Next Article in Journal
Carbon Nitride-Perovskite Composites: Evaluation and Optimization of Photocatalytic Hydrogen Evolution in Saccharides Aqueous Solution
Next Article in Special Issue
Characterization of Anaerobic Biofilms Growing on Carbon Felt Bioanodes Exposed to Air
Previous Article in Journal
Self-Regeneration Effect of Three-Way Catalysts during Thermal Aging Procedure
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Photocatalysis for Organic Wastewater Treatment: From the Basis to Current Challenges for Society

by
Salma Izati Sinar Mashuri
1,
Mohd Lokman Ibrahim
1,2,*,
Muhd Firdaus Kasim
2,
Mohd Sufri Mastuli
2,
Umer Rashid
3,*,
Abdul Halim Abdullah
4,
Aminul Islam
5,
Nurul Asikin Mijan
6,
Yie Hua Tan
7,
Nasar Mansir
8,9,
Noor Haida Mohd Kaus
10 and
Taufiq-Yap Yun Hin
11
1
School of Chemistry and Environment, Faculty of Applied Sciences, Universiti Teknologi MARA, Shah Alam 40450, Malaysia
2
Centre for Functional Materials and Nanotechnology, Institute of Science, Universiti Teknologi MARA, Shah Alam 40450, Malaysia
3
Institute of Advanced Technology, Universiti Putra Malaysia (UPM), Serdang 43400, Malaysia
4
Department of Chemistry, Faculty of Science, Universiti Putra Malaysia (UPM), Serdang 43400, Malaysia
5
Department of Petroleum and Mining Engineering, Jessore University of Science and Technology, Jashore Sadar 7408, Bangladesh
6
Department of Chemical Sciences, Faculty of Science and Technology, Universiti Kebangsaan Malaysia, Bangi 43600, Malaysia
7
Department of Chemical Engineering, Faculty of Engineering and Science, Curtin University Malaysia, CDT 250, Miri 98009, Malaysia
8
Catalysis Science and Technology Research Centre, Faculty of Science, Universiti Putra Malaysia, Serdang 43400, Malaysia
9
Department of Chemistry, Federal University, Dutse 7156, Nigeria
10
Nano|Hybrid|Materials Research Group, School of Chemical Sciences, Universiti Sains Malaysia, Penang 11800, Malaysia
11
Faculty of Science and Natural Resources, Universiti Malaysia Sabah, Kota Kinabalu 88400, Malaysia
*
Authors to whom correspondence should be addressed.
Catalysts 2020, 10(11), 1260; https://doi.org/10.3390/catal10111260
Submission received: 14 September 2020 / Revised: 23 October 2020 / Accepted: 25 October 2020 / Published: 30 October 2020

Abstract

:
Organic pollutants such as dyes, antibiotics, analgesics, herbicides, pesticides, and stimulants become major sources of water pollution. Several treatments such as absorptions, coagulation, filtration, and oxidations were introduced and experimentally carried out to overcome these problems. Nowadays, an advanced technique by photocatalytic degradation attracts the attention of most researchers due to its interesting and promising mechanism that allows spontaneous and non-spontaneous reactions as they utilized light energy to initiate the reaction. However, only a few numbers of photocatalysts reported were able to completely degrade organic pollutants. In the past decade, the number of preparation techniques of photocatalyst such as doping, morphology manipulation, metal loading, and coupling heterojunction were studied and tested. Thus, in this paper, we reviewed details on the fundamentals, common photocatalyst preparation for coupling heterojunction, morphological effect, and photocatalyst’s characterization techniques. The important variables such as catalyst dosage, pH, and initial concentration of sample pollution, irradiation time by light, temperature system, durability, and stability of the catalyst that potentially affect the efficiency of the process were also discussed. Overall, this paper offers an in-depth perspective of photocatalytic degradation of sample pollutions and its future direction.

1. Introduction

Nowadays, about 300 to 400 million tons of untreated organic pollutants are produced annually, which leads to water pollution problems, especially near industrial areas [1]. To overcome this issue, the majority of countries introduced strict regulations to control environmental pollution. Moreover, it attracts scientists’ attention to intensively study the best technology in this scope of research with the hope to control environmental pollution and improve the wellbeing of the environment.
One of the attractive solutions to degrade organic pollutants is by using photocatalysts. This is because of its promising, effective, and efficient degradation activity of pollutants, which occur by allowing both spontaneous and non-spontaneous reactions to optimize the whole process. Photocatalysis was introduced in 1972 by Fujishima and Honda in their research on the electrochemical photolysis of water at a semiconductor electrode and published in Nature [2]. The process is defined as a chronological succession of advanced oxidation processes (AOPs), which improve their drawbacks such as high cost, incomplete mineralization, and require high hydroxyl radical [3]. The photocatalysis process requires light energy to activate the photocatalyst, hence this shows interest because the reaction can be controlled using the light or photon sources.
Currently, the most studied photocatalysts are titanium dioxide (TiO2) [4,5,6] and zinc oxide (ZnO) [7,8,9]. However, they suffer from the recombination of the electrons (e) and holes (h+), which are considered as the major drawback to these materials. Improving the efficiency and stability of photocatalyst become the target goal among researchers.
In this review, we focused on the hybridization type, preparation technique, effect of parameters, and current challenges of photocatalysis sciences in organic wastewater treatment. As has been mentioned before, the recombination of the electrons (e) and holes (h+) are common problems for the photocatalyst. Hence, the hybridization technique was proposed by several researchers to overcome this problem. Examples of such techniques are doping, coupling heterojunction, and supporting materials. Further mechanisms and explanations on the effect of hybridization type on the performance of catalysts are discussed in this review.
Several techniques were also proposed by researchers to synthesize photocatalyst, however, there is no single study that reviewed the performance of each method and quality of produced photocatalyst. In this paper, there are two types of green and simple techniques identified that could affectively affect the performance of the photocatalyst such as physical- and chemical-based techniques. Moreover, this paper also discussed extensively the effect of operating parameters involved during the photocatalysis process. For example, the effect of catalyst amount, pH of pollutant, irradiation intensity, the temperature of pollutant, initial concentration, and effect of size and shape of photocatalyst toward the rate of degradation of organic pollutants.

2. Photocatalysis

Catalysis is the study involving synthesis, modification, and mechanism of a substance that can increase or accelerate the rate of a chemical reaction due to the participation of a substance called a catalyst, which remains unchanged at the end of the reaction. The reaction will occur faster with the catalyst because they require less activation energy than that of a normal reaction. The activation energy (Ea) is threshold energy, which must be overcome for a reaction to occur as illustrated in Figure 1. Despite the normal reaction pathway, the presence of the catalyst opens up an alternative approach with lower activation energy, thus, the rate of reaction will increase. Nevertheless, the result and the overall thermodynamics are the same [10].
Nowadays, photocatalysis shows high potential in reclaiming the environmental practice and green technology. It drives researchers to enhance this technology by looking forward to the best type of photocatalyst and the reactors. Currently, it has been applied in many applications such as coating technology [11,12], environmental pollution [13,14], and air pollution treatment [15,16].
Theoretically, photocatalysis requires light to activate the photocatalyst to initiate the reaction. According to Ohtani [17], the difference between photocatalytic and catalytic reaction depends on the preferential crystal facets of the photocatalyst. Therefore, the intensity of irradiated light will affect the kinetics of the reaction, to produce electron–hole pairs, which makes the reaction happens, whereas the catalytic reaction depends on the active sites for the reaction to occur [18,19,20]. In catalysis, the density of active sites concurrent with the reaction kinetic [21,22] as illustrated in Figure 2. Generally, the catalytic reaction is limited to spontaneous reaction only, in which the Gibbs free energy is more than zero. However, photocatalyst allows both spontaneous and non-spontaneous reactions, which relied on the photo-absorption ability of the material that can provide energy source and turn it into chemical energy [17].
As illustrated in Figure 3, the activity of photocatalyst (i.e., semiconductor) depends on the ability to create e and h+ pairs to generate free radicals, which are needed to initiate the reaction. An electron from the valence band (VB) will be excited to the conductive band (CB) by absorption of the light energy equally or more than its band gap, which is an energy difference between VB and CB in the semiconductor.
Theoretically, the reaction starts when enough photons (hv) from the light source hit the e on VB. The e excites the CB, leaving the h+ on VB. Both e and h+ will migrate to the surface of the photocatalyst. Simultaneously on the surface, h+ will oxidize the water to form hydroxyl radicals (OH·) that initiate the chain reaction to oxidize the organic pollutants. While e will donate to the electron acceptor such as oxygen (O2) extending the formation of superoxide or metal ion that is reduced to its lower valence state and deposited on the photocatalyst surface [23]. Both oxidation and reduction processes produce a degraded product of organic pollutant that is more environmentally friendly.
Table 1 shows recent publications on the catalytic study of the photocatalyst in the degradation of (i) dyes like methyl orange and methylene blue, (ii) antibiotics such as doxycycline, levofloxacin, tetracycline hydrochloride, and tetracycline, (iii) analgesics such as acetaminophen, (iv) herbicides such as atrazine, (v) pesticides such as naphthalene, and (vi) stimulants such as caffeine—all of which could benefit from further research and applications. According to Gogate and Pandit [24] and Low et al. [25], organic contaminants from industries were found to be the major source of water pollution. Globally, around 15 g pharmaceutical products were consumed per capita and 3 to 10 times higher in developed countries [26]. The remaining untreated refractory organic contaminants could stimulate microbial growth, leading to oxygen depletion and disturb the entire water bodies’ ecosystem [27]. Numerous research reported in this area is scientifically important. For example, Cheshme et al. [28] and Vaiano et al. [4] in their research, succeeded in completely degrading the methyl orange within 50 min and paracetamol sample in 120 min using Cu doped ZnO/Al2O3 and TiO2-graphite composite photocatalyst, respectively.

3. Types of Hybridization Photocatalyst

The single semiconductor was used as a photocatalyst, facing the problem of recombination of e and h+ due to limited diffusion length and large band gap [44]. In the past decade, researchers put effort into hybridizing the photocatalyst to enhance the light absorption in the range of visible sunlight. Therefore, several types of hybridization of photocatalysts such as doping, metal loading, and coupling heterojunction have been explored to solve the problem.

3.1. Metal-Doped Photocatalyst

Doping is a method used to add other substance that has energy levels almost the same as the valence band or conducting band edge of the main semiconductor, that enables it to enhance the concentration of charge carriers either by donating or accepting electrons. However, the opposite effect has been observed, which could promote e and h+ recombination [45]. Thus, morphology manipulation is important to produce nano-sized photocatalyst to reduce travel distance from the e to the surface hence creating h+ in the bulk phase [46]. Nevertheless, it also has a side effect, which leads to an increase in the bandgap due to the particle-in-box model [46]. According to Ge et al. [47], metal loading requires a balanced amount of loading to favor the effective reaction to occur effectively. When an excess of metal is loaded, it will cover the active site on the surface, yet it decreases the separation capacity: it may also act as the center of recombination.
Theoretically, the introduction of metal dopant on photocatalyst could improve both adsorption and photocatalytic degradation efficiencies. As been reported by Cao et al. [48] in Table 2, the addition of 1 wt.% to 8 wt.% of Co Zr6O4(OH4)BDC12 in photocatalyst improved the adsorption and degradation process from 9.9% to 78.5%. This is because of the photocatalyst’s crystallinity level, which improves the light adsorption and increases the efficiency of charge separation. Fundamentally, the introduced dopant or metal on photocatalyst has introduced a sub-energy level below the conductive band. Moreover, it acts as a trapping site for excitation and delay of the recombination of the excited electron as shown in Figure 4.
Furthermore, the synergy between dopant and photocatalyst is also an important criterion that must be observed. Based on the reported studies presented in Table 2, where Mg-doped ZnO-Al2O3 without or with minimum metal dopant showing low photocatalytic degradation. Further addition of dopant to the optimum level (1 wt.%) improves the photocatalysis degradation up to 98.9%. However, the addition of dopant (3–5 wt.%) exceeded the optimum level causes the decrease of the photocatalysis degradation lower than 98.9%. The decrease of the photocatalytic degradation might be due to several factors: (1) the addition of excess dopant, increases the crystallinity of catalyst, which hinder their active site on the surface [43] and (2) the excess of dopant tends to act as the center of recombination, which accelerates the recombination between e and h+ [28].

3.2. Coupling Heterojunction Photocatalyst

A coupling heterojunction is a combination of semiconductor to other semiconductor(s). It creates variation towards interfacial interactions between the semiconductors, which resulted in a new unique photocatalyst. Based on the band alignment, the heterojunctions are categorized into three types as shown in Figure 5a.
In semiconductor/semiconductor heterojunction, the e and h+ flow towards less negative potential and less positive potential, respectively. Type I shows the CB in the first semiconductor (SC-1), which is more negative than the second semiconductor (SC-2) while VB is more positive. Therefore, both e and h+ are accumulated in SC-2 and make them potentially recombine and reduce photocatalytic degradation. Subsequently, the type II band alignment is the most preferred because e in CB of SC-1 will flow to SC-2, while h+ in SC-2 will flow to SC-1 and both will transfer to the surface to undergo redox reaction. On the other hand, type III shows no heterojunction because SC-1 and SC-2 will work as a single semiconductor. In each semiconductor, the e tends to recombine to its own h+ [51].
Meanwhile, Figure 5b–d show the binary, ternary, and quaternary coupling heterojunctions. The same concept was applied as in Figure 5a, which depends on the level of band gaps for the coupled metals which determine the types of band alignment.

3.3. Supported Material Photocatalyst

Currently, the trend in the development of active catalyst’s support has attracted much attention due to its ability to disperse the active site on the catalyst surface. The support could also enhance the process due to its responsibility to provide a high surface area for depositing the primary photocatalyst test solutions [55,56]. The support’s material can be inert or active during the photocatalytic process, which could act as a co-catalyst or secondary catalyst [57]. It is a requisite to overcome the difficulty in separating the catalyst usually in the form of powder, microcrystalline, and nanocrystalline after mixing with an aqueous sample and make it a milky dispersion [58]. To minimize the inconvenient, researchers discovered that the support material such as polymer membrane [59,60], silica [61,62], metal oxide [63,64], graphene [65,66], zeolite [67], carbon nanotube (CNT) [68], ceramics [69] and aluminum oxide [70] play key roles in maximizing the efficiency of the catalyst.
Several reported research papers studied the efficiency of metal oxides like the catalyst’s support. For example, titanium dioxide [71,72], zinc oxide [73,74], silicon dioxide [75,76], and aluminum oxide [77,78]. However, most of the introduced supports suffer from having a low surface area and a lack of physical and chemical stabilities. Besides, some of these oxides offer good support; these include aluminum oxide, which offers very high surface area, high thermal conductivity, as well as promising chemical and physical stabilities based on the previous research by Shi et al. [78], Larimi and Khorasheh [77], and Sun et al. [79].
The chemical illustration of support materials of graphene, carbon nanotube, and aluminum oxide are presented in Figure 6. From this Figure, it is observed that the support material is one of the most important materials used to increase the absorption of the sample molecules, which allow the dispersion of photocatalyst on their exposed surface. Hence, it is potentially increasing the performance of the degradation process.

4. Preparation Techniques of Photocatalyst

In the past few years, various approaches of catalyst preparation techniques such as hydrothermal, electrospinning, impregnation, co-precipitation, solid-state, sonication, microemulsion, thermal evaporation, and sol-gel have been introduced, as shown in Table 3. These methods can be divided into solid-based and solution-based methods.

4.1. Preparation of Photocatalyst by Physical Techniques

The preparation of photocatalyst using the physical technique is commonly based on the machine, force, energy, and pressure to mix or to modify the structure of the materials such as the sonication [28], solid-state [43], and thermal evaporation [85]. These methods offer several benefits such as simple technical aspects and a high yield of purity compared to chemical-based techniques. However, the composition, shape, and size of the catalyst are difficult to control [96].
A Sonication bath is one of the instruments that can be used to synthesize the photocatalyst using sound energy. The ultrasonic probe delivers the ultrasonic waves that come from different power capacities, pressure, and structural form as shown in Figure 7. According to Warner et al. [97] and Taylor [98], different products will be obtained with different frequency, amplitude, power supply, and processing time. It was also reported that this product has a high potential to produce a pure product with high yield and minimal waste from the process.
According to Sobhani-Nasab et al. [80], the nanosheet of Ce(MoO4)2 photocatalyst as shown in Figure 8a has been successfully synthesized by the initial reactant of Ce(NO3)3·6H2O and (NH4)6Mo7O24·4H2O solution with the addition of glucose as surfactant using sonication. The surfactant played a key role in modeling the photocatalyst into a nanosheet. The research reported that the product was of high purity based on the X-ray diffraction (XRD) and energy dispersive spectrometer (EDS) analysis results. However, it was found that the increase of ultrasonic power could increase the nanosheet size of the catalyst. Therefore, the power needs to be analyzed so that to be maintained on the nano size scale. Similarly, Liang et al. [81], reported that ultrasonic power was used in the bonding of g-C3N4 and SnO. At a lower amount of g-C3N4, the observed morphology was in the form of a nanosheet as shown in Figure 8b; however, at a higher amount of g-C3N4, the morphology became irregular due to excess g-C3N4, which inhibited the formation of nanosheet and reduced the specific surface area of the photocatalysts.
Meanwhile, the solid-state is a facile additive-free method that is promising in synthesizing the product in a large quantity [100]. This simple technique is normally employed for commercial production in the industry. The temperature is usually applied until the unwanted substances decompose and produce the product [82]. Without contact with any additive, the product produces a high yield with high purity. Despite that, the promising benefit of this method is that it is a simple and easy technique to operate. Figure 9 shows the illustration of the solid-state technique for the preparation of the photocatalyst.
Elhalil et al. [43] in their report, found that the solid-state method produced an Mg doped ZnO-Al2O3 photocatalyst when the calcination process was able to remove the carbonate group based on their FTIR result, thus, coupling the semiconductors. Their findings also showed that there are no impurities involved during the process, which was confirmed by the Joint Committee on Powder Diffraction Standards (JCPDS) data of XRD analysis. The porosity of the photocatalyst is presented in Figure 10a. Similarly, according to Rokesh et al. [82], the ZnO produced using the solid-state method was found to be of high purity as well as high crystallinity as proved by XRD data and energy dispersive x-ray (EDX) spectrum. The calcination process made the structure of microcrystal folding as shown in Figure 10b, thereby producing the rough surface and large quantity of surface defect. It also indicated the high surface area of the ZnO. Consequently, this method is strongly convinced to be good considering it is simplicity and ability to produce products with high purity and crystallinity.

4.2. Preparation of Photocatalyst by Chemical Techiques

The solution-based method is the most common technique used for the preparation of photocatalysts because it offers numerous advantages such as being environmentally friendly, having affordable reagents, and requires incredibly low energy input. From the previous study, the efficiency of the produced photocatalyst can be controlled by manipulating the operating parameters. The variables involved during the process design could result in different compositions, shapes, and sizes, which potentially affects the performance of photocatalyst. Therefore, this method is preferable compared to a solid-based method. The common examples of these methods are sol-gel [36], co-precipitation [34], electrospinning [37], and hydrothermal [29], see Table 3.

4.2.1. Electrospinning Technique

Electrospinning is the process that requires the precursors in the form of solution or suspension, in order to be transferred into the syringe pump, so that the spinning tip will eject the sample drop by drop. The droplets are aided by the electrical field, which will be charged. The surface tension will subsequently be overtaken by electrostatic repulsion. The droplet repulse and elongate until it is deposited on a collector plate or drum [51]. The synthesized product by this technique is usually in nanofibers form. Figure 11 shows an illustration of the electrospinning technique.
According to Sekar et al. [102], the product obtained was smooth and has a uniform surface of nanofibers as shown in Figure 12a. Therefore, additional calcination was compulsorily needed to increase the surface area and its porosity. Meanwhile, Li et al. [87] reported that the electrospinning method creates a lack of uniformity of product diameter due to the interference of the inner needle. Hence, the best viscosity of precursor is necessary to avoid wrinkled nanofibers as shown in Figure 12b.

4.2.2. Sol-Gel Technique

There are few steps involved in the sol-gel technique, which include hydrolysis, condensation, and the drying process [8]. Generally, a metal precursor undergoes a rapid hydrolysis process to form the metal hydroxide. Rapid condensation is then applied to produce the gel followed by a drying process to solidify the gel [103]. Figure 13 shows the illustrations of the sol-gel technique in the preparation ZnO [104].
Research conducted by Li et al. [95] reported that high purity of Bi2Mo3O12 photocatalyst was produced in the Nano foam shaped using a sol-gel technique. Figure 14a presents the uniform and compacted size nanoparticles. The nanoparticles provide a large surface area in the catalyst, which is beneficial for catalytic activity. Meanwhile, Vinoth et al. [36] claimed that they obtained high purity crystal of Nd2Mo3O9 without any redundant impurities. In addition, the flower-like photocatalyst shown in Figure 14b below contains a fairly uniform distribution of Nd, Mo, and O elements. Thus, the sol-gel method is believed to produce high purity and uniformly size photocatalyst. Moreover, the steps involved are quite simple and support sustainable green technology.

4.2.3. Co-Precipitation Technique

The co-precipitation method is used to design different morphologies of photocatalysts based on the operating parameters. Both anionic and cationic solutions are mixed and stirred together to form a uniform mixture. The nucleation occurs in the formed mixture when one ion is replaced by another ion, forming the crystal lattice. The growing process will continue until the precipitation agent is added to agglomerate and formed the stable colloid suspension or precipitant. The obtained product from this method can be control either by solution concentration, pH, washing medium, and calcination temperature [96]. Figure 15 shows the illustration of the co-precipitation technique of Fe3O4 [105].
As reported by Pan et al. [91], the ZnO photocatalyst obtained has a flower-like shape as shown in Figure 16a. They control the morphology of the catalyst by manipulating the concentration of Zn2+ and OH. Therefore, the concentration of aqueous ZnCl2 and NaOH is increased to obtain a high reactant concentration. In line with Cao et al. [34], the porous hollow cube shape of ZnFe2O4 shown in Figure 16b was controlled by the synthesis of Prussian Blue precursor. During nucleation and growth, polyvinylpyrrolidone (PVP) was added as a capping agent, to produce the cube shape. In this process, calcination is needed to decompose PVP and cyanide ligands, while gas was also diffused to produce the small voids.

5. Degradation of Various Source of Pollutants

Generally, the sources of organic pollutants in water originally comes from domestic sewage, municipal sewage, and industrial wastewater, which commonly contained organic and inorganic types of pollutants that partially biodegrade. Moreover, even at low concentrations, it tends to poison aquatic organisms [106]. Therefore, chemical degradation process such as photocatalysis is required to remove the organic pollutant into more environmentally friendly substances. The pollutants that were reportedly studied were dye-based, antibiotic-based, and other-based pollutants such as herbicides and toxic substances.
Truong et al. [5], reported having used graphene@Fe-Ti binary oxide composites under the exposure of 350 W Xenon sunlight simulated lamp to degrade 10 mg/L methylene blue dye solution. About 6 mg photocatalyst was used and successfully degraded by about 100% for 20 min. According to Cheshme et al. [28], the Cu doped ZnO/Al2O3 under visible light of 400 W and a high-pressure mercury-vapor lamp with a wavelength of 546.8 nm, could degrade 100% of methyl orange (15 ppm) within 50 min with 0.6 g/L photocatalyst.
Based on Cao et al. [34], in their work, they successfully degraded about 84.08% of 500 ppm tetracycline hydrochloride using ZnFe2O4 photocatalyst within 50 min. 300 W Xenon lamp was used to supply the light to 100 mL solution with a 40 mg/L photocatalyst dosage. In keeping with Osotsi et al. [35], ZnWO4−x photocatalyst was used to degrade 20 ppm of tetracycline solution. About 2 g/L photocatalyst dosage was used under the exposure of 300 W of a ultra violet (UV) lamp. Therefore, 91% of degradation was achieved within 80 min.
As reported by Sekar et al. [102], 40 ppm of naphthalene was successfully degraded under a 16 W UV lamp using Fe-ZnO/PVA photocatalyst. About 96% of naphthalene solution was also degraded in 4 h with 0.06 g/L of Fe-ZnO/PVA. Moreover, Moyet et al. [41], used 0.1 g/L of Cu-BiOCl photocatalyst under the exposure of mercury UV lamp with 254 nm wavelength, degrading 10 ppm of atrazine solution in 30 min. Other examples are shown in Table 4.

6. Effect of Parameters on the Efficiency of Photocatalytic Degradation Process

Several variables can affect the performance of the photocatalysis process. The variables include catalyst dosage, pH, irradiation time, temperature, and initial concentration. These variables might be affecting the photodegradation of pollutants. Therefore, an analysis with a series of experiments is needed to get the optimum operational parameters for the process.

6.1. Photocatalyst Dosage

Catalyst dosage has a major impact on the photocatalysis process as described by many researchers. According to Cheshme et al. [28], the dosage range of 0.005 to 0.04 g of Al2O3/ZnO:Cu photocatalyst was used for the degradation of 15 mg/L of methyl orange. However, the increment in photodegradation of methyl orange solution occurs with the addition of only 0.005 to 0.03 g of catalyst dosage. However, the addition of 0.04 g of catalyst exceeded the optimum amount of the amount of catalyst used and the abundance of catalysts will only reduce its performance. Therefore, the optimum photocatalyst loading was found to be 0.03 g. According to Elhalil et al. [43], about 0.1 to 0.3 g/L showed an increment in degradation efficiency of 20 mg/L caffeine from 69.42% to 98.9%. As the amount of photocatalyst increased to more than 0.3 g/L, the photocatalytic degradation showed a slight decrease. This was due to the excess amount of photocatalyst, which scattered the light and reduced its penetration into the solution. Therefore, their optimum photocatalyst dosage was found to be 0.3 g with 98.9% caffeine solution degradation.

6.2. pH of Wastewater Sample

The pH plays a major role in photocatalytic degradation. This is because it has an important role to play on the surface charge of the material in the aqueous media, especially in adsorption studies. If the surface charge of the material is opposing the adsorption due to the fact of having the same charge as the adsorbate, which needs to be modified and find the pH conditions that show the best adsorption. Therefore, there is a need to find out the pH at which the surface charge of the material is zero in the aqueous media or the pHpzc (pH point of zero charge) of the adsorbent material [117,118].
Elhalil et al. [43], discovered the optimum pH value of 9.5 for Mg-ZnO-Al2O3 photocatalyst to degrade 20 mg/L caffeine solution. The dramatically decreased was found with a pH of 3.5. The researchers reported that the pH solution affects the surface charge and ionization of caffeine molecules, as a result, it enhanced hydroxyl radical formation. Compared to a pH of 3.5, many factors contribute simultaneously such as non-favorable adsorption, dissolution, and decomposition of the photocatalyst. However, Subash et al. [119], reported that the optimum pH of 9 was recorded for the higher adsorption efficiency for photodegradation of Acid Black 1 dye solution recording up to 90.1% using ZnO photocatalyst. At acidic pH level, the removal efficiency was less due to the dissolution of the photocatalyst. Therefore, the solution of pH must be tested as the photocatalyst surface charge is inverse to the solution charge.

6.3. Irradiation Intensity

Irradiation of light intensity affects the photocatalysis process via the production of more hydroxyl radicals as intensity increase [120]. Tekin and Saygi [121] reported that at a higher intensity (132 W/m2), 25 mg/L of acid black 1 dye solution was degraded completely for 40 min reactions. Compared to low intensity (44 W/m2), which has only an 80% degradation rate. Bhatia et al. [122] also reported the degradation of atenolol solution (25 mg/L) with 1.5 g/L TiO2-graphene photocatalyst under 1000 mW/cm2, the degradation almost completed in 60 min compared to under irradiation of 250 mW/cm2, the solution degraded to 20 mg/L only. An increase in the irradiation light intensity can produce more photons to interact with a photocatalyst, thus, releasing more hydroxyl radicals and superoxide ions. However, the effective cost should be calculated to determine the best light source intensity and high efficiency of photodegradation of pollutants.

6.4. Temperature of Wastewater Sample

Usually, the elevation of temperature will increase the photodegradation within a certain range. It is agreeable to Tambat et al. [123] findings, which reported that at room temperature and high temperature (60 °C), the methyl orange solutions were completely degraded using CeO2 photocatalyst. In contrary to low temperature (20 °C), the solution was not degraded completely. Moreover, according to Ateş et al. [124], the degradation of methyl orange solution using Al2O3-NP/SnO2 photocatalyst increased from 60.90% to 93.95% as the temperature increased from 10 to 55 °C. However, beyond that, the rate of reaction temperature could give a negative effect when electron–hole recombine and cause a decrease in degradation reaction rate [121]. Thus, it is crucial to control the temperature in the range of reaction temperature.

6.5. Initial Concentration of Wastewater Sample

It is important to specify the optimum initial concentration of solution in the photocatalytic process due to the synergistic effect between photocatalyst, the formation of the oxidizing agent, and the pollutant. Subash et al. [119] reported a decrease in removal efficiency when they increased the Acid Black 1 dye concentration. The report claimed that the path length of the photons from the light source decreases, thus, it did not fully penetrate the solution. On the contrary, Elhalil et al. [43] increased the initial concentration of caffeine solution, which lead to an increase in the degradation rate. This was due to the higher ability of collision between hydroxyl radicals and that of caffeine molecules. Therefore, the initial concentration is one of the variables that can influence the photocatalysis.

6.6. Stability and Durability of Photocatalyst

The stability and durability of the photocatalyst are important to ensure the efficiency and the quality of the synthesized catalyst. Based on the previous studies, the morphological characteristics affect the stability and durability of the photocatalyst. The reduction of performance of the photocatalyst is known as the deactivation process. This phenomenon is also depending on the type of reactions, solvents, and the mechanical process. Most of the photocatalyst deactivation is discovered after several cycles of the reaction process. Several factors lead to the deactivation of the photocatalyst. These include (1) loss of photocatalyst mass especially during the washing/purification process, for example, the research reported by Taddesse et al. [125], where the Rod-like ZnO stacking on Cu2O/Ag3PO4, which has lost some quantity during filtration; (2) leaching of dopants is also a common phenomenon that normally happens during the reaction due to photo-etching as experienced by Zhang et al. [126], when the synthesized ZnO catalyst undergo photodecomposition after three cycles in degradation of Rhodamine B; (3) residual pollutant and/or organic intermediates that adsorbed on the surface of photocatalyst. This normally occurs for the nanosized-photocatalyst with high porosity surface. For example, the nanocomposite of WO3-ZnO has uniform surface morphology with a concomitant irregular distribution with an average thickness of 27 nm, which showed a gradual decrease in degradation until the fifth cycle. The researchers discovered that the methylene blue of organic intermediates adsorbed on the photocatalyst surface even on every cycle [127,128]. In summary, the strength, durability, and stability of the catalyst towards a harsh environment need to be analyzed and carefully studied. This is to ensure that the prepared catalyst has high reusability and recycle the ability to minimize the cost of the process.

7. Future Direction

Currently, several photocatalysts were developed and introduced, especially for water treatment technologies. This is due to their inexpensive cost, efficiency, and being environmentally friendly. However, every technique introduced in the reported papers in past decades have their own advantages and weaknesses in terms of catalyst stability, efficiency, cost production, structure, and performance of the catalyst. In this review, the recent reported common techniques used to prepare the catalyst to provide a better understanding to the readers and researchers were analyzed.
For example, the most common technique used was co-precipitation, which is easy to control the morphology of the catalyst. However, this technique involves several problems such as being easy to be contaminated and complex purification process. Tremendous attention has been put on the improvement of the discussed techniques. The factors that need to focus to improve the techniques such as (1) bandgap of the semi-conductors; (2) type of carrier transport; (3) crystallinity of the materials; (4) surface area; (5) stability of photocatalyst, which could be controlled using the proper technique of preparation. All these factors potentially affect the electron–hole recombination, synergic between pollutant and photocatalyst, purities of the materials, availability of active sites, and reusability of the photocatalyst, respectively. These factors are the clear paths of directions for the next researchers who studied the sciences behind photocatalysis technology.
The hurdles, limitations, challenges, and research gap in this evolution of studies become an opportunity for researchers to explore and find a clear direction for them. Figure 17 shows the key factors of different types of hybridization of photocatalyst, which might be the main guide or reference for other researchers to study the effect of phase structure towards the catalytic activity.
This review offers a comprehensive summary and novel insight into photocatalysis: These insights include (i) hybridization of photocatalyst; (ii) preparation techniques; (iii) degradation of various sources; (iv) effects of parameter towards photocatalysis reaction, thus aiming the future study to design and develop improved photocatalysts that are functional and hence facilitating the photocatalysis process for industrial scale. Figure 18 shows the importance of optimum reaction parameters during photodegradation processes. This is crucial to ensure the maximum capacity and performance of the catalyst–reactant relationship.

8. Summary

Photocatalyst shows promising techniques to degrade organic pollutants to environmentally friendly substances by allowing both spontaneous and non-spontaneous reactions. From the previous studies, the single semiconductor normally facing the recombination problem of electron–hole pairs. The hybridization photocatalyst was proposed to overcome the problems. However, it has slight weaknesses such as promoting the center of recombination, absence of interfacial interaction, and imposing lower photocatalytic degradation. Consequently, the proper preparation and hybridization techniques are important to ensure the maximum potential of the catalyst, which was discussed in this review. Adequate methods were proposed by researchers to synthesize the photocatalyst, however, to the best of our knowledge, no study compares the methods and quality of photocatalyst production. Therefore, this manuscript provides detailed observation, comparison, and conclusion on the previously reported papers working on photocatalysis, especially for organic wastewater treatment in recent years. Moreover, clear direction to overcome the challenges to the researchers and society from the basis of key factors and parameters were highlighted and discussed.

Author Contributions

Writing—original draft preparation, S.I.S.M.; conceptualization, S.I.S.M. and U.R.; visualization, S.I.S.M.; methodology, S.I.S.M. and M.F.K.; writing—review and editing, M.L.I., N.A.M., and A.I.; investigation, S.I.S.M. and Y.H.T.; validation, M.L.I., M.F.K, M.S.M., N.M., and N.H.M.K.; supervision, M.L.I., A.H.A., and T.-Y.Y.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Malaysia Higher Education for the FRGS research fund File No. FRGS/1/2018/STG07/UITM/03/5.

Acknowledgments

The authors would acknowledge the Malaysia Higher Education and Universiti Teknologi MARA (UiTM) for the FRGS research fund File No. FRGS/1/2018/STG07/UITM/03/5. Special thanks to the Institute of Science (IOS), Universiti Teknologi MARA for all the facilities provided throughout this work.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Palaniappan, M.; Gleick, P.H.; Allen, L.; Cohen, M.J.; Christian-Smith, J.; Smith, C. References. In Clearing the Waters. A Focus on Water Quality Solutions; Ross, N., Ed.; Pacific Institute: Nairobi, Kenya, 2010; pp. 1–88. ISBN 978-92-807-3074-6. [Google Scholar]
  2. Fujushima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef] [PubMed]
  3. Ong, C.B.; Ng, L.Y.; Mohammad, A.W. A review of ZnO nanoparticles as solar photocatalysts: Synthesis, mechanisms and applications. Renew. Sustain. Energy Rev. 2018, 81, 536–551. [Google Scholar] [CrossRef]
  4. Vaiano, V.; Sacco, O.; Matarangolo, M. Photocatalytic degradation of paracetamol under UV irradiation using TiO2-graphite composites. Catal. Today 2018, 315, 230–236. [Google Scholar] [CrossRef]
  5. Truong, N.T.; Thi, H.P.N.; Ninh, H.D.; Phung, X.T.; Van Tran, C.; Nguyen, T.T.; Pham, T.D.; Dang, T.D.; Chang, S.W.; Rene, E.R.; et al. Facile fabrication of graphene@Fe-Ti binary oxide nanocomposite from ilmenite ore: An effective photocatalyst for dye degradation under visible light irradiation. J. Water Process Eng. 2020, 37, 101474. [Google Scholar] [CrossRef]
  6. He, Y.; Sutton, N.B.; Rijnaarts, H.H.H.; Langenhoff, A.A.M. Degradation of pharmaceuticals in wastewater using immobilized TiO2 photocatalysis under simulated solar irradiation. Appl. Catal. B Environ. 2016, 182, 132–141. [Google Scholar] [CrossRef]
  7. Kosera, V.S.; Cruz, T.M.; Chaves, E.S.; Tiburtius, E.R.L. Triclosan degradation by heterogeneous photocatalysis using ZnO immobilized in biopolymer as catalyst. J. Photochem. Photobiol. A Chem. 2017, 344, 184–191. [Google Scholar] [CrossRef]
  8. Hasnidawani, J.N.; Azlina, H.N.; Norita, H.; Bonnia, N.N.; Ratim, S.; Ali, E.S. Synthesis of ZnO nanostructures using sol-gel method. Procedia Chem. 2016, 19, 211–216. [Google Scholar] [CrossRef] [Green Version]
  9. Shaban, M.; Abdallah, S.; Khalek, A.A. Characterization and photocatalytic properties of cotton fibers modified with ZnO nanoparticles using sol–gel spin coating technique. Beni-Suef Univ. J. Basic Appl. Sci. 2016, 5, 277–283. [Google Scholar] [CrossRef] [Green Version]
  10. Abas, M.T. Physical Chemistry II, 3rd ed.; UiTM Press: Shah Alam, Malaysia, 2009. [Google Scholar]
  11. Wu, B.; Lu, S.; Xu, W.; Cui, S.; Li, J.; Han, P.F. Study on corrosion resistance and photocatalysis of cobalt superhydrophobic coating on aluminum substrate. Surf. Coat. Technol. 2017, 330, 42–52. [Google Scholar] [CrossRef]
  12. Stojadinović, S.; Tadić, N.; Radić, N.; Stojadinović, B.; Grbić, B.; Vasilić, R. Synthesis and characterization of Al2O3/ZnO coatings formed by plasma electrolytic oxidation. Surf. Coat. Technol. 2015, 276, 573–579. [Google Scholar] [CrossRef]
  13. Bansal, P.; Verma, A.; Talwar, S. Detoxification of real pharmaceutical wastewater by integrating photocatalysis and photo-Fenton in fixed-mode. Chem. Eng. J. 2018, 349, 838–848. [Google Scholar] [CrossRef]
  14. Bharatvaj, J.; Preethi, V.; Kanmani, S. Hydrogen production from sulphide wastewater using Ce3+–TiO2 photocatalysis. Int. J. Hydrogen Energy 2018, 43, 3935–3945. [Google Scholar] [CrossRef]
  15. Mera, A.C.; Martínez-de la Cruz, A.; Pérez-Tijerina, E.; Meléndrez, M.F.; Valdés, H. Nanostructured BiOI for air pollution control: Microwave-assisted synthesis, characterization and photocatalytic activity toward NO transformation under visible light irradiation. Mater. Sci. Semicond. Process. 2018, 88, 20–27. [Google Scholar] [CrossRef]
  16. Faraldos, M.; Kropp, R.; Anderson, M.A.; Sobolev, K. Photocatalytic hydrophobic concrete coatings to combat air pollution. Catal. Today 2016, 259, 228–236. [Google Scholar] [CrossRef]
  17. Ohtani, B. Photocatalysis A to Z—What we know and what we do not know in a scientific sense. J. Photochem. Photobiol. C Photochem. Rev. 2010, 11, 157–178. [Google Scholar] [CrossRef] [Green Version]
  18. Ahmad, F.M.A.; Hassan, M.A.; Taufiq-Yap, Y.H.; Ibrahim, M.L.; Othman, M.R.; Ali, A.A.M.; Shirai, Y. Production of methyl esters from waste cooking oil using a heterogeneous biomass-based catalyst. Renew. Energy 2017, 114, 638–643. [Google Scholar] [CrossRef]
  19. Ibrahim, M.L.; Nik, A.K.N.N.A.; Islam, A.; Rashid, U.; Ibrahim, S.F.; Mashuri, S.I.S.; Taufiq-Yap, Y.H. Preparation of Na2O supported CNTs nanocatalyst for efficient biodiesel production from waste-oil. Energy Convers. Manag. 2020, 205, 112445. [Google Scholar] [CrossRef]
  20. Mansir, N.; Hwa, S.T.; Lokman, I.M.; Taufiq-Yap, Y.H. Synthesis and application of waste egg shell derived CaO supported W-Mo mixed oxide catalysts for FAME production from waste cooking oil: Effect of stoichiometry. Energy Convers. Manag. 2017, 151, 216–226. [Google Scholar] [CrossRef]
  21. Ahmad, F.M.A.; Hassan, M.A.; Taufiq-Yap, Y.H.; Ibrahim, M.L.; Hasan, M.Y.; Ali, A.A.M.; Othman, M.R.; Shirai, Y. Kinetic and thermodynamic of heterogeneously K3PO4/AC-catalysed transesterification via pseudo-first order mechanism and Eyring-Polanyi equation. Fuel 2018, 232, 653–658. [Google Scholar] [CrossRef]
  22. Ibrahim, S.F.; Asikin-Mijan, N.; Ibrahim, M.L.; Abdulkareem-Alsultan, G.; Izham, S.M.; Taufiq-Yap, Y.H. Sulfonated functionalization of carbon derived corncob residue via hydrothermal synthesis route for esterification of palm fatty acid distillate. Energy Convers. Manag. 2020, 210, 112698. [Google Scholar] [CrossRef]
  23. Mahlambi, M.M.; Ngila, C.J.; Mamba, B.B. Recent developments in environmental photocatalytic degradation of organic pollutants: The case of titanium dioxide nanoparticles-A review. J. Nanomater. 2015, 2015, 1–29. [Google Scholar] [CrossRef] [Green Version]
  24. Gogate, P.R.; Pandit, A.B. A review of imperative technologies for wastewater treatment I: Oxidation technologies at ambient conditions. Adv. Environ. Res. 2004, 8, 501–551. [Google Scholar] [CrossRef]
  25. Low, J.; Yu, J.; Ho, W. Graphene-based photocatalysts for CO2 reduction to solar fuel. J. Phys. Chem. Lett. 2015, 6, 4244–4251. [Google Scholar] [CrossRef] [PubMed]
  26. Zhang, Y.; Geißen, S.U.; Gal, C. Carbamazepine and diclofenac: Removal in wastewater treatment plants and occurrence in water bodies. Chemosphere 2008, 73, 1151–1161. [Google Scholar] [CrossRef] [PubMed]
  27. Sirota, J.; Baiser, B.; Gotelli, N.J.; Ellison, A.M. Organic-matter loading determines regime shifts and alternative states in an aquatic ecosystem. Proc. Natl. Acad. Sci.USA 2013, 110, 7742–7747. [Google Scholar] [CrossRef] [Green Version]
  28. Cheshme, K.A.H.; Mahjoub, A.; Bayat, R.M. Low temperature one-pot synthesis of Cu-doped ZnO/Al2O3 composite by a facile rout for rapid methyl orange degradation. J. Photochem. Photobiol. B Biol. 2017, 175, 37–45. [Google Scholar] [CrossRef]
  29. Nasr, M.; Viter, R.; Eid, C.; Habchi, R.; Miele, P.; Bechelany, M. Optical and structural properties of Al2O3 doped ZnO nanotubes prepared by ALD and their photocatalytic application. Surf. Coat. Technol. 2018, 343, 24–29. [Google Scholar] [CrossRef]
  30. Gao, J.; Gao, Y.; Sui, Z.; Dong, Z.; Wang, S.; Zou, D. Hydrothermal synthesis of BiOBr/FeWO4 composite photocatalysts and their photocatalytic degradation of doxycycline. J. Alloys Compd. 2018, 732, 43–51. [Google Scholar] [CrossRef]
  31. Al-Balushi, B.S.M.; Al-Marzouqi, F.; Al-Wahaibi, B.; Kuvarega, A.T.; Al-Kindy, S.M.Z.; Kim, Y.; Selvaraj, R. Hydrothermal synthesis of CdS sub-microspheres for photocatalytic degradation of pharmaceuticals. Appl. Surf. Sci. 2018, 457, 559–565. [Google Scholar] [CrossRef]
  32. Gong, Y.; Wu, Y.; Xu, Y.; Li, L.; Li, C.; Liu, X.; Niu, L. All-solid-state Z-scheme CdTe/TiO2 heterostructure photocatalysts with enhanced visible-light photocatalytic degradation of antibiotic wastewater. Chem. Eng. J. 2018, 350, 257–267. [Google Scholar] [CrossRef]
  33. Wang, D.; Jia, F.; Wang, H.; Chen, F.; Fang, Y.; Dong, W.; Zeng, G.; Li, X.; Yang, Q.; Yuan, X. Simultaneously efficient adsorption and photocatalytic degradation of tetracycline by Fe-based MOFs. J. Colloid Interface Sci. 2018, 519, 273–284. [Google Scholar] [CrossRef] [PubMed]
  34. Cao, Y.; Lei, X.; Chen, Q.; Kang, C.; Li, W.; Liu, B. Enhanced photocatalytic degradation of tetracycline hydrochloride by novel porous hollow cube ZnFe2O4. J. Photochem. Photobiol. A Chem. 2018, 364, 794–800. [Google Scholar] [CrossRef]
  35. Osotsi, M.I.; Macharia, D.K.; Zhu, B.; Wang, Z.; Shen, X.; Liu, Z.; Zhang, L.; Chen, Z. Synthesis of ZnWO4−x nanorods with oxygen vacancy for efficient photocatalytic degradation of tetracycline. Prog. Nat. Sci. Mater. Int. 2018, 28, 408–415. [Google Scholar] [CrossRef]
  36. Vinoth, K.J.; Karthik, R.; Chen, S.M.; Chen, K.H.; Sakthinathan, S.; Muthuraj, V.; Chiu, T.W. Design of novel 3D flower-like neodymium molybdate: An efficient and challenging catalyst for sensing and destroying pulmonary toxicity antibiotic drug nitrofurantoin. Chem. Eng. J. 2018, 346, 11–23. [Google Scholar] [CrossRef]
  37. Tu, H.; Li, D.; Yi, Y.; Liu, R.; Wu, Y.; Dong, X.; Shi, X.; Deng, H. Incorporation of rectorite into porous polycaprolactone/TiO2 nanofibrous mats for enhancing photocatalysis properties towards organic dye pollution. Compos. Commun. 2019, 15, 58–63. [Google Scholar] [CrossRef]
  38. Zhou, F.; Yan, C.; Liang, T.; Sun, Q.; Wang, H. Photocatalytic degradation of Orange G using sepiolite-TiO2 nanocomposites: Optimization of physicochemical parameters and kinetics studies. Chem. Eng. Sci. 2018, 183, 231–239. [Google Scholar] [CrossRef]
  39. Liu, S.J.; Li, F.T.; Li, Y.L.; Hao, Y.J.; Wang, X.J.; Li, B.; Liu, R.H. Fabrication of ternary g-C3N4/Al2O3/ZnO heterojunctions based on cascade electron transfer toward molecular oxygen activation. Appl. Catal. B Environ. 2017, 212, 115–128. [Google Scholar] [CrossRef]
  40. Stojadinović, S.; Vasilić, R.; Radić, N.; Tadić, N.; Stefanov, P.; Grbić, B. The formation of tungsten doped Al2O3/ZnO coatings on aluminum by plasma electrolytic oxidation and their application in photocatalysis. Appl. Surf. Sci. 2016, 377, 37–43. [Google Scholar] [CrossRef]
  41. Moyet, M.A.; Arthur, R.B.; Lueders, E.E.; Breeding, W.P.; Patterson, H.H. The role of Copper (II) ions in Cu-BiOCl for use in the photocatalytic degradation of atrazine. J. Environ. Chem. Eng. 2018, 6, 5595–5601. [Google Scholar] [CrossRef]
  42. Singh, P.; Mondal, K.; Sharma, A. Reusable electrospun mesoporous ZnO nanofiber mats for photocatalytic degradation of polycyclic aromatic hydrocarbon dyes in wastewater. J. Colloid Interface Sci. 2013, 394, 208–215. [Google Scholar] [CrossRef]
  43. Elhalil, A.; Elmoubarki, R.; Farnane, M.; Machrouhi, A.; Sadiq, M.; Mahjoubi, F.Z.Z.; Qourzal, S.; Barka, N. Photocatalytic degradation of caffeine as a model pharmaceutical pollutant on Mg doped ZnO-Al2O3 heterostructure. Environ. Nanotechnol. Monit. Manag. 2018, 10, 63–72. [Google Scholar] [CrossRef]
  44. Takanabe, K. Photocatalytic water splitting: Quantitative approaches toward photocatalyst by design. ACS Catal. 2017, 7, 8006–8022. [Google Scholar] [CrossRef]
  45. Li, J.; Cushing, S.K.; Zheng, P.; Senty, T.; Meng, F.; Bristow, A.D.; Manivannan, A.; Wu, N. Solar hydrogen generation by a CdS-Au-TiO2 sandwich nanorod array enhanced with au nanoparticle as electron relay and plasmonic photosensitizer. J. Am. Chem. Soc. 2014, 136, 8438–8449. [Google Scholar] [CrossRef]
  46. Zhang, Z.; Wang, C.C.; Zakaria, R.; Ying, J.Y. Role of particle size in nanocrystalline TiO2-based photocatalysts. J. Phys. Chem. B 1998, 102, 10871–10878. [Google Scholar] [CrossRef]
  47. Ge, L.; Han, C.; Liu, J.; Li, Y. Enhanced visible light photocatalytic activity of novel polymeric g-C3N4 loaded with Ag nanoparticles. Appl. Catal. A Gen. 2011, 409–410, 215–222. [Google Scholar] [CrossRef]
  48. Cao, J.; Yang, Z.H.; Xiong, W.P.; Zhou, Y.Y.; Peng, Y.R.; Li, X.; Zhou, C.Y.; Xu, R.; Zhang, Y.R. One-step synthesis of Co-doped UiO-66 nanoparticle with enhanced removal efficiency of tetracycline: Simultaneous adsorption and photocatalysis. Chem. Eng. J. 2018, 353, 126–137. [Google Scholar] [CrossRef]
  49. Gupta, V.K.; Fakhri, A.; Azad, M.; Agarwal, S. Synthesis and characterization of Ag doped ZnS quantum dots for enhanced photocatalysis of Strychnine as a poison: Charge transfer behavior study by electrochemical impedance and time-resolved photoluminescence spectroscopy. J. Colloid Interface Sci. 2018, 510, 95–102. [Google Scholar] [CrossRef]
  50. Nguyen, C.H.; Fu, C.C.; Juang, R.S. Degradation of methylene blue and methyl orange by palladium-doped TiO2 photocatalysis for water reuse: Efficiency and degradation pathways. J. Clean. Prod. 2018, 202, 413–427. [Google Scholar] [CrossRef]
  51. Zhang, L.; Jaroniec, M. Toward designing semiconductor-semiconductor heterojunctions for photocatalytic applications. Appl. Surf. Sci. 2018, 430, 2–17. [Google Scholar] [CrossRef]
  52. Li, C.; Ma, Y.; Zheng, S.; Hu, C.; Qin, F.; Wei, L.; Zhang, C.; Duo, S.; Hu, Q. One-pot synthesis of Bi2O3/Bi2O4 p-n heterojunction for highly efficient photocatalytic removal of organic pollutants under visible light irradiation. J. Phys. Chem. Solids 2020, 140, 109376. [Google Scholar] [CrossRef]
  53. Zhou, T.; Zhang, H.; Ma, X.; Zhang, X.; Zhu, Y.; Zhang, A.; Cao, Y.; Yang, P. Construction of AgI/Bi2MoO6/AgBi(MoO4)2 multi-heterostructure composite nanosheets for visible-light photocatalysis. Mater. Today Commun. 2020, 23, 100903. [Google Scholar] [CrossRef]
  54. Lin, X.; Guo, X.; Shi, W.; Zhai, H.; Yan, Y.; Wang, Q. Quaternary heterostructured Ag-Bi2O2CO3/Bi3.64Mo0.36O6.55/Bi2MoO6 composite: Synthesis and enhanced visible-light-driven photocatalytic activity. J. Solid State Chem. 2015, 229, 68–77. [Google Scholar] [CrossRef]
  55. Sakata, Y.; Tamaura, Y.; Imamura, H.; Watanabe, M. Preparation of a new type of CaSiO3 with high surface area and property as a catalyst support. In Studies in Surface Science and Catalysis; Elsevier: Amsterdam, The Netherlands, 2006; Volume 162, pp. 331–338. [Google Scholar]
  56. Zheng, Y.; Liu, J.; Cheng, B.; You, W.; Ho, W.; Tang, H. Hierarchical porous Al2O3@ZnO core-shell microfibres with excellent adsorption affinity for Congo red molecule. Appl. Surf. Sci. 2019, 473, 251–260. [Google Scholar] [CrossRef]
  57. IUPAC. Compendium of Chemical Terminology; Nič, M., Jirát, J., Košata, B., Jenkins, A., McNaught, A., Eds.; IUPAC: Research Triagle Park, NC, USA, 2009. [Google Scholar]
  58. Loddo, V.; Marcì, G.; Martín, C.; Palmisano, L.; Rives, V.; Sclafani, A. Preparation and characterisation of TiO2(anatase) supported on TiO2(rutile) catalysts employed for 4-nitrophenol photodegradation in aqueous medium and comparison with TiO2(anatase) supported on Al2O3. Appl. Catal. B Environ. 1999, 20, 29–45. [Google Scholar] [CrossRef]
  59. Li, Y.; Yeung, K.L. Polymeric catalytic membrane for ozone treatment of DEET in water. Catal. Today 2019, 331, 53–59. [Google Scholar] [CrossRef]
  60. Korolkov, I.V.; Mashentseva, A.A.; Güven, O.; Gorin, Y.G.; Kozlovskiy, A.L.; Zdorovets, M.V.; Zhidkov, I.S.; Cholach, S.O. Electron/gamma radiation-induced synthesis and catalytic activity of gold nanoparticles supported on track-etched poly(ethylene terephthalate) membranes. Mater. Chem. Phys. 2018, 217, 31–39. [Google Scholar] [CrossRef]
  61. Popa, A.; Sasca, V.; Verdes, O.; Oszko, A. Preparation and catalytic properties of cobalt salts of Keggin type heteropolyacids supported on mesoporous silica. Catal. Today 2018, 306, 233–242. [Google Scholar] [CrossRef]
  62. Ren, J.; Hao, P.; Sun, W.; Shi, R.; Liu, S. Ordered mesoporous silica-carbon-supported copper catalyst as an efficient and stable catalyst for catalytic oxidative carbonylation. Chem. Eng. J. 2017, 328, 673–682. [Google Scholar] [CrossRef]
  63. Liu, P.; Wei, G.; He, H.; Liang, X.; Chen, H.; Xi, Y.; Zhu, J. The catalytic oxidation of formaldehyde over palygorskite-supported copper and manganese oxides: Catalytic deactivation and regeneration. Appl. Surf. Sci. 2019, 464, 287–293. [Google Scholar] [CrossRef]
  64. Hu, E.; Wu, X.; Shang, S.; Tao, X.M.; Jiang, S.X.; Gan, L. Catalytic ozonation of simulated textile dyeing wastewater using mesoporous carbon aerogel supported copper oxide catalyst. J. Clean. Prod. 2016, 112, 4710–4718. [Google Scholar] [CrossRef]
  65. Pawar, R.C.; Lee, C.S. Single-step sensitization of reduced graphene oxide sheets and CdS nanoparticles on ZnO nanorods as visible-light photocatalysts. Appl. Catal. B Environ. 2014, 144, 57–65. [Google Scholar] [CrossRef]
  66. Zhang, D.; Zhang, L. Ultrasonic-assisted sol-gel synthesis of rugby-shaped SrFe2O4/reduced graphene oxide hybrid as versatile visible light photocatalyst. J. Taiwan Inst. Chem. Eng. 2016, 69, 156–162. [Google Scholar] [CrossRef]
  67. Sacco, O.; Vaiano, V.; Matarangolo, M. ZnO supported on zeolite pellets as efficient catalytic system for the removal of caffeine by adsorption and photocatalysis. Sep. Purif. Technol. 2018, 193, 303–310. [Google Scholar] [CrossRef]
  68. Sharifi, A.; Montazerghaem, L.; Naeimi, A.; Abhari, A.R.; Vafaee, M.; Ali, G.A.M.; Sadegh, H. Investigation of photocatalytic behavior of modified ZnS:Mn/MWCNTs nanocomposite for organic pollutants effective photodegradation. J. Environ. Manag. 2019, 247, 624–632. [Google Scholar] [CrossRef]
  69. Zhang, G.; Jin, W.; Xu, N. Design and fabrication of ceramic catalytic membrane reactors for green chemical engineering applications. Engineering 2018, 4, 848–860. [Google Scholar] [CrossRef]
  70. Neves, T.M.; Frantz, T.S.; do Schenque, E.C.C.; Gelesky, M.A.; Mortola, V.B. An investigation into an alternative photocatalyst based on CeO2/Al2O3 in dye degradation. Environ. Technol. Innov. 2017, 8, 349–359. [Google Scholar] [CrossRef]
  71. Boreriboon, N.; Jiang, X.; Song, C.; Prasassarakich, P. Fe-based bimetallic catalysts supported on TiO2 for selective CO2 hydrogenation to hydrocarbons. J. CO2 Util. 2018, 25, 330–337. [Google Scholar] [CrossRef]
  72. Hu, D.; Liu, C.; Li, L.; Lv, K.L.; Zhang, Y.H.; Li, J.L. Carbon dioxide reforming of methane over nickel catalysts supported on TiO2 (001) nanosheets. Int. J. Hydrogen Energy 2018, 3, 21345–21354. [Google Scholar] [CrossRef]
  73. Nouri, F.; Rostamizadeh, S.; Azad, M. Synthesis of a novel ZnO nanoplates supported hydrazone-based palladacycle as an effective and recyclable heterogeneous catalyst for the Mizoroki-Heck cross-coupling reaction. Inorg. Chim. Acta 2018, 471, 664–673. [Google Scholar] [CrossRef]
  74. Fujita, S.I.; Mitani, H.; Zhang, C.; Li, K.; Zhao, F.; Arai, M. Pd and PdZn supported on ZnO as catalysts for the hydrogenation of cinnamaldehyde to hydrocinnamyl alcohol. Mol. Catal. 2017, 442, 12–19. [Google Scholar] [CrossRef]
  75. Mitoraj, D.; Lamdab, U.; Kangwansupamonkon, W.; Pacia, M.; Macyk, W.; Wetchakun, N.; Beranek, R. Revisiting the problem of using methylene blue as a model pollutant in photocatalysis: The case of InVO4/BiVO4 composites. J. Photochem. Photobiol. A Chem. 2018, 366, 103–110. [Google Scholar] [CrossRef]
  76. Li, J.; Liu, Z.; Wang, R. Support structure and reduction treatment effects on CO oxidation of SiO2 nanospheres and CeO2 nanorods supported ruthenium catalysts. J. Colloid Interface Sci. 2018, 531, 204–215. [Google Scholar] [CrossRef] [PubMed]
  77. Larimi, A.; Khorasheh, F. Renewable hydrogen production by ethylene glycol steam reforming over Al2O3 supported Ni-Pt bimetallic nano-catalysts. Renew. Energy 2018, 128, 188–199. [Google Scholar] [CrossRef]
  78. Shi, D.; Wang, H.; Kovarik, L.; Gao, F.; Wan, C.; Hu, J.Z.; Wang, Y. WOx supported on γ-Al2O3 with different morphologies as model catalysts for alkanol dehydration. J. Catal. 2018, 363, 1–8. [Google Scholar] [CrossRef] [Green Version]
  79. Sun, J.; Wang, Y.; Zou, H.; Guo, X.; Wang, Z. Ni catalysts supported on nanosheet and nanoplate γ-Al2O3 for carbon dioxide methanation. J. Energy Chem. 2019, 29, 3–7. [Google Scholar] [CrossRef] [Green Version]
  80. Sobhani-Nasab, A.; Maddahfar, M.; Hosseinpour-Mashkani, S.M. Ce(MoO4)2 nanostructures: Synthesis, characterization, and its photocatalyst application through the ultrasonic method. J. Mol. Liq. 2016, 216, 1–5. [Google Scholar] [CrossRef]
  81. Liang, B.; Han, D.; Sun, C.; Zhang, W.; Qin, Q. Synthesis of SnO/g-C3N4 visible light driven photocatalysts via grinding assisted ultrasonic route. Ceram. Int. 2018, 44, 7315–7318. [Google Scholar] [CrossRef]
  82. Rokesh, K.; Nithya, A.; Jeganathan, K.; Jothivenkatachalam, K. A facile solid state synthesis of cone-like ZnO microstructure an efficient solar-light driven photocatalyst for Rhodamine B degradation. Mater. Today Proc. 2016, 3, 4163–4172. [Google Scholar] [CrossRef]
  83. Miwa, T.; Kaneco, S.; Katsumata, H.; Suzuki, T.; Ohta, K.; Verma, S.C.; Sugihara, K. Photocatalytic hydrogen production from aqueous methanol solution with CuO/Al2O3/TiO2 nanocomposite. Int. J. Hydrogen Energy 2010, 35, 6554–6560. [Google Scholar] [CrossRef]
  84. Yang, Z.X.; Zhu, F.; Zhou, W.M.; Zhang, Y.F. Novel nanostructures of β-Ga2O3 synthesized by thermal evaporation. Phys. E Low-Dimens. Syst. Nanostruct. 2005, 30, 93–95. [Google Scholar] [CrossRef]
  85. Zou, C.; Liang, F.; Xue, S. Synthesis and oxygen vacancy-related photocatalytic properties of ZnO nanotubes grown by thermal evaporation. Res. Chem. Intermed. 2015, 41, 5167–5176. [Google Scholar] [CrossRef]
  86. Pan, Z.W. Nanobelts of semiconducting oxides. Science 2001, 291, 1947–1949. [Google Scholar] [CrossRef] [Green Version]
  87. Li, S.; Zhao, Y.; Wang, C.; Li, D.; Gao, K. Fabrication and characterization unique ribbon-like porous Ag/LaFeO3 nanobelts photocatalyst via electrospinning. Mater. Lett. 2016, 170, 122–125. [Google Scholar] [CrossRef]
  88. Ahluwalia, S.; Prakash, N.T.; Prakash, R.; Pal, B. Improved degradation of methyl orange dye using bio-co-catalyst Se nanoparticles impregnated ZnS photocatalyst under UV irradiation. Chem. Eng. J. 2016, 306, 1041–1048. [Google Scholar] [CrossRef]
  89. Parida, K.M.; Pradhan, A.C.; Das, J.; Sahu, N. Synthesis and characterization of nano-sized porous gamma-alumina by control precipitation method. Mater. Chem. Phys. 2009, 113, 244–248. [Google Scholar] [CrossRef]
  90. Amouzegar, Z.; Naghizadeh, R.; Rezaie, H.R.; Ghahari, M.; Aminzare, M. Cubic ZnWO4 nano-photocatalysts synthesized by the microwave-assisted precipitation technique. Ceram. Int. 2015, 41, 1743–1747. [Google Scholar] [CrossRef]
  91. Li, B.; Wang, Y. Facile synthesis and enhanced photocatalytic performance of flower-like ZnO hierarchical microstructures. J. Phys. Chem. C 2010, 114, 890–896. [Google Scholar] [CrossRef]
  92. Ayyub, P.; Multani, M.; Barma, M.; Palkar, V.R.; Vijayaraghavan, R. Size-induced structural phase transitions and hyperfine properties of microcrystalline Fe2O3. J. Phys. C Solid State Phys. 1988, 21, 2229–2245. [Google Scholar] [CrossRef]
  93. Kumar, H.; Rani, R. Structural and optical characterization of ZnO nanoparticles synthesized by microemulsion route. Int. Lett. Chem. Phys. Astron. 2013, 19, 26–36. [Google Scholar] [CrossRef] [Green Version]
  94. Karbassi, M.; Zarrintaj, P.; Ghafarinazari, A.; Saeb, M.R.; Mohammadi, M.R.; Yazdanpanah, A.; Rajadas, J.; Mozafari, M. Microemulsion-based synthesis of a visible-light-responsive Si-doped TiO2 photocatalyst and its photodegradation efficiency potential. Mater. Chem. Phys. 2018, 220, 374–382. [Google Scholar] [CrossRef]
  95. Li, P.; Wang, L.; Liu, H.; Li, R.; Xue, M.; Zhu, G. Facile sol-gel foaming synthesized nano foam Bi2Mo3O12 as novel photocatalysts for Microcystis aeruginosa treatment. Mater. Res. Bull. 2018, 107, 8–13. [Google Scholar] [CrossRef]
  96. Król, A.; Pomastowski, P.; Rafińska, K.; Railean-Plugaru, V.; Buszewski, B. Zinc oxide nanoparticles: Synthesis, antiseptic activity and toxicity mechanism. Adv. Colloid Interface Sci. 2017, 249, 37–52. [Google Scholar] [CrossRef] [PubMed]
  97. Warner, J.H.; Schäffel, F.; Bachmatiuk, A.; Rümmeli, M.H. Methods for Obtaining Graphene. In Graphene; Elsevier: Amsterdam, The Netherlands, 2013; pp. 129–228. [Google Scholar]
  98. Taylor, A.C. Advances in nanoparticle reinforcement in structural adhesives. In Advances in Structural Adhesive Bonding; Elsevier: Amsterdam, The Netherlands, 2010; pp. 151–182. [Google Scholar]
  99. Cheaburu-Yilmaz, C.N.; Karasulu, H.Y.; Yilmaz, O. Nanoscaled Dispersed Systems Used in Drug-Delivery Applications; Elsevier: Philadelphia, PA, USA, 2018. [Google Scholar]
  100. Tian, C.; Zhang, Q.; Wu, A.; Jiang, M.; Liang, Z.; Jiang, B.; Fu, H. Cost-effective large-scale synthesis of ZnO photocatalyst with excellent performance for dye photodegradation. Chem. Commun. 2012, 48, 2858–2860. [Google Scholar] [CrossRef] [PubMed]
  101. Bhardwaj, N.; Kundu, S.C. Electrospinning: A fascinating fiber fabrication technique. Biotechnol. Adv. 2010, 28, 325–347. [Google Scholar] [CrossRef] [PubMed]
  102. Sekar, A.D.; Muthukumar, H.; Chandrasekaran, N.I.; Matheswaran, M. Photocatalytic degradation of naphthalene using calcined Fe–ZnO/ PVA nanofibers. Chemosphere 2018, 205, 610–617. [Google Scholar] [CrossRef] [PubMed]
  103. Rao, B.G.; Mukherjee, D.; Reddy, B.M. Novel Approaches for Preparation of Nanoparticles. In Nanostructures for Novel Therapy; Elsevier: Amsterdam, The Netherlands, 2017; pp. 1–36. [Google Scholar]
  104. Mahdavi, R.; Ashraf Talesh, S.S. The effect of ultrasonic irradiation on the structure, morphology and photocatalytic performance of ZnO nanoparticles by sol-gel method. Ultrason. Sonochem. 2017, 39, 504–510. [Google Scholar] [CrossRef] [PubMed]
  105. Fang, K.; Shi, L.; Yao, L.; Cui, L. Synthesis of novel magnetically separable Fe3O4/Bi12O17Cl2 photocatalyst with boosted visible-light photocatalytic activity. Mater. Res. Bull. 2020, 129, 110888. [Google Scholar] [CrossRef]
  106. Salman, S.R. Chemical Reactions Studied by Electronic Spectroscopy. In Encyclopedia of Spectroscopy and Spectrometry; Elsevier: Amsterdam, The Netherlands, 1999; pp. 246–252. [Google Scholar]
  107. Zhang, Y.; Li, G.; Yang, X.; Yang, H.; Lu, Z.; Chen, R. Monoclinic BiVO4 micro-/nanostructures: Microwave and ultrasonic wave combined synthesis and their visible-light photocatalytic activities. J. Alloys Compd. 2013, 551, 544–550. [Google Scholar] [CrossRef]
  108. Bharathi, P.; Harish, S.; Archana, J.; Navaneethan, M.; Ponnusamy, S.; Muthamizhchelvan, C.; Shimomura, M.; Hayakawa, Y. Enhanced charge transfer and separation of hierarchical CuO/ZnO composites: The synergistic effect of photocatalysis for the mineralization of organic pollutant in water. Appl. Surf. Sci. 2019, 484, 884–891. [Google Scholar] [CrossRef]
  109. Wang, X.; Zhou, J.; Zhao, S.; Chen, X.; Yu, Y. Synergistic effect of adsorption and visible-light photocatalysis for organic pollutant removal over BiVO4/carbon sphere nanocomposites. Appl. Surf. Sci. 2018, 453, 394–404. [Google Scholar] [CrossRef]
  110. Wei, X.N.; Ou, C.L.; Guan, X.X.; Peng, Z.K.; Zheng, X.C. Facile assembly of CdS-reduced graphene oxide heterojunction with enhanced elimination performance for organic pollutants in wastewater. Appl. Surf. Sci. 2019, 469, 666–673. [Google Scholar] [CrossRef]
  111. Li, H.; Chen, Y.; Zhou, W.; Gao, H.; Tian, G. Tuning in BiVO4/Bi4V2O10 porous heterophase nanospheres for synergistic photocatalytic degradation of organic pollutants. Appl. Surf. Sci. 2019, 470, 631–638. [Google Scholar] [CrossRef]
  112. Li, K.; Zhong, Y.; Luo, S.; Deng, W. Fabrication of powder and modular H3PW12O40/Ag3PO4 composites: Novel visible-light photocatalysts for ultra-fast degradation of organic pollutants in water. Appl. Catal. B Environ. 2020, 278, 119313. [Google Scholar] [CrossRef]
  113. Deng, Y.; Xiao, Y.; Zhou, Y.; Zeng, T.; Xing, M.; Zhang, J. A structural engineering-inspired CdS based composite for photocatalytic remediation of organic pollutant and hexavalent chromium. Catal. Today 2019, 335, 101–109. [Google Scholar] [CrossRef]
  114. Mbiri, A.; Taffa, D.H.; Gatebe, E.; Wark, M. Zirconium doped mesoporous TiO2 multilayer thin films: Influence of the zirconium content on the photodegradation of organic pollutants. Catal. Today 2019, 328, 71–78. [Google Scholar] [CrossRef]
  115. Adhikari, S.; Sarath, C.K.; Kim, D.H.; Madras, G.; Sarkar, D. Understanding the morphological effects of WO3 photocatalysts for the degradation of organic pollutants. Adv. Powder Technol. 2018, 29, 1591–1600. [Google Scholar] [CrossRef]
  116. Bai, X.; Du, Y.; Hu, X.; He, Y.; He, C.; Liu, E.; Fan, J. Synergy removal of Cr(VI) and organic pollutants over RP-MoS2/rGO photocatalyst. Appl. Catal. B Environ. 2018, 239, 204–213. [Google Scholar] [CrossRef]
  117. Haque, M.M.; Muneer, M. Photodegradation of norfloxacin in aqueous suspensions of titanium dioxide. J. Hazard. Mater. 2007, 145, 51–57. [Google Scholar] [CrossRef] [PubMed]
  118. Xiao, J.; Xie, Y.; Cao, H. Organic pollutants removal in wastewater by heterogeneous photocatalytic ozonation. Chemosphere 2015, 121, 1–17. [Google Scholar] [CrossRef]
  119. Subash, B.; Krishnakumar, B.; Swaminathan, M.; Shanthi, M. Highly active Zr co-doped Ag-ZnO photocatalyst for the mineralization of Acid Black 1 under UV-A light illumination. Mater. Chem. Phys. 2013, 141, 114–120. [Google Scholar] [CrossRef]
  120. Daneshvar, N.; Rabbani, M.; Modirshahla, N.; Behnajady, M.A. Kinetic modeling of photocatalytic degradation of Acid Red 27 in UV/TiO2 process. J. Photochem. Photobiol. A Chem. 2004, 168, 39–45. [Google Scholar] [CrossRef]
  121. Tekin, D.; Saygi, B. Photoelectrocatalytic decomposition of Acid Black 1 dye using TiO2 nanotubes. J. Environ. Chem. Eng. 2013, 1, 1057–1061. [Google Scholar] [CrossRef]
  122. Bhatia, V.; Malekshoar, G.; Dhir, A.; Ray, A.K. Enhanced photocatalytic degradation of atenolol using graphene TiO2 composite. J. Photochem. Photobiol. A Chem. 2017, 332, 182–187. [Google Scholar] [CrossRef]
  123. Tambat, S.; Umale, S.; Sontakke, S. Photocatalytic degradation of Milling Yellow dye using sol-gel synthesized CeO2. Mater. Res. Bull. 2016, 76, 466–472. [Google Scholar] [CrossRef]
  124. Ateş, S.; Baran, E.; Yazıcı, B. Fabrication of Al2O3 nanopores/SnO2 and its application in photocatalytic degradation under UV irradiation. Mater. Chem. Phys. 2018, 214, 17–27. [Google Scholar] [CrossRef]
  125. Taddesse, A.M.; Alemu, M.; Kebede, T. Enhanced photocatalytic activity of p-n-n heterojunctions ternary composite Cu2O/ZnO/Ag3PO4 under visible light irradiation. J. Environ. Chem. Eng. 2020, 8, 104356. [Google Scholar] [CrossRef]
  126. Zhang, J.Y.; Mei, J.Y.; Yi, S.S.; Guan, X.X. Constructing of Z-scheme 3D g-C3N4-ZnO@graphene aerogel heterojunctions for high-efficient adsorption and photodegradation of organic pollutants. Appl. Surf. Sci. 2019, 492, 808–817. [Google Scholar] [CrossRef]
  127. Tayebee, R.; Esmaeili, E.; Maleki, B.; Khoshniat, A.; Chahkandi, M.; Mollania, N. Photodegradation of methylene blue and some emerging pharmaceutical micropollutants with an aqueous suspension of WZnO-NH2@H3PW12O40 nanocomposite. J. Mol. Liq. 2020, 317, 113928. [Google Scholar] [CrossRef]
  128. Mahat, M.M.; Aris, A.H.M.; Jais, U.S.; Yahya, M.F.Z.R.; Ramli, R.; Bonnia, N.N.; Mamat, M.T. A preliminary study on microbiologically influenced corrosion (MIC) of mild steel by Pseudomonas aeruginosa by using infinite focus microscope (IFM). AIP Conf. Proceed. J. 2012, 1455, 117–123. [Google Scholar]
Figure 1. Effect of catalyst on energy diagram profile.
Figure 1. Effect of catalyst on energy diagram profile.
Catalysts 10 01260 g001
Figure 2. The illustration of catalytic and photocatalytic reaction processes.
Figure 2. The illustration of catalytic and photocatalytic reaction processes.
Catalysts 10 01260 g002
Figure 3. The basic mechanism of photocatalysis [23].
Figure 3. The basic mechanism of photocatalysis [23].
Catalysts 10 01260 g003
Figure 4. Example of doping as an electron trapping site of photocatalyst [50].
Figure 4. Example of doping as an electron trapping site of photocatalyst [50].
Catalysts 10 01260 g004
Figure 5. Illustrated (a) type of band alignment between two semiconductors in a heterojunction [51], (b) binary [52], (c) ternary [53], and (d) quaternary coupling heterojunctions [54].
Figure 5. Illustrated (a) type of band alignment between two semiconductors in a heterojunction [51], (b) binary [52], (c) ternary [53], and (d) quaternary coupling heterojunctions [54].
Catalysts 10 01260 g005
Figure 6. Example of recent studies on supported material photocatalyst such as (a) graphene [65], (b) carbon nanotube [68], and (c) aluminum oxide [70].
Figure 6. Example of recent studies on supported material photocatalyst such as (a) graphene [65], (b) carbon nanotube [68], and (c) aluminum oxide [70].
Catalysts 10 01260 g006aCatalysts 10 01260 g006b
Figure 7. Illustration of sonication effect on the shape and size of the catalyst [99].
Figure 7. Illustration of sonication effect on the shape and size of the catalyst [99].
Catalysts 10 01260 g007
Figure 8. SEM images of (a) nanosheet Ce(MoO4)2 [80] and (b) g-C3N4/SnO photocatalyst [81].
Figure 8. SEM images of (a) nanosheet Ce(MoO4)2 [80] and (b) g-C3N4/SnO photocatalyst [81].
Catalysts 10 01260 g008
Figure 9. The solid-state technique for the preparation of ZnO.
Figure 9. The solid-state technique for the preparation of ZnO.
Catalysts 10 01260 g009
Figure 10. SEM images of (a) Mg doped ZnO-Al2O3 photocatalyst [43] and (b) cone-like ZnO shape [82].
Figure 10. SEM images of (a) Mg doped ZnO-Al2O3 photocatalyst [43] and (b) cone-like ZnO shape [82].
Catalysts 10 01260 g010
Figure 11. Visual of the formation of nanofibers by high voltage and spinneret injection of the solution [101].
Figure 11. Visual of the formation of nanofibers by high voltage and spinneret injection of the solution [101].
Catalysts 10 01260 g011
Figure 12. SEM images of (a) Fe-ZnO/PVA nanofibers [102]; (b) the wrinkled Ag/LaFeO3 nanofibers [87].
Figure 12. SEM images of (a) Fe-ZnO/PVA nanofibers [102]; (b) the wrinkled Ag/LaFeO3 nanofibers [87].
Catalysts 10 01260 g012
Figure 13. Illustration of ZnO photocatalyst preparation using a sol-gel technique.
Figure 13. Illustration of ZnO photocatalyst preparation using a sol-gel technique.
Catalysts 10 01260 g013
Figure 14. SEM images of (a) Bi2Mo3O12 photocatalyst [95] (b) uniformly size and shape of Nd2Mo3O9 [36].
Figure 14. SEM images of (a) Bi2Mo3O12 photocatalyst [95] (b) uniformly size and shape of Nd2Mo3O9 [36].
Catalysts 10 01260 g014
Figure 15. Illustration of the preparation of Fe3O4 by the co-precipitation technique.
Figure 15. Illustration of the preparation of Fe3O4 by the co-precipitation technique.
Catalysts 10 01260 g015
Figure 16. SEM images of (a) flower-like ZnO [91] (b) the cubic shaped of porous hollow ZnFe2O4 [34].
Figure 16. SEM images of (a) flower-like ZnO [91] (b) the cubic shaped of porous hollow ZnFe2O4 [34].
Catalysts 10 01260 g016
Figure 17. Different types of hybridization photocatalyst and the key factor to focus.
Figure 17. Different types of hybridization photocatalyst and the key factor to focus.
Catalysts 10 01260 g017
Figure 18. Important remarks: Parameter effects on photocatalytic degradation efficiency.
Figure 18. Important remarks: Parameter effects on photocatalytic degradation efficiency.
Catalysts 10 01260 g018
Table 1. Research on photocatalysis; parameter condition and results reported from 2016 to 2018.
Table 1. Research on photocatalysis; parameter condition and results reported from 2016 to 2018.
Type of Organic PollutantTreatPhotocatalystLight SourceBand GapDegradation Rate
(%)
Ref.
AnalgesicsAcetaminophenCdS
sub-microspheres
50 W LED visible light λ = 455 nm2.16 eVAcetaminophen: 85%
Levofloxacin: 70%
in 240 min
[29]
ParacetamolTiO2-graphite compositesUV lamp 8 W
λ = 365 nm
3.24 eVUV: 100%
in 120 min
TOC: 88%
in 180 min
[30]
AntibioticsDoxycyclineBiOBr/FeWO4300 W Xenon lamp with a 400-nm cutoff filter2.46 eV90.4% in
60 min
[31]
LevofloxacinCdS
sub-microspheres
50 W LED visible light λ = 455 nm2.16 eVLevofloxacin: 70%
in 240 min
[29]
Tetracycline hydrochlorideCdTe/TiO2400 W halogen lamp equipped with a cutoff filter (λ > 400 nm)1.39 eV78%
in 30 min
[32]
tetracyclineFe-based metal–organic frameworks300 W Xenon lamp visible lamp
λ > 420 nm
1.88 eV96.6%
in 3 h
[33]
Tetracycline hydrochlorideZnFe2O4
porous hollow cube
300 W Xe
Lamp equipped with 350 nm–780 nm reflection filter and 420 nm cutoff filter (irradiation wavelength of 420 nm–780 nm)
1.5 eV84.08%
in 60 min
[34]
TetracyclineZnWO4−x nanorodsUV lamp mercury 300 W
Xenon
lamp 300 W
3.1 eV91%
in 80 min
[35]
NitrofurantoinNd2Mo3O9300 W tungsten incandescent lamp
lamp intensity is 150 mW/cm2
2.82 eV99%
in 45 min
[36]
DyesRhodamine Bpolycaprolactone/TiO2 nanofibrous25 W of 254 nm UV light-100% in
300 min
[37]
Methyl orangeCu-doped ZnO/Al2O3Visible light
400 W high-pressure mercury-vapor lamp
λ = 546.8 nm
2.18 eV100%
in 50 min
[28]
Orange GSepiolite-TiO2
nanocomposites
300 W Xe lamp-98.8%
in 150 min
[38]
Nitroblue tetrazolium
Methylene blue
Ternary g-C3N4/Al2O3/ZnOVisible light
300 W xenon lamp with λ > 420 nm cut-off filter
ZnO 3.20 eV
Al2O3 4.86 eV
g-C3N4 2.76 eV
85%
in 50 min
[39]
Methyl orangeTungsten doped Al2O3/ZnO coating aluminumSimulate solar irradiation 300 W-95%
in 10 h
[40]
HerbicideatrazineCu-BiOClMercury UV lamp
Λ = 254 nm
3.0 eV35%
in 30 min
[41]
PesticidenaphthaleneZnO254 nm irradiation under 50 W mercury lamp2.98 eV70%
in 2 h
[42]
StimulantCaffeineMg doped ZnO-Al2O3UV mercury lamp 400 W-89.18%
in 70 min
[43]
Table 2. Effect of different loading of dopant on the adsorption and photocatalytic degradation.
Table 2. Effect of different loading of dopant on the adsorption and photocatalytic degradation.
CatalystsQuantity of DopantMorphological CharacterizationAdsorption Capacity
(wt.% Dopant: % Adsorption)
Photocatalytic Degradation
(wt.% Dopant: % Degradation)
Ref.
Cu-doped ZnO/Al2O3Minimum dopant loadingUndoped ZnO/Al2O3 observed as spherical morphology-0.0 wt.%: 6.4%
2.5 wt.%: 39.5%
5.0 wt.%: ~93.0%
[28]
Optimum dopant loadingAfter doping, the surface became lamellar morphology and XRD peaks relative intensity showed slightly decreased-7.5 wt.%: 100%
Maximum dopant loading--10.0 wt.%: ~93.0%
Mg-doped ZnO-Al2O3Minimum dopant loadingCrystallite size 21 nm;
Observable porosity on surface
0.0 wt.%: 7.0%0.0 wt.%: 89.2%[43]
Optimum dopant loadingCrystallite size 8 nm;
The surface increase in grain size
1.0 wt.%: 11.1%1.0 wt.%: 98.9%
Maximum dopant loadingCrystallite size ≤ 35 nm3.0 wt.%: 6.7%
5.0 wt.%: 1.2%
3.0 wt.%: <98.9%
5.0 wt.%: <98.9%
Co-doped Zr6O4(OH4)BDC12Minimum dopant loadingAgglomerated cubic morphology with diameter of 230 nm; surface area was 584 m2g−10.0 wt.%: 9.9%0.0 wt.%: <78.5%[48]
Optimum dopant loadingDispersive and uniform cubic with diameter of 170 nm; increase surface area to 815 m2g−11.0 wt.%: 68.1%1.0 wt.%: 78.5%
Maximum dopant loadingThe higher amount of Co doped, the higher surface area observed2.0 wt.%: 61.3%
4.0 wt.%: 58.6%
8.0 wt.%: 55.4%
2.0 wt.%: <78.5%
4.0 wt.%: <78.5%
8.0 wt.%: <78.5%
Ag-doped ZnSMinimum dopant loadingAverage diameter was 3.0–5.0 nm; surface area was 78 m2g−1-No dopant: 79.7%[49]
Optimum dopant loadingAverage diameter was 3.0–5.3 nm; increase surface area to 89 m2g−1-With dopant: 92.8%
Maximum dopant loading---
Table 3. Published preparation methods for photocatalysts.
Table 3. Published preparation methods for photocatalysts.
Type of MethodMethodPhotocatalystReference
Solid-based methodSonicationCu-doped ZnO/Al2O3[28]
Ce(MoO4)2[80]
SnO/g-C3N4[81]
Solid-stateMg doped ZnO-Al2O3[43]
ZnO[82]
CuO/Al2O3/TiO2[83]
Thermal evaporationSnO2[84]
ZnO[85]
ZnO[86]
Solution-based methodHydrothermalSepiolite-TiO2[38]
CdS[29]
CdTe/TiO2[32]
ElectrospinningZnO[42]
polycaprolactone/TiO2[37]
Ag/LaFeO3[87]
ImpregnationMg-ZnO/Al2O3[43]
Cu-BiOCl[41]
Se-ZnS[88]
PrecipitationTiO2-graphite[30]
γ-Al2O3[89]
ZnWO4[90]
Co-precipitationg-C3N4/Al2O3/ZnO[39]
Mg doped ZnO-Al2O3[43]
ZnFe2O4[34]
Cu-doped ZnO/Al2O3[28]
ZnO[91]
MicroemulsionFe2O3[92]
ZnO[93]
Si doped TiO2[94]
Sol-gelNd2Mo3O9[36]
TiO2[94]
Bi2Mo3O12[95]
Table 4. Example of a current study of the photocatalyst in degradation from a different source of organic pollutants.
Table 4. Example of a current study of the photocatalyst in degradation from a different source of organic pollutants.
PhotocatalystType of Light UsedSource of PollutionCondition SetDegradationRef.
Catalyst LoadingInitial ConcentrationReaction TimepH
BiVO4/CHCOO(BiO)Visible light using 300 W Xenon lampSulfamethoxazole1 g/L10 mg/L5 h685%[107]
Bisphenol A1 g/L10 mg/L5 h699%
4-aminoantipyrine1 g/L10 mg/L5 h646%
Ibuprofen1 g/L10 mg/L5 h665%
CuO/ZnO500 W visible lampMethylene blue1 g/L10 mg/L25 min-96.57%[108]
BiVO4/carbon350 W Xenon lamp with 400 nm cut off filterMethylene blue1.0 g/L0.0001 mol/L180 min-95%[109]
Rhodamine B1.0 g/L1 × 10−5 mol/L180 min-80%
Phenol1.0 g/L5 mg/L5 h-50.13%
CdS-reduced graphene oxide300 W Xenon lamp with UV cut off filterRhodamine B0.4 g/L20 mg/L60 min-97.2%[110]
Acid chrome blue K0.4 g/L20 mg/L60 min-65.7%
BiVO4/Bi4V2O10300 W Xenon lamp with 400 nm cut off filterRhodamine B1 g/L20 mg/L15 min-100%[111]
Methylene blue250 g/L20 mg/L60 min-75%
Phenol1 g/L30 mg/L60 min-95%
H3PW12O40/Ag3PO4300 W Xe lamp λ > 420 nm 4-fluorophenol3 g/L10 mg/L720 s7100%[112]
Methyl orange3 g/L10 mg/L720 s7100%
ZnS:Mn/MWCNTLow-pressure mercury lampAR18 dye0.1 g/L20 mg/L180 min-70%[68]
CdS/TiO2300 W xenon lamp with 420 nm cut off filterHexavalent chromium2 g/L10 mg/L5 h3.5100%[113]
Phenol1 g/L10 mg/L3 h-78%
Rhodamine B1 g/L10 mg/L60 min-83%
Zr/TiO2300 W Xenon lamp with 320 nm cut off filterChloridazon0.1 g/L0.005 mM4 h5100%[114]
Phenol0.1 g/L0.001 mM4 h5.4590%
4-chlorophenol0.1 g/L0.001 mM4 h5.5395%
WO3400 W metal halide lamp. Light intensity ≈ 86,800 lxRhodamine B1 g/L20 mg/L3 h-95%[115]
RP-MoS2/rGO300 W Xenon lamp 420 nm cut off filterRhodamine B0.4 g/L20 mg/L30 min-99.3%[116]
Hexavalent chromium0.4 g/L40 mg/L30 min-98%
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Sinar Mashuri, S.I.; Ibrahim, M.L.; Kasim, M.F.; Mastuli, M.S.; Rashid, U.; Abdullah, A.H.; Islam, A.; Asikin Mijan, N.; Tan, Y.H.; Mansir, N.; et al. Photocatalysis for Organic Wastewater Treatment: From the Basis to Current Challenges for Society. Catalysts 2020, 10, 1260. https://doi.org/10.3390/catal10111260

AMA Style

Sinar Mashuri SI, Ibrahim ML, Kasim MF, Mastuli MS, Rashid U, Abdullah AH, Islam A, Asikin Mijan N, Tan YH, Mansir N, et al. Photocatalysis for Organic Wastewater Treatment: From the Basis to Current Challenges for Society. Catalysts. 2020; 10(11):1260. https://doi.org/10.3390/catal10111260

Chicago/Turabian Style

Sinar Mashuri, Salma Izati, Mohd Lokman Ibrahim, Muhd Firdaus Kasim, Mohd Sufri Mastuli, Umer Rashid, Abdul Halim Abdullah, Aminul Islam, Nurul Asikin Mijan, Yie Hua Tan, Nasar Mansir, and et al. 2020. "Photocatalysis for Organic Wastewater Treatment: From the Basis to Current Challenges for Society" Catalysts 10, no. 11: 1260. https://doi.org/10.3390/catal10111260

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