Impact of various cocatalysts on TiO2 photocatalytic activity.
Table 2. Impact of various cocatalysts on TiO2 photocatalytic activity.
CocatalystLight Source/PollutantConditionsPE for TiO2 (%)PE for Anchored TiO2 (%)Ref.
PdUV lamp/2,2′,4,4′-tetrabromodiphenylether5% Pd, 300 WUV lamp-100[71]
BUV light/4-nitrophenol5% B in TiO2, 1 g/L catalyst dose, 1 mg/L 4-nitrophenol7990[72]
Au/UIUV irradiation/2,4 dinitrophenol20 mg/L contaminant concentration, 120 min, 1 g/L catalyst dose6037[73]
Au/IWIUV irradiation/2,4 dinitrophenol20 mg/L contaminant concentration, 120 min, 1 g/L catalyst dose6050[73]
Pd/IWIUV irradiation/2,4 dinitrophenol20 mg/L contaminant concentration, 120 min, 1 g/L catalyst dose6067[73]
Pd/IWIUV irradiation/Rhodamine 6G20 mg/L contaminant concentration, 120 min, 1 g/L catalyst dose8896[73]
AgUV-A illumination/2,4,6-trichlorophenol0.5 wt.% Ag, 120 min-95[74]
AgHalogen lamp/MB2 Wt.% Ag, 120 min-82.3[75]
NiUltraviolet/DipterexpH 6, Dipterex concentration = 40 mg/L, 2 h-83.5[76]
CeUV lamp, crystal violet0.8 mol Ce in TiO2, 0.2 g/L catalyst, 30 ppm dye concentration, pH 6.5, intensity of 2000 W/cm27092[15]
FeUV lamp/crystal violet1.2 mol Fe in TiO2, 0.2 g/L catalyst, 30 ppm dye concentration, pH 6.5, intensity of 2000 W/cm27080[15]
AuUV lamp/total organic carbon8.71 mg/L total organic carbon, 15W UV lamp-93[77]
SiUV lamp/MB20 h for TiO2 and 2 h for Si/TiO2, 10 ppm MB6886.7[65]
Table 3. Advantages and disadvantages of some important techniques for the synthesis of ZnO.
Table 3. Advantages and disadvantages of some important techniques for the synthesis of ZnO.
ProcessAdvantages and DisadvantagesRef.
HydrothermalAdvantages: high crystallinity
Disadvantages: high temperature and pressure needed, long reaction time
[83,85]
Sol-gel processAdvantages: uniform distribution of particles, low temperature needed, simple synthesis, controlling the particle size and morphology, low cost
Disadvantages: agglomeration of particles
[85,86,88]
Chemical coprecipitationAdvantages: short time for synthesis, excellent reproducibility, low temperature, technical simplicity
Disadvantages: low crystallinity nanoparticles, produces high waste, expensive
[78,84]
MicrowaveAdvantages: simple, fast synthesis, more energy efficient, uniform distribution of particles
Disadvantages: high cost of microwave reactors
[86,89]
PyrolysisAdvantages: cost-effectiveness, easy way to stick to any element, easy control of the thickness of the films
Disadvantages: operation at moderate temperature
[90]
Pulsed-laser depositionAdvantages: simple setup, high flexibility, good adaptability, high process speed, and generating high-quality transparent films
Disadvantages: difficulty scaling up
[87,91]
Table 4. Comparing different metallic cocatalysts in the ZnO structure on the photocatalytic activity.
Table 4. Comparing different metallic cocatalysts in the ZnO structure on the photocatalytic activity.
CocatalystLight Source/PollutantOperating Conditions* PE (%) for ZnOPE (%) for
Anchored ZnO
Ref.
CoVisible light irradiation/MO10 wt.% Co, 130 min, 100 mg/L MO4693[101]
AgVisible irradiation/MB5 wt.% Ag, 120 min2.745.1[102]
CoVisible light irradiation/MB5 wt.% Co, 10 ppm dye concentration, 140 min2.762.6[103]
CuVisible light irradiation/MB5 wt.% Cu, 10 ppm dye concentration, 140 min2.742.5[103]
CrUV-vis light illumination/MO1 wt.% Cr, 100 min-99.8[104]
SnSunlight/brilliant green120 min72.696.52[105]
FeSunlight/MBTime = 3 h9095[106]
TaVisible light irradiation/MB20 min, 1 g/L catalyst dosage, pH 8, 10 mg/L dye concentration-97.5[107]
* PE: photodegradation efficiency.
Table 5. Impact of various non-metallic cocatalysts on the ZnO photocatalytic activity.
Table 5. Impact of various non-metallic cocatalysts on the ZnO photocatalytic activity.
CocatalystLight Source/PollutantConditionsPE (%) ZnOPE (%) Anchored ZnORef.
NVisible light irradiation/Rhodamine 6G60 min, 0.01 g N76.281.6[108]
CVisible light/MB60 min54.398.1[109]
NVisible light irradiation/Rhodamine B10 mg/L dye concentration, room temperature, 2 h-97[110]
NVisible light irradiation/MB10 mg/L dye concentration, room temperature, 2 h-99[110]
CVisible light/MB2.5% catalyst, 200 °C2680[111]
NUV light or visible light irradiation/MB--99.6[112]
CUV irradiation/Bisphenol A24 h-100[113]
CSunlight irradiation/Rhodamine B2.5 h54.692.9[114]
PE: photodegradation efficiency.
Table 6. Stability and recyclability of various composites of ZnO and TiO2.
Table 6. Stability and recyclability of various composites of ZnO and TiO2.
CatalystPollutantPE * (%)PE (%) After n CyclesRef.
ZnO nanorodsimidazole83%n = 4, 80%[119]
Ag-anchored ZnO nanocompositeMB95%n = 4, 89.5%[120]
Fe3O4@S-ZnOofloxacinAbove 90%n = 6, above 90%[121]
ZnO NPsMB93.25%n = 5, 86.63%[22]
ZnO NPsRhodamine B91.06%n = 5, 83.61%[22]
Mg-ZnO nanorodsMB and ciprofloxacin82%n = 4, 75%[122]
2%Au-anchored TiO2 nanocatalystMO100%n = 11, 100%[123]
Sm/N co-doped TiO2/diatomitetetracycline87.1%n = 5, 83.2%[124]
C-anchored TiO2/carbon nanofibrousRhodamine B94.2%n = 6, 92%[125]
TiO2MO58.3% n = 10, 36.1%[126]
TiO2/polyanilineMO86%n = 10, 46.2%[126]
Fe3O4/AC/TiO2MB98%n = 7, 93%[127]
2-(methacryloyloxy) ethyltrimethylammonium chloride/TiO2MB99.66%n = 20, 98.7%[128]
TiO2/2NiOMB100%n = 5, 72.6%[115]
* PE: Photodegradation efficiency.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

AlMohamadi, H.; Awad, S.A.; Sharma, A.K.; Fayzullaev, N.; Távara-Aponte, A.; Chiguala-Contreras, L.; Amari, A.; Rodriguez-Benites, C.; Tahoon, M.A.; Esmaeili, H. Photocatalytic Activity of Metal- and Non-Metal-Anchored ZnO and TiO2 Nanocatalysts for Advanced Photocatalysis: Comparative Study. Catalysts 2024, 14, 420. https://doi.org/10.3390/catal14070420

AMA Style

AlMohamadi H, Awad SA, Sharma AK, Fayzullaev N, Távara-Aponte A, Chiguala-Contreras L, Amari A, Rodriguez-Benites C, Tahoon MA, Esmaeili H. Photocatalytic Activity of Metal- and Non-Metal-Anchored ZnO and TiO2 Nanocatalysts for Advanced Photocatalysis: Comparative Study. Catalysts. 2024; 14(7):420. https://doi.org/10.3390/catal14070420

Chicago/Turabian Style

AlMohamadi, Hamad, Sameer A. Awad, Ashwani Kumar Sharma, Normurot Fayzullaev, Arístides Távara-Aponte, Lincoln Chiguala-Contreras, Abdelfattah Amari, Carlos Rodriguez-Benites, Mohamed A. Tahoon, and Hossein Esmaeili. 2024. "Photocatalytic Activity of Metal- and Non-Metal-Anchored ZnO and TiO2 Nanocatalysts for Advanced Photocatalysis: Comparative Study" Catalysts 14, no. 7: 420. https://doi.org/10.3390/catal14070420

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