Next Article in Journal
Mono-, Bis-, and Tris-Chelate Zn(II) Complexes with Imidazo[1,5-a]pyridine: Luminescence and Structural Dependence
Next Article in Special Issue
Applications of Catalytic Nanomaterials in Energy and Environment
Previous Article in Journal
Preparation of Porous Hydroxyethyl Cellulose Materials to Utilize Lactic Acid with Vacuum-Assisted Process
Previous Article in Special Issue
Comparative Studies of g-C3N4 and C3N3S3 Organic Semiconductors—Synthesis, Properties, and Application in the Catalytic Oxygen Reduction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Application of Photocatalysis and Sonocatalysis for Treatment of Organic Dye Wastewater and the Synergistic Effect of Ultrasound and Light

MOE Key Laboratory for Earth Surface Processes, College of Urban and Environmental Sciences, Peking University, Beijing 100871, China
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(9), 3706; https://doi.org/10.3390/molecules28093706
Submission received: 23 March 2023 / Revised: 12 April 2023 / Accepted: 18 April 2023 / Published: 25 April 2023
(This article belongs to the Special Issue Catalytic Nanomaterials: Energy and Environment)

Abstract

:
Organic dyes play vital roles in the textile industry, while the discharge of organic dye wastewater in the production and utilization of dyes has caused significant damage to the aquatic ecosystem. This review aims to summarize the mechanisms of photocatalysis, sonocatalysis, and sonophotocatalysis in the treatment of organic dye wastewater and the recent advances in catalyst development, with a focus on the synergistic effect of ultrasound and light in the catalytic degradation of organic dyes. The performance of TiO2-based catalysts for organic dye degradation in photocatalytic, sonocatalytic, and sonophotocatalytic systems is compared. With significant synergistic effect of ultrasound and light, sonophotocatalysis generally performs much better than sonocatalysis or photocatalysis alone in pollutant degradation, yet it has a much higher energy requirement. Future research directions are proposed to expand the fundamental knowledge on the sonophotocatalysis process and to enhance its practical application in degrading organic dyes in wastewater.

Graphical Abstract

1. Introduction

In recent years, the textile industry has played a vital role in the global economy, but it is also a major contributor to environmental pollution, particularly in terms of organic dye wastewater [1,2,3,4,5]. The discharge of organic dye wastewater from textile production and utilization can cause significant damage to the aquatic ecosystem, and therefore, it is imperative to develop effective treatment methods to degrade these pollutants [6,7,8]. Among the various technologies geared toward the treatment of organic dye wastewater, photocatalysis, sonocatalysis, and sonophotocatalysis have received increasing attention owing to their high efficiency and potential for large-scale industrial applications [9,10,11,12,13].
Photocatalysis is a well-established technology for the treatment of organic pollutants in wastewater [14,15,16,17]. Qutub et al. investigated CdS/TiO2 nanocomposites for photocatalytic degradation of organic pollutants in wastewater [18]. The results showed that CdS-TiO2 nanocomposites exhibited the highest photocatalytic activity in the degradation of AB-29 dye, with a degradation efficiency of 84%, compared to 68% and 9% achieved by CdS and TiO2 under comparable conditions, respectively. The enhanced photocatalytic performance of CdS-TiO2 was attributed to reduced charge carrier recombination, improved charge separation, and expansion of the response of TiO2 to visible light. In photocatalytic systems, semiconductor photocatalysts, such as TiO2, are irradiated with light, generating electron-hole pairs that react with water or oxygen to form reactive species, e.g., hydroxyl radicals. These reactive species can later degrade the organic pollutants into harmless products [19,20,21,22]. Sonocatalysis, on the other hand, utilizes ultrasonic wave to generate cavitation bubbles in the solution, which collapse and produce high-energy conditions that can promote chemical reactions. Wang et al. developed a recyclable WO3/NiFe2O4/BiOBr (WNB) composite with dual Z-scheme heterojunction for the degradation of levofloxacin (LEV) in aqueous solution [23]. The WNB composite showed the highest removal efficiency (97.97%) for LEV within 75 min under ultrasonic irradiation. The ternary composite comprises three different semiconductors suitable for harvesting full-spectrum light. The combination of sonocatalysis and photocatalysis, known as sonophotocatalysis, can further improve the efficiency of both processes, as the cavitation bubbles can create local “hot spots” that increase the photocatalytic activity for the catalyst [24,25,26,27,28]. Despite the promising results of sonophotocatalysis in pollutant degradation, there are still significant challenges that need to be addressed in order to enhance its performance for the treatment of organic dye wastewater [29]. For example, the high energy requirement of sonophotocatalysis limits its practical application, as it has high electricity consumption for generating the ultrasonic wave and producing the light [30,31,32,33,34,35,36]. In addition, mechanistic understanding on the synergistic effect of ultrasound and light in sonophotocatalysis is still not lacking, and more studies are required to clarify the underlying mechanism for optimization of the process [37,38]. Wang et al. synthesized Fe3O4@SiO2/PAEDTC@MIL-101(Fe), a mesoporous composite with a core-shell structure, and evaluated its sonophotocatalytic performance in degrading acid red 14 (AR14) [39]. The results showed that Fe3O4@SiO2/PAEDTC@MIL-101 (Fe)/UV/US exhibited excellent activity in the removal of AR14 and total organic carbon.
In this review, we aim to present an overview on the mechanisms of photocatalysis, sonocatalysis, and sonophotocatalysis in the treatment of organic dye wastewater [40,41]. We compare the performance of TiO2-based catalysts in photocatalytic, sonocatalytic, and sonophotocatalytic systems, with a focus on the synergistic effect of ultrasound and light in the catalytic degradation of organic dyes [42]. We further discuss the recent advances in catalyst development for sonophotocatalysis and point out the potential risks associated with the sonophotocatalytic process, such as the generation of toxic byproducts and the potential releases of nanoparticles into the environment [43,44,45,46,47,48]. It is essential to monitor the reaction products and assess their toxicity to ensure that the sonophotocatalytic process is safe for both the environment and human health. Finally, we propose future research directions to expand the fundamental knowledge on the sonophotocatalysis process and enhance its practical application in degrading organic pollutants [49,50,51,52,53]. The information presented in this review can provide valuable insights into the mechanisms and performance of photocatalysis, sonocatalysis, and sonophotocatalysis, and contribute to the development of more efficient and cost-effective treatment methods for organic dye wastewater. This review will be of great interest to researchers and practitioners in the field of environmental science and engineering, especially those involved in the development of sustainable wastewater treatment technologies.

2. Sonocatalytic and Photocatalytic Mechanisms

2.1. Sonocatalytic Mechanism

The sonocatalytic process is believed to be predominantly based on the “hot spots” and “sonoluminescence” that originate from the ultrasonic cavitation phenomenon [54]. Ultrasonic wave of a specific frequency and intensity can produce numerous small bubbles in liquids [55]. These minute bubbles trigger various physical and chemical transformations during their formation, oscillation, expansion, contraction, and ultimate collapse [56,57,58]. Figure 1 scehematically depicts the phenomenon of ultrasonic cavitation. It promotes the production of light with a range of wavelengths, called “sonoluminescence,” and a large number of localized “hot spots” with very high temperatures (up to ~5000 K) and pressures (up to ~1000 atm) [59,60]. These localized “hot spots” can cause pyrolysis of H2O molecules, producing hydroxyl radicals (•OH) [61], which can effectively oxidize organic pollutants and even mineralize them into CO2 and H2O [62,63,64,65,66].
In general, the sonolytic removal of organic pollutants involves oxidation through both pyrolysis and free radical attack [67]. However, due to the significant energy loss that occurs during thermal dissipation (exceeding 50%), radpid degradation often cannot occur when relying solely on ultrasound [68]. In recent years, the use of ultrasound in the presence of appropriate catalysts, known as sonocatalysis, has been increasingly used to degrade organic pollutants due to its numerous benefits, including convenient handling and low cost, as well as environmental friendliness [69]. Sonocatalytic degradation involves the use of a sonocatalyst to create additional active sites for the cavitation effect, leading to the formation of greater numbers with highly reactive radicals [70]. In general, these radicals could recombine to form H2O, •OH, H2O2, and •O2 in water [71,72]:
H2O + ))) → •H + •OH
•H + •OH → H2O2
O2 + •H → •HO2
2•OH → H2O2
2•HO2 → O2 + H2O2
H2O + •OH → H2O2 + •H
Sonocatalysis is a crucial technology in the degradation of pollutants, and radicals play an essential role in this process [73]. These radicals can initiate chain reactions that lead to the degradation of pollutants. To enhance the efficiency of the sonocatalytic process, it is essential to understand the underlying mechanisms of sonocatalysis [74]. Figure 2 depicts the major processes involved in sonocatalysis elucidated by extensive studies conducted in the field [75,76].

2.1.1. Heterogeneous Nucleation Mechanism

Semiconductor particles have been observed to induce preferential formation of nuclei at solid surfaces or phase boundaries, leading to increased formation of cavitation bubbles and free radicals, such as •OH [54]. The phenomenon of heterogeneous nucleation has been found to be more applicable than homogeneous cavitation in sonocatalysis [56]. This can be attributed to the fact that the thermodynamic nucleation barriers on interfaces are generally lower than their bulk counterparts, promoting surface nucleation [18]. The relationship between the maximum energy barriers of heterogeneous and homogeneous nucleation processes can be expressed as [57,60]:
Δ G * het = 16 π σ 3 3 P 2   f ( θ ) = Δ G * hom   f ( θ )  
where ΔGhet and ΔGhom are the maximum energy barrier for heterogeneous and homogeneous reactions, respectively, σ is the surface tension of water (J/m2), θ is the contact angle between the liquid and solid, and P is the sum of the partial pressure of the entrapped gas [47].
It is expected that preferential nucleation will transpire on hydrophobic surfaces, notably on solid surfaces [63]. Furthermore, the rate of bubble nucleation at the solid surface can be significantly influenced by many factors. Sonication parameters, such as ultrasonic power, frequency, as well as changes in surface energy, aqueous temperature, and type of absorbed gas, can greatly affect this process [59]. In addition, physicochemical properties of the solid particles, such as roughness, particle size, pore size, and wettability, can play a crucial role in influencing the nucleation rate.

2.1.2. Photo-Excitation Mechanism

Sonoluminescence (SL) is a light-emitting phenomenon caused by the collapse of cavitational bubbles. The light emitted during sonoluminescence has high intensity and covers a wide range of wavelengths, typically between 200 and 700 nm [77]. In the presence of a semiconductor catalyst during ultrasonication, the energy from the light generated can exceed the band gap of the semiconductor, leading to the excitation of electrons from the valence band (VB) to the conduction band (CB) [78]. This process generates holes in the valence band, which are caused by the excited electrons [79]. When photogenerated electron-hole pairs react with dissolved oxygen, they create highly reactive radicals. This process in sonocatalysis is akin to that in photocatalysis [80].

2.1.3. Thermal Excitation Mechanism

The “hot spots” hypothesis proposes that elevated temperatures in a specific area may result in thermal excitation of the semiconductor, causing the formation of electron-hole pairs [81]. This phenomenon has been observed in numerous studies [82], demonstrating that certain semiconductors can be stimulated by high temperatures to generate electron-hole pairs. At room temperature, TiO2 displays low catalytic activity, but its performance improves significantly after being heated to temperatures ranging between 350 and 500 °C [83]. Such enhancement is attributed to the abundant highly oxidative holes that arise due to the thermal excitation of semiconductors [84].

2.2. Photocatalytic Mechanism

The photocatalytic process occurs when a semiconductor catalyst is exposed to light of greater energy than the semiconductor’s bandgap [85], as depicted in Figure 3. When this happens, electrons in the VB may become excited and jump into the CB, forming a hole (h+VB) (Equation (8)). Subsequently, the electron-hole pairs that are generated by the absorption of light recombine together, leading to the emission of energy (Equation (9)) [86]. The poor quantum efficiency of the semiconductor is attributed to this recombination, which leads to low light-to-energy conversion rates [87,88]. If the photogenerated carriers do not recombine, light-generated electron-hole (e, h+) pairs separate and move to the material’s surface, reacting with the adsorbed molecules [89]. When photo-excited electrons come into contact with dissolved oxygen molecules (O2) in an aqueous solution, they can react and form superoxide radical anions (•O2), as indicated by Equation (10) [90]. At the same time, the holes may directly oxidize pollutants or H2O molecules to produce hydroxyl radicals (•OH) (Equation (11)). The reactive radicals generated (•OH, •O2) are highly reactive oxidizing agents [91], andthey may readily mineralize many organic molecules, producing water and carbon dioxide (Equations (12) and (13)).
Semiconductor + hv → eCB + h+VB
eCB + h+VB → energy
e + O2 → •O2
h+ + H2O → •H + •OH
O2 + Pollutant → H2O + CO2
•OH + Pollutant → H2O + CO2

2.3. Comparison of Sonocatalytic and Photocatalytic Mechanisms

Comparison of the mechanisms of sonocatalysis and photocatalysis can help better understand the unique features and advantages of sonocatalysis in promoting efficient and sustainable chemical transformations. The similarity and difference between these two types of mechanisms are detailed below.

2.3.1. Similarity

Semiconductor catalyst plays a vital role in lowering the energy barrier for the formation of cavitation bubbles, which is similar to the way that a traditional catalyst reduces the activation energy of a chemical reaction [92]. This is achieved by providing a surface for the accumulation and stabilization of gas or vapor pockets within the fluid medium, effectively reducing the threshold pressure required for bubble nucleation [93].
Undoubtedly, photocatalysts have the potential to serve as effective sonocatalysts, leveraging the phenomenon of sonoluminescence generated by cavitation [94]. Given their inherent properties and unique chemical compositions, photocatalysts can harness the energy released by cavitation bubbles to enhance catalytic reactions and promote efficient chemical transformations [95].

2.3.2. Difference

The formation of cavitation bubbles is primarily driven by physical processes, involving the rapid formation and collapse of small pockets of gas or vapor within a fluid medium [96]. This can occur due to the changes in pressure and temperature that cause the fluid to reach its boiling point, resulting in the generation of these bubbles [97]. The effect of cavitation can be significant, leading to the erosion of solid surfaces and the generation of shockwave that can have profound impacts on the surrounding environment [98].
Acoustic cavitation is a key phenomenon in sonocatalysis, whereby high-intensity sound wave generates microscopic bubbles in a liquid medium [99]. During the cavitation process, these bubbles release energy in the form of heat, shockwave, and free radicals, which can induce chemical reactions in the solution [100]. As the bubbles collapse, they generate extremely high temperatures and pressures in localized regions of the solution [101]. The sudden and intense energy release can result in large increases in temperature, which can accelerate the rate of chemical reactions in the solution [102]. Moreover, the high temperatures generated by acoustic cavitation can lead to thermal excitation of the catalyst, thereby promoting the generation of reactive species, such as electron-hole pairs [103]. This, in turn, can lead to enhanced catalytic activity and selectivity in sonocatalysis.

3. Sonophotocatalytic Process

3.1. Sonophotocatalytic Mechanism

Sonophotocatalysis is essentially a combination of light, ultrasound, and catalyst that accelerates the degradation rates of organic pollutants via increasing the production of active radicals [104,105,106]. The highly efficient degradation of organic pollutants in sonophotocatalytic process is principally based on the synergistic effect of sonocatalysis and photocatalysis [107,108]. Figure 4 depicts the mechanism for the synergistic effect of photocatalysis and sonocatalysis. The key advantage of combining these two technologies is the greater number of cavitation bubbles generated via ultrasound, and more radicals generated via electron-hole pair separation in semiconductor photocatalysts. In addition, ultrasound continuously cleans the surface of the photocatalyst, which helps maintain the catalyst activity for extended periods. The combination of these two technologies can degrade hydrophobic and hydrophilic organic pollutants [109].
The sonophotocatalytic degradation of organic pollutants include the phenomena of both sonocatalysis and photocatalysis. Upon irradiation of ultrasonic wave, sonoluminescence and “hot spot” are generated due to cavitation in the aqueous solution. Moreover, the “hot spots” formed through ultrasonic cavitation may cause the pyrolysis of water molecules in contact with the surface of the sonocatalyst, generating hydroxyl radicals (•OH) and hydrogen radicals (•H) (Equations (14)–(19)) [110]. Subsequently, the light generated with a wide range of wavelength from sonoluminescence can excite the catalyst, facilitating charge carriers’ formation and the generation of electron-hole pairs in the CB and VB (Equation (20)). Additionally, irradiation of the catalyst’s surface with light increases the generation of electron-hole pairs and active radical species during sonophotocatalysis (Equation (21)) [111,112,113]. In the VB, holes react with water molecules adsorbed on the catalyst surface to generate •OH (Equation (22)). At the same time, electrons generated in the CB react with dissolved oxygen to generate •O2, •OH, and H2O2 (Equations (23)–(25)). Subsequently, these active species react with organic pollutants to generate different degradation intermediates and even mineralization products (i.e., H2O and CO2) (Equation (26)) [114].
H2O + heat (hot spot) → •OH + •H
H2O + H• → H2 +•OH
•OH + •OH → H2O2
O2 + H• → HO2
HO2 + HO2 → O2 + H2O2
H2O2 + heat (hot spot) → 2•OH
Semiconductor Catalyst + US → h+/VB + e/CB
Semiconductor Catalyst + hν → h+/VB + e/CB
h+/VB + H2O → •OH
e/CB + O2 → •O2
2H2O + 2•O2 → 2OH- + H2O2 + O2
2•O2 + H2O2 → OH- + •OH + O2
Organic pollutants + Reactive oxidative species →H2O + CO2

3.2. Summary of the Synergistic Effect during Sonophotocatalytic Process

In order to compare the effects of sonophotocatalysis with those of separate processes (sonocatalysis and photocatalysis), it is necessary to assess the synergistic contribution to the elimination of organic pollutants during the degradation process by sonophotocatalysis. The synergistic effect of a sonophotocatalysis process can be assessed using the synergistic index. This index is calculated as the ratio of the rate constant of sonophotocatalysis to the sum of the rate constants of the individual processes, and is commonly employed to analyze the degree of synergistic enhancement in dye decolorization. The generic expression of the synergy index can be expressed as ([115,116]):
Synergy   Index = k sonophotocatalysis k sonocatalysis + k photocatalysis
where k represents the pseudo-first-order rate constants of the photocatalytic, sonophotocatalytic, and sonocatalytic degradation processes, a synergistic index value of >1 means the efficiency of the sonophotocatalytic degradation is higher than the cumulative value of the individual processes (sonocatalytic or photocatalytic).
The synergistic effect of sonophotocatalysis in organic pollutant degradation has been demonstrated in many studies. Mosleh et al. reported that the pseudo-first-order rate constant for sonophotocatalytic degradation of trypan blue was 26.33 × 10−2 min−1, while the sum of the rate constants of photocatalysis and sonocatalysis was only 9.88 × 10−2 min−1, resulting in a synergistic index of 2.53 [117]. Babu et al. reported a synergistic index of 3.7 for the sonophotocatalytic degradation of Methyl orange using CuO-TiO2/rGO nanocatalysts [118]. The authors concluded that the high synergy probably resulted from the combined action of hydroxyl radicals generated by the sonolytic and photocatalytic systems. Benomara et al. reported that the pseudo-first-order rate constants for the degradation of methyl violet 2B were 6.8 × 10−3 for sonocatalysis, 22.9 × 10−3 for photocatalysis, and 39.7 × 10−3 min−1 for sonophotocatalysis, demonstrating the significant synergistic effect of sonophotocatalysis [119]. Ahmad et al. investigated the degradation of Rhodamine B (RhB) in photocatalytic, sonocatalytic, and sonophotocatalytic systems, and found that the sonophotocatalytic process exhibiting a higher rate constant compared to the sum of the photocatalytic and sonocatalytic processes [120]. Sonophotocatalytic process was more effective in degrading RhB compared to photocatalytic and sonocatalytic processes due to the presence of more reactive radicals and the increased active surface area of the ZnO/CNT photocatalyst. Togther, these findings highlight the potential of sonophotocatalysis as a promising approach for the efficient degradation of organic dyes in wastewater.
During the sonophotocatalytic process, the combination of ultrasonic wave, light, and photocatalyst can lead to synergistic effect that enhances the degradation of organic pollutants in wastewater. The synergistic effect is attributed to several factors, including the increased production of reactive radicals and the improved mass transfer of the pollutants to the photocatalyst surface. One of the key advantages of sonophotocatalysis is the increased production of reactive radicals, such as •OH, which is highly effective in breaking down organic pollutants. Ultrasonic wave can induce cavitation, which generates high-energy bubbles that collapse and release shockwave and heat, leading to the formation of reactive radicals. Similarly, when a photocatalyst is illuminated with light, electrons are excited, leading to the production of reactive radicals. The combination of ultrasonic wave and light in sonophotocatalysis can lead to a higher production of reactive radicals, as the ultrasonic wave can promote the separation of electron-hole pairs, which are the precursors of reactive radicals, while also enhancing the mass transfer of the pollutants to the photocatalyst surface. Another factor that contributes to the synergistic effect of sonophotocatalysis is the improved mass transfer of the pollutants to the photocatalyst surface. In traditional photocatalysis, the efficiency of pollutant degradation is often limited by the mass transfer of the pollutants from the bulk solution to the photocatalyst surface. The use of ultrasonic wave in sonophotocatalysis can enhance the mass transfer of the pollutants by promoting the formation of micro-scale streams and turbulence, which increase the contact between the pollutants and the photocatalyst surface. In summary, the synergistic effect of sonophotocatalysis in the degradation of organic pollutants can be attributed to the increased production of reactive radicals and the improved mass transfer of the pollutants to the photocatalyst surface.

4. Degradation of Dyes Using TiO2-Based Semiconductor Catalysts

TiO2 has been widely examined among numerous photocatalysts because of its chemical stability, non-toxicity, strong oxidation ability, low cost, high catalytic activity, and photo corrosion resistance. It has been the focus of research in the field of photocatalysis and is one of the most commonly used and most promising photocatalysts [121,122,123]. The photocatalytic activity of anatase TiO2 is limited to ultraviolet light with wavelength shorter than 387 nm due to its wide band gap of 3.23 eV. As the energy of UV light accounts for only 4% of the total energy of sunlight, TiO2 cannot effieiciently utilize sunlight, which seriously limits its application in photocatalysis [18]. In practical applications, researchers have modified TiO2 to enhance its catalytic activity. There are several primary methods for TiO2 modification, such as noble metal deposition, compound semiconductor, ion doping, and photosensitization. The primary objective of modification is to expand the light-absorption of TiO2 to the visible light spectrum and inhibit the recombination of electron-hole pairs [124]. Additionally, the incorporation of other materials into the TiO2 photocatalyst can enhance its performance. For example, graphene oxide (GO) has been used as a support material for TiO2 nanoparticles to form GO-TiO2 composites. The incorporation of GO can improve the adsorption capacity and photocatalytic activity of TiO2 by increasing the specific surface area and promoting the separation of photogenerated electron-hole pairs [125]. GO also has excellent electrical conductivity, which can facilitate the transfer of electrons and improve the efficiency of photocatalytic reactions. Moreover, metal ions, such as Fe, Cu, and Ag, can be doped into the TiO2 lattice to form metal-doped TiO2 photocatalysts. The incorporation of metal ions can modify the band gap of TiO2 and enhance its photocatalytic activity [126,127,128]. The metal ions can also act as active sites for the adsorption and degradation of organic dyes [129]. Therefore, the combination of TiO2 with other materials can enhance the its photocatalystic performance and broaden its application in the treatment of organic dye wastewater.
TiO2-based catalysts have shown promising performance in the degradation of organic dyes in various processes, including photocatalysis, sonocatalysis, and sonophotocatalysis. The efficiency of these processes depends largely on the generation of free radicals, such as •OH, •O2. Table 1 summarizes the performance of TiO2-based catalysts in the degradation of organic dyes in recent studies.
Nuengmatcha et al. showed that the ZnO/graphene/TiO2 hybrid catalyst prepared using solvothermal method was more efficient at degrading ZGT dye compared to the indiviudal components [130]. The high surface area of graphene allows for better dispersion of ZnO and TiO2, leading to increased absorption of ultrasonic irradiation and the generation of more electron-hole pairs. Yao et al. synthesized TiO2/BiOBr heterojunctions with N/Ti3+ co-doping using one-step in situ hydrothermal method and demonstrated that they exhibited higher sonocatalytic activity in degrading methylene blue compared to pristine TiO2 [131]. Specifically, NT-TB0.3 exhibited the highest degradation efficiency of 98.2% after 50 min of ultrasound irradiation. The improved catalytic activity was attributed to the formation of a heterojunction between TiO2/BiOBr, which enhances the separation of electron-hole pairs. These studies highlight the potential of hybrid and composite catalysts in enhancing the performance of sonocatalytic and photocatalytic reactions and provide insight into the mechanisms underlying their improved activity.
Sriramoju et al. synthesized RGO-TiO2−x nanocomposites using one-step in situ hydrothermal method and observed that these nanocomposites displayed exceptional photocatalytic degradation performance against diverse organic dyes when exposed to UV-visible irradiation [132]. The rate constants for Rhodamine-B, methylene blue, and rose red dye were 0.083 min−1, 0.075 min−1, and 0.093 min−1, respectively. The superior photocatalytic performance observed in TiO2−x samples was linked to the presence of highly conductive RGO, which improves the mobility of photo-generated charge carriers and reduces electron-hole pair recombination. The oxygen vacancy/Ti3+ was also identified as an important contributor to the enhanced photocatalytic activity. Mousavi and colleagues developed a Z-scheme heterojunction photocatalyst consisting of Black-TiO2 and CoTiO3, which exhibits visible-light responsiveness and is capable of decomposing a variety of organic dyes, includimg rhodamine B, methylene blue, and methyl orange [133]. The much higher photocatalytic activity of B-TiO2/CTO nanocomposites compared to B-TiO2 and CTO is attributed to the improved generation, separation, and transportation of charge carriers. Moreover, the combination of B-TiO2 with CTO increased the specific surface area of the nanocomposites, which increases the active sites on the catalyst surface and the generation of photo-generated electron-hole pairs. These studies highlight the potential of constructing hybrid nanocomposites of TiO2 to enhance the photocatalytic removal of organic dyes.
Lozano et al. synthesized a novel Ag-graphene oxide/TiO2 catalyst and showed that it effectively degraded Black 5 and orange II dyes in a sonophotocatalysis system under ultrasonic and UV irradiation [134]. Ultrasound and UV light were observed to have significant synergistic effect on the degradation of organic dyes in the presence of the catalyst. Sun et al. investigated the sonophotocatalytic removal of organic pollutants in water using N/Ti3+-doped biphasic TiO2/Bi2WO6 heterojunctions, and found that the catalytic activity of NT-TBWx in the sonophotocatalytic system for the degradation of methyl blue was much higher than that in the photocatalytic and sonocatalytic systems [135]. Compared to TiO2 and NT-TiO2, the NT-TBWx heterojunctions exhibited superior sonophotocatalytic activity. The improved sonophotocatalytic efficiency of the NT-TBWx composites is likely due to the synergistic effect of photocatalysis and sonocatalysis, as well as the N/Ti3+ co-doping and heterophase junctions.
The performance of sonocatalytic, photocatalytic, and sonophotocatalytic processes is influenced by a range of factors, including catalyst dose, solution pH, and the type and concentration of organic dyes. These factors must be optimized to improve the overall dye removal efficiency in practical applications. For instance, pH plays a critical role in determining the surface charge potential of the catalyst, which can significantly impact its interaction with the organic dye molecules. Under acidic conditions, the surface of TiO2 is positively charged, allowing the adsorption of negatively charged dye molecules, thus increases the efficiency of the photocatalytic degradation. On the other hand, the surface of TiO2 becomes negatively charged under alkaline conditions, which reduces photocatalytic activityin the degradation of negatively charged dyes. The catalyst dose is another crucial factor that affects the performance of TiO2-based catalysts. The amount of catalyst used affects the number of active sites available for the adsorption of dye molecules, which directly influences the overal degradation rate. However, high catalyst doses may cause light shielding in the solution and reduce photocatalytic performance. In addition, the type and concentration of organic dyes also play important roles in the efficiency of TiO2-based catalysts. The adsorption of dye molecules on the surface of TiO2 is influenced by the size, structure, and chemical composition of the dye molecules, which affect the overall degradation rate. High concentrations of organic dyes can lead to increased light scattering and lower photocatalytic activity.
Taken together, photocatalytic, sonocatalytic, and sonophotocatalytic activity of TiO2-based catalysts is influenced by various factors, such as pH, catalyst dose, and the type and concentration of organic dyes. In practical applications, it is necessary to optimize these factors to achieve efficient treatment of organic pollutants in water and wastewater. Addressing the challenges of high cost and limited efficacy under visible light is crucial for the widespread adoption of TiO2-based catalysts in water and wastewater treatment.

5. Further Research Trends

While the general mechanism of sonophotocatalysis has been relativly well understood, there are several challenges that need to be addressed to make it a practical and effective method for dye decolorization. One of the major challenges is the scale-up of the sonophotocatalytic process from laboratory to industrial scale, as the reaction conditions and equipment used in the laboratory may not be suitable for large-scale applications. Thus, it is necessary to develop and optimize the sonophotocatalytic process for industrial applications, which may require innovative catalyst and reactor design, as well as novel ways of supplying the ultrasound and light energy.
The cost-effectiveness of the sonophotocatalytic process is a crucial aspect that needs to be considered. Despite the tremendous photocatalytic activity of noble metal/TiO2 systems, their practical utilization is remarkably constrained due to the high cost as well as limited accessibility for precious metals. This hinders the widespread application of sonophotocatalysis for dye decolorization, necessitating the development of cost-effective catalysts with high activity and stability, such as non-noble metal-based catalysts or composites of TiO2 with other materials. To enhance the properties of TiO2-based heterojunction photocatalysts, more efficient synthesis techniques must be explored to produce catalysts with tailored morphologies and compositions. However, it is challenging to mass produce high-quality, homogeneous TiO2-based heterostructure photocatalysts. Therefore, the design as well as performance for TiO2-based heterojunctions must be further improved, which requires better understanding of the photocatalytic reaction mechanism. Additional investigation is required to explore both the thermodynamics and kinetics of surface catalytic processes, as well as the mechanism of charge carrier transfer. To enable the effective utilization of TiO2-based heterojunction photocatalysts in natural environments with sunlight, it is crucial to extend the excitation wavelength of photocatalysts, particularly by broadening their light-response from UV to visible light, which can enhance their solar conversion efficiency. Additionally, the effect of environmental factors, such as temperature, pH, and the presence of other pollutants, on the sonophotocatalytic process need to be investigated. Changes in these factors can affect the performance of sonophotocatalytic processes, necessitating the optimization of treatment conditions to aqequate decolorization efficiency. In brief, the cost-effectiveness, synthesis techniques, photocatalytic reaction mechanism, and environmental factors are critical aspects that must be considered to enhance the practical application of sonophotocatalytic processes for dye decolorization.
TiO2 is only responsive to UV light, which accounts for a small portion of the solar spectrum. Therefore, there is a need for the development of visible-light-responsive TiO2-based catalysts to expand their applications in water and wastewater treatment. Several strategies have been proposed to improve the visible-light responsiveness of TiO2-based catalysts, such as doping with transition metals, modifying with carbon materials, and forming heterojunctions with other semiconductors.
In addition to the technical challenges of sonophotocatalytic treatment of organic pollutants, more in-depth understanding of the mechanism of sonophotocatalytic degradation of dyes is essential. Investigating the interactions between the catalyst, organic dye, and environmental factors is critical for developing an effective and efficient sonophotocatalytic process. Characterization of the reaction pathways and intermediates of dyes enables the prediction of the toxicity and environmental impact of the degradation products. Therefore, future research should focus on developing detailed mechanistic models that can predict the reaction pathways and intermediates in sonophotocatalytic degradation of dyes. This requires a combination of experimental and theoretical approaches to elucidate the complex interplay between the catalyst, organic dye, and environmental factors.
Besides the above technical and mechanistic challenges for the sonophotocatalytic treatment of organic pollutants, there are potential risks associated with the process that must be assessed. The release of nanoparticles from catalyst breakdown may have adverse effects on human health and the ecosystem. This should be thoroughly explored, and appropriate measures should be taken to mitigate their impact on the ecosystem and human health. Nanoparticle release can be minimized by optimizing the sonophotocatalytic process and designing catalysts with minimal nanoparticle release. Overall, a comprehensive risk assessment of the sonophotocatalytic process is essential to ensure that it is safe and sustainable for practical applications.

6. Conclusions

The technology of sonophotocatalysis has become an important method for treating organic pollutants in water and wastewater. Sonophotocatalysis has significant synergistic effect, resulting in faster pollutant removal compared to sonocatalysis and photocatalysis. With significant improvements in terms of efficiency and treatment time, sonophotocatalysis has the potential to be a practical and effective method for dye decolorization,. However, several challenges need to be addressed, including scale-up, cost-effectiveness, optimization of process conditions, mechanistic understanding, and risk assessment, to ensure that sonophotocatalysis can be widely applied in the treatment of dye wastewater.
The synergistic effect of sonophotocatalysis offers unique opportunity to overcome some of the limitations of other treatment technologies, including sonocatalysis and photocatalysis. Therefore, further research in this field could lead to the development of new and efficient water treatment technologies that can address a wide range of environmental problems.

Author Contributions

G.W.: Conceptualization, Writing—original draft, Writing—review & editing; H.C.: Conceptualization, Resources, Supervision, Writing—original draft, Writing—review & editing. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Natural Science Foundation of China (Grant Nos.: U2006212 and 41725015).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No data are associated with this article.

Acknowledgments

This work was supported in parts by the Natural Science Foundation of China (Grant Nos. 41725015 and U2006212).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yang, W.; Ding, K.; Chen, J.; Wang, H.; Deng, X. Synergistic Multisystem Photocatalytic Degradation of Anionic and Cationic Dyes Using Graphitic Phase Carbon Nitride. Molecules 2023, 28, 2796. [Google Scholar] [CrossRef] [PubMed]
  2. Wei, X.; Feng, H.; Li, L.; Gong, J.; Jiang, K.; Xue, S.; Chu, P.K. Synthesis of tetragonal prismatic γ-In2Se3 nanostructures with predominantly {110} facets and photocatalytic degradation of tetracycline. Appl. Catal. B-Environ. 2020, 260, 118218. [Google Scholar] [CrossRef]
  3. Zou, X.; Zhang, J.; Zhao, X.; Zhang, Z. MoS2/RGO composites for photocatalytic degradation of ranitidine and elimination of NDMA formation potential under visible light. Chem. Eng. J. 2020, 383, 123084. [Google Scholar] [CrossRef]
  4. Huang, D.; Wang, H.; Wu, Y. Photocatalytic Aerobic Oxidation of Biomass-Derived 5-HMF to DFF over MIL-53(Fe)/g-C3N4 Composite. Molecules 2022, 27, 8537. [Google Scholar] [CrossRef] [PubMed]
  5. Heidari, S.; Haghighi, M.; Shabani, M. Sunlight-activated BiOCl/BiOBr–Bi24O31Br10 photocatalyst for the removal of pharmaceutical compounds. J. Clean. Prod. 2020, 259, 120679. [Google Scholar] [CrossRef]
  6. Zhou, Y.; Yu, M.; Liang, H.; Chen, J.; Xu, L.; Niu, J. Novel dual-effective Z-scheme heterojunction with g-C3N4, Ti3C2 MXene and black phosphorus for improving visible light-induced degradation of ciprofloxacin. Appl. Catal. B-Environ. 2021, 291, 120105. [Google Scholar] [CrossRef]
  7. Liu, K.; Tong, Z.; Muhammad, Y.; Huang, G.; Zhang, H.; Wang, Z.; Zhu, Y.; Tang, R. Synthesis of sodium dodecyl sulfate modified BiOBr/magnetic bentonite photocatalyst with Three-dimensional parterre like structure for the enhanced photodegradation of tetracycline and ciprofloxacin. Chem. Eng. J. 2020, 388, 124374. [Google Scholar] [CrossRef]
  8. Zhang, M.; Lai, C.; Li, B.; Huang, D.; Liu, S.; Qin, L.; Yi, H.; Fu, Y.; Xu, F.; Li, M. Ultrathin oxygen-vacancy abundant WO3 decorated monolayer Bi2WO6 nanosheet: A 2D/2D heterojunction for the degradation of Ciprofloxacin under visible and NIR light irradiation. J. Colloid Interface Sci. 2019, 556, 557–567. [Google Scholar] [CrossRef]
  9. Irshad, A.; Warsi, M.F.; Agboola, P.O.; Dastgeer, G.; Shahid, M. Sol-gel assisted Ag doped NiAl2O4 nanomaterials and their nanocomposites with g-C3N4 nanosheets for the removal of organic effluents. J. Alloys Compd. 2022, 902, 163805. [Google Scholar] [CrossRef]
  10. Gao, P.; Cui, J.; Deng, Y. Direct regeneration of ion exchange resins with sulfate radical-based advanced oxidation for enabling a cyclic adsorption–regeneration treatment approach to aqueous perfluorooctanoic acid (PFOA). Chem. Eng. J. 2021, 405, 126698. [Google Scholar] [CrossRef]
  11. Wu, J.; Wang, T.; Wang, J.; Zhang, Y.; Pan, W.-P. A novel modified method for the efficient removal of Pb and Cd from wastewater by biochar: Enhanced the ion exchange and precipitation capacity. Sci. Total Environ. 2021, 754, 142150. [Google Scholar] [CrossRef] [PubMed]
  12. Vapnik, H.; Elbert, J.; Su, X. Redox-copolymers for the recovery of rare earth elements by electrochemically regenerated ion-exchange. J. Mater. Chem. A 2021, 9, 20068–20077. [Google Scholar] [CrossRef]
  13. Chen, J.; Li, Y.; Li, M.; Shi, J.; Wang, L.; Luo, S.; Liu, H. Chemical Flocculation-Based Green Algae Materials for Photobiological Hydrogen Production. ACS Appl. Bio. Mater. 2022, 5, 897–903. [Google Scholar] [CrossRef] [PubMed]
  14. Kurniawan, S.B.; Imron, M.F.; Sługocki, Ł.; Nowakowski, K.; Ahmad, A.; Najiya, D.; Abdullah, S.R.S.; Othman, A.R.; Purwanti, I.F.; Hasan, H.A. Assessing the effect of multiple variables on the production of bioflocculant by Serratia marcescens: Flocculating activity, kinetics, toxicity, and flocculation mechanism. Sci. Total Environ. 2022, 836, 155564. [Google Scholar] [CrossRef]
  15. Wang, X.; Wang, D.; Xu, J.; Fu, J.; Zheng, G.; Zhou, L. Modified chemical mineralization-alkali neutralization technology: Mineralization behavior at high iron concentrations and its application in sulfur acid spent pickling solution. Water Res. 2022, 218, 118513. [Google Scholar] [CrossRef]
  16. Xie, L.; Du, T.; Wang, J.; Ma, Y.; Ni, Y.; Liu, Z.; Zhang, L.; Yang, C.; Wang, J. Recent advances on heterojunction-based photocatalysts for the degradation of persistent organic pollutants. Chem. Eng. J. 2021, 426, 130617. [Google Scholar] [CrossRef]
  17. Wang, L.; Bahnemann, D.W.; Bian, L.; Dong, G.; Zhao, J.; Wang, C. Two-dimensional layered zinc silicate nanosheets with excellent photocatalytic performance for organic pollutant degradation and CO2 conversion. Angew. Chem. Int. Ed. 2019, 131, 8187–8192. [Google Scholar] [CrossRef]
  18. Qutub, N.; Singh, P.; Sabir, S.; Sagadevan, S.; Oh, W.-C. Enhanced photocatalytic degradation of Acid Blue dye using CdS/TiO2 nanocomposite. Sci. Rep. 2022, 12, 5759. [Google Scholar] [CrossRef]
  19. Han, B.; Xie, A.; Yu, Q.; Huang, F.; Shen, Y.; Zhu, L. Synthesis of PbSO4 crystals by hydrogel template on postprocessing strategy for secondary pollution. Appl. Surf. Sci. 2012, 261, 623–627. [Google Scholar] [CrossRef]
  20. Bao, S.; Li, K.; Ning, P.; Peng, J.; Jin, X.; Tang, L. Highly effective removal of mercury and lead ions from wastewater by mercaptoamine-functionalised silica-coated magnetic nano-adsorbents: Behaviours and mechanisms. Appl. Surf. Sci. 2017, 393, 457–466. [Google Scholar] [CrossRef]
  21. Xin, S.; Zeng, Z.; Zhou, X.; Luo, W.; Shi, X.; Wang, Q.; Deng, H.; Du, Y. Recyclable Saccharomyces cerevisiae loaded nanofibrous mats with sandwich structure constructing via bio-electrospraying for heavy metal removal. J. Hazard. Mater. 2017, 324, 365–372. [Google Scholar] [CrossRef] [PubMed]
  22. Mamba, G.; Mishra, A. Advances in magnetically separable photocatalysts: Smart, recyclable materials for water pollution mitigation. Catalysts 2016, 6, 79. [Google Scholar] [CrossRef]
  23. Wang, G.; Cheng, H. Facile synthesis of a novel recyclable dual Z-scheme WO3/NiFe2O4/BiOBr composite with broad-spectrum response and enhanced sonocatalytic performance for levofloxacin removal in aqueous solution. Chem. Eng. J. 2023, 461, 141941. [Google Scholar] [CrossRef]
  24. Vasseghian, Y.; Dragoi, E.-N.; Almomani, F. A comprehensive review on MXenes as new nanomaterials for degradation of hazardous pollutants: Deployment as heterogeneous sonocatalysis. Chemosphere 2022, 287, 132387. [Google Scholar] [CrossRef] [PubMed]
  25. Nas, M.S. AgFe2O4/MWCNT nanoparticles as novel catalyst combined adsorption-sonocatalytic for the degradation of methylene blue under ultrasonic irradiation. J. Environ. Chem. Eng. 2021, 9, 105207. [Google Scholar] [CrossRef]
  26. Dulta, K.; Koşarsoy Ağçeli, G.; Chauhan, P.; Jasrotia, R.; Chauhan, P.; Ighalo, J.O. Multifunctional CuO nanoparticles with enhanced photocatalytic dye degradation and antibacterial activity. Sustain. Environ. Res. 2022, 32, 1–15. [Google Scholar] [CrossRef]
  27. Ramamoorthy, S.; Das, S.; Balan, R.; Lekshmi, I. TiO2-ZrO2 nanocomposite with tetragonal zirconia phase and photocatalytic degradation of Alizarin Yellow GG azo dye under natural sunlight. Mater. Today Proc. 2021, 47, 4641–4646. [Google Scholar] [CrossRef]
  28. Wani, S.I.; Ganie, A.S. Ag2O incorporated ZnO-TiO2 nanocomposite: Ionic conductivity and photocatalytic degradation of an organic dye. Inorg. Chem. Commun. 2021, 128, 108567. [Google Scholar] [CrossRef]
  29. Vellingiri, K.; Vikrant, K.; Kumar, V.; Kim, K.-H. Advances in thermocatalytic and photocatalytic techniques for the room/low temperature oxidative removal of formaldehyde in air. Chem. Eng. J. 2020, 399, 125759. [Google Scholar] [CrossRef]
  30. Wang, J.; Zhang, T.; Jiang, S.; Ma, X.; Shao, X.; Liu, Y.; Wang, D.; Li, X.; Li, B. Controllable self-assembly of BiOI/oxidized mesocarbon microbeads core-shell composites: A novel hierarchical structure facilitated photocatalytic activities. Chem. Eng. Sci. 2020, 221, 115653. [Google Scholar] [CrossRef]
  31. Lei, X.; Ouyang, C.; Huang, K. A first-principles investigation of Janus MoSSe as a catalyst for photocatalytic water-splitting. Appl. Surf. Sci. 2021, 537, 147919. [Google Scholar] [CrossRef]
  32. Majumder, S.; Chatterjee, S.; Basnet, P.; Mukherjee, J. ZnO based nanomaterials for photocatalytic degradation of aqueous pharmaceutical waste solutions–A contemporary review. Environ. Nanotechnol. Monit. Manag. 2020, 14, 100386. [Google Scholar] [CrossRef]
  33. Costarramone, N.; Kartheuser, B.; Pecheyran, C.; Pigot, T.; Lacombe, S. Efficiency and harmfulness of air-purifying photocatalytic commercial devices: From standardized chamber tests to nanoparticles release. Catal. Today 2015, 252, 35–40. [Google Scholar] [CrossRef]
  34. Cushing, S.K.; Li, J.; Meng, F.; Senty, T.R.; Suri, S.; Zhi, M.; Li, M.; Bristow, A.D.; Wu, N. Photocatalytic activity enhanced by plasmonic resonant energy transfer from metal to semiconductor. J. Am. Chem. Soc. 2012, 134, 15033–15041. [Google Scholar] [CrossRef]
  35. Sheikh, M.; Pazirofteh, M.; Dehghani, M.; Asghari, M.; Rezakazemi, M.; Valderrama, C.; Cortina, J.-L. Application of ZnO nanostructures in ceramic and polymeric membranes for water and wastewater technologies: A review. Chem. Eng. J. 2020, 391, 123475. [Google Scholar] [CrossRef]
  36. Zhou, D.; Wu, S.; Cheng, G.; Che, C.-M. A gold (iii)–TADF emitter as a sensitizer for high-color-purity and efficient deep-blue solution-processed OLEDs. J. Mater. Chem. C 2022, 10, 4590–4596. [Google Scholar] [CrossRef]
  37. Nemati, F.; Nikkhah, S.H.; Elhampour, A. An environmental friendly approach for the catalyst-free synthesis of highly substituted pyrazoles promoted by ultrasonic radiation. Chin. Chem. Lett. 2015, 26, 1397–1399. [Google Scholar] [CrossRef]
  38. Ali El-Remaily, M.A.E.A.A.; El-Dabea, T.; Alsawat, M.; Mahmoud, M.H.; Alfi, A.A.; El-Metwaly, N.; Abu-Dief, A.M. Development of new thiazole complexes as powerful catalysts for synthesis of pyrazole-4-carbonitrile derivatives under ultrasonic irradiation condition supported by DFT studies. ACS Omega 2021, 6, 21071–21086. [Google Scholar] [CrossRef]
  39. Wojcieszyńska, D.; Łagoda, K.; Guzik, U. Diclofenac Biodegradation by Microorganisms and with Immobilised Systems—A Review. Catalysts 2023, 13, 412. [Google Scholar]
  40. Zhu, Z.-H.; Liu, Y.; Song, C.; Hu, Y.; Feng, G.; Tang, B.Z. Porphyrin-Based Two-Dimensional Layered Metal–Organic Framework with Sono-/Photocatalytic Activity for Water Decontamination. ACS Nano 2021, 16, 1346–1357. [Google Scholar] [CrossRef]
  41. Guo, L.; Chen, Y.; Ren, Z.; Li, X.; Zhang, Q.; Wu, J.; Li, Y.; Liu, W.; Li, P.; Fu, Y. Morphology engineering of type-II heterojunction nanoarrays for improved sonophotocatalytic capability. Ultrason. Sonochem. 2021, 81, 105849. [Google Scholar] [CrossRef] [PubMed]
  42. Hosseini, M.; Kahkha, M.R.R.; Fakhri, A.; Tahami, S.; Lariche, M.J. Degradation of macrolide antibiotics via sono or photo coupled with Fenton methods in the presence of ZnS quantum dots decorated SnO2 nanosheets. J. Photochem. Photobiol. B 2018, 185, 24–31. [Google Scholar] [CrossRef] [PubMed]
  43. Preeyanghaa, M.; Vinesh, V.; Neppolian, B. Construction of S-scheme 1D/2D rod-like g-C3N4/V2O5 heterostructure with enhanced sonophotocatalytic degradation for Tetracycline antibiotics. Chemosphere 2022, 287, 132380. [Google Scholar] [CrossRef]
  44. Liu, J.; Ma, N.; Wu, W.; He, Q. Recent progress on photocatalytic heterostructures with full solar spectral responses. Chem. Eng. J. 2020, 393, 124719. [Google Scholar] [CrossRef]
  45. Rodríguez-González, V.; Obregón, S.; Patrón-Soberano, O.A.; Terashima, C.; Fujishima, A. An approach to the photocatalytic mechanism in the TiO2-nanomaterials microorganism interface for the control of infectious processes. Appl. Catal. B-Environ. 2020, 270, 118853. [Google Scholar] [CrossRef] [PubMed]
  46. Wen, X.-J.; Shen, C.-H.; Fei, Z.-H.; Fang, D.; Liu, Z.-T.; Dai, J.-T.; Niu, C.-G. Recent developments on AgI based heterojunction photocatalytic systems in photocatalytic application. Chem. Eng. J. 2020, 383, 123083. [Google Scholar] [CrossRef]
  47. Theerthagiri, J.; Lee, S.J.; Karuppasamy, K.; Arulmani, S.; Veeralakshmi, S.; Ashokkumar, M.; Choi, M.Y. Application of advanced materials in sonophotocatalytic processes for the remediation of environmental pollutants. J. Hazard. Mater. 2021, 412, 125245. [Google Scholar] [CrossRef]
  48. Qiu, P.; Park, B.; Choi, J.; Thokchom, B.; Pandit, A.B.; Khim, J. A review on heterogeneous sonocatalyst for treatment of organic pollutants in aqueous phase based on catalytic mechanism. Ultrason. Sonochem. 2018, 45, 29–49. [Google Scholar] [CrossRef] [PubMed]
  49. Liu, P.; Wu, Z.; Abramova, A.V.; Cravotto, G. Sonochemical processes for the degradation of antibiotics in aqueous solutions: A review. Ultrason. Sonochem. 2021, 74, 105566. [Google Scholar] [CrossRef]
  50. He, Y.; Ma, Z.; Junior, L.B. Distinctive binary g-C3N4/MoS2 heterojunctions with highly efficient ultrasonic catalytic degradation for levofloxacin and methylene blue. Ceram. Int. 2020, 46, 12364–12372. [Google Scholar] [CrossRef]
  51. Waheed, I.F.; Al-Janabi, O.Y.T.; Foot, P.J. Novel MgFe2O4-CuO/GO heterojunction magnetic nanocomposite: Synthesis, characterization, and batch photocatalytic degradation of methylene blue dye. J. Mol. Liq. 2022, 357, 119084. [Google Scholar] [CrossRef]
  52. Xu, X.; Xu, X.; Wang, T.; Xu, M.; Yang, X.; Hou, J.; Cao, D.; Wang, Q. Construction of Z-scheme CdS/Ag/TiO2 NTs photocatalysts for photocatalytic dye degradation and hydrogen evolution. Spectrochim. Acta A 2022, 276, 121215. [Google Scholar] [CrossRef]
  53. Abazari, R.; Sanati, S.; Morsali, A.; Kirillov, A.M. Instantaneous sonophotocatalytic degradation of tetracycline over NU-1000@ ZnIn2S4 core–shell nanorods as a robust and eco-friendly catalyst. Inorg. Chem. 2021, 60, 9660–9672. [Google Scholar] [CrossRef] [PubMed]
  54. Hoo, D.Y.; Low, Z.L.; Low, D.Y.S.; Tang, S.Y.; Manickam, S.; Tan, K.W.; Ban, Z.H. Ultrasonic cavitation: An effective cleaner and greener intensification technology in the extraction and surface modification of nanocellulose. Ultrason. Sonochem. 2022, 90, 106176. [Google Scholar] [CrossRef] [PubMed]
  55. Moftakhari Anasori Movahed, S.; Calgaro, L.; Marcomini, A. Trends and characteristics of employing cavitation technology for water and wastewater treatment with a focus on hydrodynamic and ultrasonic cavitation over the past two decades: A Scientometric analysis. Sci. Total Environ. 2023, 858, 159802. [Google Scholar] [CrossRef]
  56. He, L.-L.; Zhu, Y.; Qi, Q.; Li, X.-Y.; Bai, J.-Y.; Xiang, Z.; Wang, X. Synthesis of CaMoO4 microspheres with enhanced sonocatalytic performance for the removal of Acid Orange 7 in the aqueous environment. Sep. Purif. Technol. 2021, 276, 119370. [Google Scholar] [CrossRef]
  57. Wang, G.; Ma, X.; Liu, J.; Qin, L.; Li, B.; Hu, Y.; Cheng, H. Design and performance of a novel direct Z-scheme NiGa2O4/CeO2 nanocomposite with enhanced sonocatalytic activity. Sci. Total Environ. 2020, 741, 140192. [Google Scholar] [CrossRef]
  58. Gao, J.; Jiang, R.; Wang, J.; Kang, P.; Wang, B.; Li, Y.; Li, K.; Zhang, X. The investigation of sonocatalytic activity of Er3+: YAlO3/TiO2-ZnO composite in azo dyes degradation. Ultrason. Sonochem. 2011, 18, 541–548. [Google Scholar] [CrossRef]
  59. Gao, H.; Pei, K.; Hu, G.; Liu, W.; Meng, A.; Wang, H.; Shao, H.; Li, W. The influence of pressure on the acoustic cavitation in saturated CO2-expanded N, N-dimethylformamide. Ultrason. Sonochem. 2022, 83, 105934. [Google Scholar] [CrossRef]
  60. Kozmus, G.; Zevnik, J.; Hočevar, M.; Dular, M.; Petkovšek, M. Characterization of cavitation under ultrasonic horn tip—Proposition of an acoustic cavitation parameter. Ultrason. Sonochem. 2022, 89, 106159. [Google Scholar] [CrossRef]
  61. Yao, C.; Zhao, S.; Liu, L.; Liu, Z.; Chen, G. Ultrasonic emulsification: Basic characteristics, cavitation, mechanism, devices and application. Front. Chem. Sci. Eng. 2022, 16, 1560–1583. [Google Scholar] [CrossRef]
  62. Zhang, H.; Qiao, J.; Li, G.; Li, S.; Wang, G.; Wang, J.; Song, Y. Preparation of Ce4+-doped BaZrO3 by hydrothermal method and application in dual-frequent sonocatalytic degradation of norfloxacin in aqueous solution. Ultrason. Sonochem. 2018, 42, 356–367. [Google Scholar] [CrossRef] [PubMed]
  63. Wang, G.; Li, S.; Ma, X.; Qiao, J.; Li, G.; Zhang, H.; Wang, J.; Song, Y. A novel Z-scheme sonocatalyst system, Er3+:Y3Al5O12@Ni(Fe0.05Ga0.95)2O4-Au-BiVO4, and application in sonocatalytic degradation of sulfanilamide. Ultrason. Sonochem. 2018, 45, 150–166. [Google Scholar] [CrossRef] [PubMed]
  64. Huang, Y.; Wang, G.; Zhang, H.; Li, G.; Fang, D.; Wang, J.; Song, Y. Hydrothermal-precipitation preparation of CdS@(Er3+:Y3Al5O12/ZrO2) coated composite and sonocatalytic degradation of caffeine. Ultrason. Sonochem. 2017, 37, 222–234. [Google Scholar] [CrossRef] [PubMed]
  65. Wang, G.; Huang, Y.; Li, G.; Zhang, H.; Wang, Y.; Li, B.; Wang, J.; Song, Y. Preparation of a novel sonocatalyst, Au/NiGa2O4-Au-Bi2O3 nanocomposite, and application in sonocatalytic degradation of organic pollutants. Ultrason. Sonochem. 2017, 38, 335–346. [Google Scholar] [CrossRef]
  66. Abdurahman, M.H.; Abdullah, A.Z.; Shoparwe, N.F. A comprehensive review on sonocatalytic, photocatalytic, and sonophotocatalytic processes for the degradation of antibiotics in water: Synergistic mechanism and degradation pathway. Chem. Eng. J. 2021, 413, 127412. [Google Scholar] [CrossRef]
  67. Hu, Y.; Wei, J.; Shen, Y.; Chen, S.; Chen, X. Barrier-breaking effects of ultrasonic cavitation for drug delivery and biomarker release. Ultrason. Sonochem. 2023, 94, 106346. [Google Scholar] [CrossRef]
  68. Li, S.; Wang, G.; Qiao, J.; Zhou, Y.; Ma, X.; Zhang, H.; Li, G.; Wang, J.; Song, Y. Sonocatalytic degradation of norfloxacin in aqueous solution caused by a novel Z-scheme sonocatalyst, mMBIP-MWCNT-In2O3 composite. J. Mol. Liq. 2018, 254, 166–176. [Google Scholar] [CrossRef]
  69. Hassandoost, R.; Kotb, A.; Movafagh, Z.; Esmat, M.; Guegan, R.; Endo, S.; Jevasuwan, W.; Fukata, N.; Sugahara, Y.; Khataee, A.; et al. Nanoarchitecturing bimetallic manganese cobaltite spinels for sonocatalytic degradation of oxytetracycline. Chem. Eng. J. 2022, 431, 133851. [Google Scholar] [CrossRef]
  70. Afzal, M.Z.; Zu, P.; Zhang, C.-M.; Guan, J.; Song, C.; Sun, X.-F.; Wang, S.-G. Sonocatalytic degradation of ciprofloxacin using hydrogel beads of TiO2 incorporated biochar and chitosan. J. Hazard. Mater. 2022, 434, 128879. [Google Scholar] [CrossRef]
  71. Jorfi, S.; Pourfadakari, S.; Kakavandi, B. A new approach in sono-photocatalytic degradation of recalcitrant textile wastewater using MgO@ Zeolite nanostructure under UVA irradiation. Chem. Eng. J. 2018, 343, 95–107. [Google Scholar] [CrossRef]
  72. Isari, A.A.; Mehregan, M.; Mehregan, S.; Hayati, F.; Kalantary, R.R.; Kakavandi, B. Sono-photocatalytic degradation of tetracycline and pharmaceutical wastewater using WO3/CNT heterojunction nanocomposite under US and visible light irradiations: A novel hybrid system. J. Hazard. Mater. 2020, 390, 122050. [Google Scholar] [CrossRef] [PubMed]
  73. Liu, Y.-C.; Wang, J.-Q.; Wang, Y.; Chen, C.-L.; Wang, X.; Xiang, Z. Sonocatalytic degradation of ciprofloxacin by BiOBr/BiFeO3. Appl. Catal. A 2022, 643, 118776. [Google Scholar] [CrossRef]
  74. Xu, L.; Liu, N.-P.; An, H.-L.; Ju, W.-T.; Liu, B.; Wang, X.-F.; Wang, X. Preparation of Ag3PO4/CoWO4 S-scheme heterojunction and study on sonocatalytic degradation of tetracycline. Ultrason. Sonochem. 2022, 89, 106147. [Google Scholar] [CrossRef] [PubMed]
  75. Pang, Y.L.; Koe, A.Z.Y.; Chan, Y.Y.; Lim, S.; Chong, W.C. Enhanced Sonocatalytic Performance of Non-Metal Graphitic Carbon Nitride (g-C3N4)/Coconut Shell Husk Derived-Carbon Composite. Sustainability 2022, 14, 3244. [Google Scholar] [CrossRef]
  76. Sun, M.; Lin, X.; Meng, X.; Liu, W.; Ding, Z. Ultrasound-driven ferroelectric polarization of TiO2/Bi0.5Na0.5TiO3 heterojunctions for improved sonocatalytic activity. J. Alloys Compd. 2022, 892, 162065. [Google Scholar] [CrossRef]
  77. Li, S.; Zhang, M.; Ma, X.; Qiao, J.; Zhang, H.; Wang, J.; Song, Y. Preparation of ortho-symmetric double (OSD) Z-scheme SnO2\CdSe/Bi2O3 sonocatalyst by ultrasonic-assisted isoelectric point method for effective degradation of organic pollutants. J. Ind. Eng. Chem. 2019, 72, 157–169. [Google Scholar] [CrossRef]
  78. Lu, L.; Wang, T.; Fang, C.; Song, L.; Qian, C.; Lv, Z.; Fang, Y.; Liu, X.; Yu, X.; Xu, X.; et al. Oncolytic Impediment/Promotion Balance Disruption by Sonosensitizer-Free Nanoplatforms Unfreezes Autophagy-Induced Resistance to Sonocatalytic Therapy. ACS Appl. Mater. Interfaces 2022, 14, 36462–36472. [Google Scholar] [CrossRef]
  79. Haddadi, S.; Khataee, A.; Arefi-Oskoui, S.; Vahid, B.; Orooji, Y.; Yoon, Y. Titanium-based MAX-phase with sonocatalytic activity for degradation of oxytetracycline antibiotic. Ultrason. Sonochem. 2023, 92, 106255. [Google Scholar] [CrossRef]
  80. Wang, G.; Dou, K.; Cao, H.; Du, R.; Liu, J.; Tsidaeva, N.; Wang, W. Designing Z-scheme CdS/WS2 heterojunctions with enhanced photocatalytic degradation of organic dyes and photoreduction of Cr (VI): Experiments, DFT calculations and mechanism. Sep. Purif. Technol. 2022, 291, 120976. [Google Scholar] [CrossRef]
  81. Dharman, R.K.; Shejale, K.P.; Kim, S.Y. Efficient sonocatalytic degradation of heavy metal and organic pollutants using CuS/MoS2 nanocomposites. Chemosphere 2022, 305, 135415. [Google Scholar] [CrossRef] [PubMed]
  82. Zhou, Q.; Ma, S.; Zhan, S. Superior photocatalytic disinfection effect of Ag-3D ordered mesoporous CeO2 under visible light. Appl. Catal. B-Environ. 2018, 224, 27–37. [Google Scholar] [CrossRef]
  83. Wan, L.; Zhou, Q.; Wang, X.; Wood, T.E.; Wang, L.; Duchesne, P.N.; Guo, J.; Yan, X.; Xia, M.; Li, Y.F. Cu2O nanocubes with mixed oxidation-state facets for (photo) catalytic hydrogenation of carbon dioxide. Nat. Catal. 2019, 2, 889–898. [Google Scholar] [CrossRef]
  84. Xu, L.; Wu, X.-Q.; Li, C.-Y.; Liu, N.-P.; An, H.-L.; Ju, W.-T.; Lu, W.; Liu, B.; Wang, X.-F.; Wang, Y.; et al. Sonocatalytic degradation of tetracycline by BiOBr/FeWO4 nanomaterials and enhancement of sonocatalytic effect. J. Clean. Prod. 2023, 394, 136275. [Google Scholar] [CrossRef]
  85. Xiang, W.; Ji, Q.; Xu, C.; Guo, Y.; Liu, Y.; Sun, D.; Zhou, W.; Xu, Z.; Qi, C.; Yang, S. Accelerated photocatalytic degradation of iohexol over Co3O4/g-C3N4/Bi2O2CO3 of pn/nn dual heterojunction under simulated sunlight by persulfate. Appl. Catal. B Environ. 2021, 285, 119847. [Google Scholar] [CrossRef]
  86. Qiu, J.; Li, M.; Xu, J.; Zhang, X.-F.; Yao, J. Bismuth sulfide bridged hierarchical Bi2S3/BiOCl@ZnIn2S4 for efficient photocatalytic Cr (VI) reduction. J. Hazard. Mater. 2020, 389, 121858. [Google Scholar] [CrossRef]
  87. Liu, J.; Wang, G.; Li, B.; Ma, X.; Hu, Y.; Cheng, H. A high-efficiency mediator-free Z-scheme Bi2MoO6/AgI heterojunction with enhanced photocatalytic performance. Sci. Total Environ. 2021, 784, 147227. [Google Scholar] [CrossRef]
  88. Vaiano, V.; Iervolino, G.; Sannino, D.; Murcia, J.J.; Hidalgo, M.C.; Ciambelli, P.; Navío, J.A. Photocatalytic removal of patent blue V dye on Au-TiO2 and Pt-TiO2 catalysts. Appl. Catal. B-Environ. 2016, 188, 134–146. [Google Scholar] [CrossRef]
  89. Li, S.; Zhang, M.; Qu, Z.; Cui, X.; Liu, Z.; Piao, C.; Li, S.; Wang, J.; Song, Y. Fabrication of highly active Z-scheme Ag/g-C3N4-Ag-Ag3PO4 (1 1 0) photocatalyst photocatalyst for visible light photocatalytic degradation of levofloxacin with simultaneous hydrogen production. Chem. Eng. J. 2020, 382, 122394. [Google Scholar] [CrossRef]
  90. Zhang, D.; Yang, Z.; Hao, J.; Zhang, T.; Sun, Q.; Wang, Y. Boosted charge transfer in dual Z-scheme BiVO4@ ZnIn2S4/Bi2Sn2O7 heterojunctions: Towards superior photocatalytic properties for organic pollutant degradation. Chemosphere 2021, 276, 130226. [Google Scholar] [CrossRef]
  91. Wang, G.; Ma, X.; Wang, C.; Li, S.; Qiao, J.; Zhang, H.; Li, G.; Wang, J.; Song, Y. Highly efficient visible-light driven photocatalytic hydrogen evolution over Er3+: YAlO3/Ta2O5/rGO/MoSe2 nanocomposite. J. Mol. Liq. 2018, 260, 375–385. [Google Scholar] [CrossRef]
  92. Wang, G.; Ma, X.; Wei, S.; Li, S.; Qiao, J.; Wang, J.; Song, Y. Highly efficient visible-light driven photocatalytic hydrogen production from a novel Z-scheme Er3+: YAlO3/Ta2O5-V5+|| Fe3+-TiO2/Au coated composite. J. Power Sources 2018, 373, 161–171. [Google Scholar] [CrossRef]
  93. Zhao, G.; Ding, J.; Zhou, F.; Chen, X.; Wei, L.; Gao, Q.; Wang, K.; Zhao, Q. Construction of a visible-light-driven magnetic dual Z-scheme BiVO4/g-C3N4/NiFe2O4 photocatalyst for effective removal of ofloxacin: Mechanisms and degradation pathway. Chem. Eng. J. 2021, 405, 126704. [Google Scholar] [CrossRef]
  94. Molla, A.; Kim, A.Y.; Woo, J.C.; Cho, H.S.; Youk, J.H. Study on preparation methodology of zero-valent iron decorated on graphene oxide for highly efficient sonocatalytic dye degradation. J. Environ. Chem. Eng. 2022, 10, 107214. [Google Scholar] [CrossRef]
  95. Wang, X.; He, X.-S.; Li, C.-Y.; Liu, S.-L.; Lu, W.; Xiang, Z.; Wang, Y. Sonocatalytic removal of tetracycline in the presence of S-scheme Cu2O/BiFeO3 heterojunction: Operating parameters, mechanisms, degradation pathways and toxicological evaluation. J. Water Process Eng. 2023, 51, 103345. [Google Scholar] [CrossRef]
  96. Dharman, R.K.; Palanisamy, G.; Oh, T.H. Sonocatalytic degradation of ciprofloxacin and organic pollutant by 1T/2H phase MoS2 in Polyvinylidene fluoride nanocomposite membrane. Chemosphere 2022, 308, 136571. [Google Scholar] [CrossRef]
  97. Akdağ, S.; Sadeghi Rad, T.; Keyikoğlu, R.; Orooji, Y.; Yoon, Y.; Khataee, A. Peroxydisulfate-assisted sonocatalytic degradation of metribuzin by La-doped ZnFe layered double hydroxide. Ultrason. Sonochem. 2022, 91, 106236. [Google Scholar] [CrossRef]
  98. Liu, C.; Mao, S.; Wang, H.; Wu, Y.; Wang, F.; Xia, M.; Chen, Q. Peroxymonosulfate-assisted for facilitating photocatalytic degradation performance of 2D/2D WO3/BiOBr S-scheme heterojunction. Chem. Eng. J. 2022, 430, 132806. [Google Scholar] [CrossRef]
  99. de Jesús Ruíz-Baltazar, Á. Sonochemical activation-assisted biosynthesis of Au/Fe3O4 nanoparticles and sonocatalytic degradation of methyl orange. Ultrason. Sonochem. 2021, 73, 105521. [Google Scholar] [CrossRef]
  100. Zhang, J.; Zhao, Y.; Zhang, K.; Zada, A.; Qi, K. Sonocatalytic degradation of tetracycline hydrochloride with CoFe2O4/g-C3N4 composite. Ultrason. Sonochem. 2023, 94, 106325. [Google Scholar] [CrossRef]
  101. Wang, X.; Yu, S.; Li, Z.-H.; He, L.-L.; Liu, Q.-L.; Hu, M.-Y.; Xu, L.; Wang, X.-F.; Xiang, Z. Fabrication Z-scheme heterojunction of Ag2O/ZnWO4 with enhanced sonocatalytic performances for meloxicam decomposition: Increasing adsorption and generation of reactive species. Chem. Eng. J. 2021, 405, 126922. [Google Scholar] [CrossRef]
  102. Sadeghi Rad, T.; Ansarian, Z.; Khataee, A.; Vahid, B.; Doustkhah, E. N-doped graphitic carbon as a nanoporous MOF-derived nanoarchitecture for the efficient sonocatalytic degradation process. Sep. Purif. Technol. 2021, 256, 117811. [Google Scholar] [CrossRef]
  103. Gote, Y.M.; Sinhmar, P.S.; Gogate, P.R. Sonocatalytic Degradation of Chrysoidine R Dye Using Ultrasonically Synthesized NiFe2O4 Catalyst. Catalysts 2023, 13, 597. [Google Scholar] [CrossRef]
  104. Joseph, C.G.; Puma, G.L.; Bono, A.; Krishnaiah, D. Sonophotocatalysis in advanced oxidation process: A short review. Ultrason. Sonochem. 2009, 16, 583–589. [Google Scholar] [CrossRef] [PubMed]
  105. Malika, M.; Sonawane, S.S. The sono-photocatalytic performance of a Fe2O3 coated TiO2 based hybrid nanofluid under visible light via RSM. Colloids Surf. A 2022, 641, 128545. [Google Scholar] [CrossRef]
  106. Mosleh, S.; Rahimi, M.R.; Ghaedi, M.; Asfaram, A.; Jannesar, R.; Sadeghfar, F. A rapid and efficient sonophotocatalytic process for degradation of pollutants: Statistical modeling and kinetics study. J. Mol. Liq. 2018, 261, 291–302. [Google Scholar] [CrossRef]
  107. Karim, A.V.; Shriwastav, A. Degradation of amoxicillin with sono, photo, and sonophotocatalytic oxidation under low-frequency ultrasound and visible light. Environ. Res. 2021, 200, 111515. [Google Scholar] [CrossRef]
  108. Dinesh, G.K.; Anandan, S.; Sivasankar, T. Sonophotocatalytic treatment of Bismarck Brown G dye and real textile effluent using synthesized novel Fe (0)-doped TiO2 catalyst. RSC Adv. 2015, 5, 10440–10451. [Google Scholar] [CrossRef]
  109. Al-Musawi, T.J.; Rajiv, P.; Mengelizadeh, N.; Mohammed, I.A.; Balarak, D. Development of sonophotocatalytic process for degradation of acid orange 7 dye by using titanium dioxide nanoparticles/graphene oxide nanocomposite as a catalyst. J. Environ. Manag. 2021, 292, 112777. [Google Scholar] [CrossRef]
  110. Rameshbabu, R.; Kumar, N.; Pecchi, G.; Delgado, E.J.; Karthikeyan, C.; Mangalaraja, R. Ultrasound-assisted synthesis of rGO supported NiO-TiO2 nanocomposite: An efficient superior sonophotocatalyst under diffused sunlight. J. Environ. Chem. Eng. 2022, 10, 107701. [Google Scholar] [CrossRef]
  111. Gokul, P.; Vinoth, R.; Neppolian, B.; Anandhakumar, S. Binary metal oxide nanoparticle incorporated composite multilayer thin films for sono-photocatalytic degradation of organic pollutants. Appl. Surf. Sci. 2017, 418, 119–127. [Google Scholar] [CrossRef]
  112. Ding, Z.; Sun, M.; Liu, W.; Sun, W.; Meng, X.; Zheng, Y. Ultrasonically synthesized N-TiO2/Ti3C2 composites: Enhancing sonophotocatalytic activity for pollutant degradation and nitrogen fixation. Sep. Purif. Technol. 2021, 276, 119287. [Google Scholar] [CrossRef]
  113. Wang, S.; Gong, Q.; Liang, J. Sonophotocatalytic degradation of methyl orange by carbon nanotube/TiO2 in aqueous solutions. Ultrason. Sonochem. 2009, 16, 205–208. [Google Scholar] [CrossRef]
  114. Ahmad, R.; Ahmad, Z.; Khan, A.U.; Mastoi, N.R.; Aslam, M.; Kim, J. Photocatalytic systems as an advanced environmental remediation: Recent developments, limitations and new avenues for applications. J. Environ. Chem. Eng. 2016, 4, 4143–4164. [Google Scholar] [CrossRef]
  115. Sathishkumar, P.; Mangalaraja, R.V.; Mansilla, H.D.; Gracia-Pinilla, M.; Anandan, S. Sonophotocatalytic (42 kHz) degradation of Simazine in the presence of Au–TiO2 nanocatalysts. Appl. Catal. B-Environ. 2014, 160, 692–700. [Google Scholar] [CrossRef]
  116. Hapeshi, E.; Fotiou, I.; Fatta-Kassinos, D. Sonophotocatalytic treatment of ofloxacin in secondary treated effluent and elucidation of its transformation products. Chem. Eng. J. 2013, 224, 96–105. [Google Scholar] [CrossRef]
  117. Mosleh, S.; Rahimi, M.; Ghaedi, M.; Dashtian, K. Sonophotocatalytic degradation of trypan blue and vesuvine dyes in the presence of blue light active photocatalyst of Ag3PO4/Bi2S3-HKUST-1-MOF: Central composite optimization and synergistic effect study. Ultrason. Sonochem. 2016, 32, 387–397. [Google Scholar] [CrossRef]
  118. Babu, S.G.; Karthik, P.; John, M.C.; Lakhera, S.K.; Ashokkumar, M.; Khim, J.; Neppolian, B. Synergistic effect of sono-photocatalytic process for the degradation of organic pollutants using CuO-TiO2/rGO. Ultrason. Sonochem. 2019, 50, 218–223. [Google Scholar] [CrossRef]
  119. Benomara, A.; Guenfoud, F.; Mokhtari, M.; Boudjemaa, A. Sonolytic, sonocatalytic and sonophotocatalytic degradation of a methyl violet 2B using iron-based catalyst. React. Kinet. Mech. Catal. 2021, 132, 513–528. [Google Scholar] [CrossRef]
  120. Ahmad, M.; Ahmed, E.; Hong, Z.; Ahmed, W.; Elhissi, A.; Khalid, N. Photocatalytic, sonocatalytic and sonophotocatalytic degradation of Rhodamine B using ZnO/CNTs composites photocatalysts. Ultrason. Sonochem. 2014, 21, 761–773. [Google Scholar] [CrossRef]
  121. Abdullah, A.Z.; Ling, P.Y. Heat treatment effects on the characteristics and sonocatalytic performance of TiO2 in the degradation of organic dyes in aqueous solution. J. Hazard. Mater. 2010, 173, 159–167. [Google Scholar] [CrossRef] [PubMed]
  122. Mahanta, U.; Khandelwal, M.; Deshpande, A.S. TiO2@SiO2 nanoparticles for methylene blue removal and photocatalytic degradation under natural sunlight and low-power UV light. Appl. Surf. Sci. 2022, 576, 151745. [Google Scholar] [CrossRef]
  123. Rajagopal, S.; Paramasivam, B.; Muniyasamy, K. Photocatalytic removal of cationic and anionic dyes in the textile wastewater by H2O2 assisted TiO2 and micro-cellulose composites. Sep. Purif. Technol. 2020, 252, 117444. [Google Scholar] [CrossRef]
  124. Li, S.; Wang, J.; Xia, Y.; Li, P.; Wu, Y.; Yang, K.; Song, Y.; Jiang, S.; Zhang, T.; Li, B. Boosted electron-transfer by coupling Ag and Z-scheme heterostructures in CdSe-Ag-WO3-Ag for excellent photocatalytic H2 evolution with simultaneous degradation. Chem. Eng. J. 2021, 417, 129298. [Google Scholar] [CrossRef]
  125. Hunge, Y.M.; Yadav, A.A.; Dhodamani, A.G.; Suzuki, N.; Terashima, C.; Fujishima, A.; Mathe, V.L. Enhanced photocatalytic performance of ultrasound treated GO/TiO2 composite for photocatalytic degradation of salicylic acid under sunlight illumination. Ultrason. Sonochem. 2020, 61, 104849. [Google Scholar] [CrossRef]
  126. Ribao, P.; Corredor, J.; Rivero, M.J.; Ortiz, I. Role of reactive oxygen species on the activity of noble metal-doped TiO2 photocatalysts. J. Hazard. Mater. 2019, 372, 45–51. [Google Scholar] [CrossRef]
  127. Gogoi, D.; Namdeo, A.; Golder, A.K.; Peela, N.R. Ag-doped TiO2 photocatalysts with effective charge transfer for highly efficient hydrogen production through water splitting. Int. J. Hydrogen Energy 2020, 45, 2729–2744. [Google Scholar] [CrossRef]
  128. 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]
  129. Zhang, H.; Tang, P.; Yang, K.; Wang, Q.; Feng, W.; Tang, Y. PAA/TiO2@C composite hydrogels with hierarchical pore structures as high efficiency adsorbents for heavy metal ions and organic dyes removal. Desalination 2023, 558, 116620. [Google Scholar] [CrossRef]
  130. Nuengmatcha, P.; Chanthai, S.; Mahachai, R.; Oh, W.-C. Sonocatalytic performance of ZnO/graphene/TiO2 nanocomposite for degradation of dye pollutants (methylene blue, texbrite BAC-L, texbrite BBU-L and texbrite NFW-L) under ultrasonic irradiation. Dyes Pigm. 2016, 134, 487–497. [Google Scholar] [CrossRef]
  131. Yao, Y.; Sun, M.; Yuan, X.; Zhu, Y.; Lin, X.; Anandan, S. One-step hydrothermal synthesis of N/Ti3+ co-doping multiphasic TiO2/BiOBr heterojunctions towards enhanced sonocatalytic performance. Ultrason. Sonochem. 2018, 49, 69–78. [Google Scholar] [CrossRef]
  132. Sriramoju, J.B.; Muniyappa, M.; Marilingaiah, N.R.; Sabbanahalli, C.; Shetty, M.; Mudike, R.; Chitrabanu, C.; Shivaramu, P.D.; Nagaraju, G.; Rangappa, K.S. Carbon-based TiO2-x heterostructure nanocomposites for enhanced photocatalytic degradation of dye molecules. Ceram. Int. 2021, 47, 10314–10321. [Google Scholar] [CrossRef]
  133. Mousavi, M.; Ghasemi, J.B. Novel visible-light-responsive Black-TiO2/CoTiO3 Z-scheme heterojunction photocatalyst with efficient photocatalytic performance for the degradation of different organic dyes and tetracycline. J. Taiwan Inst. Chem. Eng. 2021, 121, 168–183. [Google Scholar] [CrossRef]
  134. May-Lozano, M.; Lopez-Medina, R.; Escamilla, V.M.; Rivadeneyra-Romero, G.; Alonzo-Garcia, A.; Morales-Mora, M.; González-Díaz, M.; Martinez-Degadillo, S. Intensification of the Orange II and Black 5 degradation by sonophotocatalysis using Ag-graphene oxide/TiO2 systems. Chem. Eng. Process. 2020, 158, 108175. [Google Scholar] [CrossRef]
  135. Sun, M.; Yao, Y.; Ding, W.; Anandan, S. N/Ti3+ co-doping biphasic TiO2/Bi2WO6 heterojunctions: Hydrothermal fabrication and sonophotocatalytic degradation of organic pollutants. J. Alloys Compd. 2020, 820, 153172. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of the acoustic generation of a cavitation bubble in water (after [66]).
Figure 1. Schematic illustration of the acoustic generation of a cavitation bubble in water (after [66]).
Molecules 28 03706 g001
Figure 2. Schematic diagram of sonocatalytic mechanism (after [23]).
Figure 2. Schematic diagram of sonocatalytic mechanism (after [23]).
Molecules 28 03706 g002
Figure 3. Schematic illustration of photocatalytic mechanism (after [23]).
Figure 3. Schematic illustration of photocatalytic mechanism (after [23]).
Molecules 28 03706 g003
Figure 4. Schematic diagram of sonophotocatalytic mechanism (after [23]).
Figure 4. Schematic diagram of sonophotocatalytic mechanism (after [23]).
Molecules 28 03706 g004
Table 1. Summary of performance of TiO2 based catalysts in the degradation of organic dyes.
Table 1. Summary of performance of TiO2 based catalysts in the degradation of organic dyes.
TiO2-Based
Catalyst
DyeCatalytic ConditionsExperiment ConditionsResult (Kinetic Constant (k) or
Degradation Efficiency (%))
Ref.
ZnO/graphene/TiO2 (ZGT)Methylene blueBath sonicator
Power = 750 W
Frequency = 20 kHz
[Catalyst] = 1.00 g/L
[Pollutant] = 20 mg/L
1.97 × 10−2 min−1[130]
N/Ti3+ TiO2/BiOBr0.3 Methylene blue, rhodamine BBath sonicator
Power = 180 W
Frequency = 30 kHz
[Catalyst] = 7.5 mg
[Pollutant] = 5 mg/L
Time = 50 min
98.2%[131]
Er3+: YAlO3/TiO2-ZnOAcid red BBath sonicator
Power = 50 W
Frequency = 40 kHz
[Catalyst] = 1.0 g/L
[Pollutant] = 10 mg/L
Time = 60 min
76.84%[58]
RGO-TiO2−xMethylene blueLight power = 150 W[Catalyst] = 20 mg
[Pollutant] = 5 ppm
0.075 min−1[132]
Black-TiO2/CoTiO3Rhodamine B, methylene blue, and methyl orange Light power = 50 W[Catalyst] = 100 mg [Pollutant] = 5 ppm
Time = 60 min
99%[133]
Au-TiO2Patent blue VLight power = 570 W/m2[Catalyst] = 23 g/L
[Pollutant] = 7 mg/L
Time = 180 min
93%[88]
TiO2_Ag_GrapheneBlack 5Bath sonicator
Power = 30 W/L
Frequency = 40 kHz
UV light power = 5 W
[Catalyst] = 0.03 g
[Pollutant] = 5 mg/L
0.05 min−1[134]
NT-TBWxMethylene blue (MB)Bath sonicator
Power = 180 W
Frequency = 35 kHz
UV light power = 100 mW/cm2
[Catalyst] = 7.5 mg
[Pollutant] = 5 mg/L
Time = 50 min
99%[135]
CNTs/TiO2methyl orange (MO)Bath sonicator
Power = 50 W
Frequency = 20kHz
UV light Power = 30 W
[Catalyst] = 50 mg
[Pollutant] = 25 ppm
0.01118 min−1[113]
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

Wang, G.; Cheng, H. Application of Photocatalysis and Sonocatalysis for Treatment of Organic Dye Wastewater and the Synergistic Effect of Ultrasound and Light. Molecules 2023, 28, 3706. https://doi.org/10.3390/molecules28093706

AMA Style

Wang G, Cheng H. Application of Photocatalysis and Sonocatalysis for Treatment of Organic Dye Wastewater and the Synergistic Effect of Ultrasound and Light. Molecules. 2023; 28(9):3706. https://doi.org/10.3390/molecules28093706

Chicago/Turabian Style

Wang, Guowei, and Hefa Cheng. 2023. "Application of Photocatalysis and Sonocatalysis for Treatment of Organic Dye Wastewater and the Synergistic Effect of Ultrasound and Light" Molecules 28, no. 9: 3706. https://doi.org/10.3390/molecules28093706

Article Metrics

Back to TopTop