Next Article in Journal
Catalytic Dehydrogenative Cracking of C4 Hydrocarbons on a Bifunctional Metal–Acid Catalyst
Previous Article in Journal / Special Issue
Sustainable Biofuel Production from Sludge by Oleaginous Fungi: Effect of Process Variables on Lipid Accumulation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Ultrasonic-Assisted Fabrication of TiO2-Based Composite Photocatalysts for Enhanced Photocatalysis of Organic Pollutants: A Review

1
Nanotechnology & Catalysis Research Center (NANOCAT), Institute for Advanced Studies, Universiti Malaya, Kuala Lumpur 50603, Malaysia
2
Institute of Art and Design, Huaihua University, Huaihua 418000, China
3
Department of Allied Sciences, School of Health Sciences and Technology, UPES, Dehradun 248007, India
4
Institute of Environmental Remediation and Human Health, School of Ecology and Environment, Southwest Forestry University, Kunming 650224, China
5
Mechanical Engineering Department, College of Engineering, King Khalid University, Abha 61421, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(11), 1010; https://doi.org/10.3390/catal15111010
Submission received: 12 September 2025 / Revised: 19 October 2025 / Accepted: 21 October 2025 / Published: 27 October 2025
(This article belongs to the Special Issue Catalysis Accelerating Energy and Environmental Sustainability)

Abstract

Water contamination and the global energy crisis are two of the most significant challenges in the world. Titanium dioxide (TiO2) has garnered attention due to its promising photocatalytic performance. However, its wide band gap energy limits its efficiency under visible light irradiation. To address this, TiO2-based composite photocatalysts have been developed to narrow the band gap energy and suppress the recombination of electron–hole pairs, thereby enhancing photocatalytic performance. The ultrasonic technique, through acoustic cavitation, facilitates a synthetic process involving localized transient high-pressure and high-temperature conditions to produce photocatalysts with superior photocatalytic capabilities. This review focuses on ultrasonication and ultrasonic-assisted fabrication method for modifying TiO2 into visible light-driven composite heterostructures. It discusses the parameters of ultrasonication that influence the synthesis and modification of these composites, along with the factors affecting photocatalytic performance.

1. Introduction

Water is a crucial resource for all living organisms, yet its scarcity and pollution are escalating global crises. According to the UN World Water Development Report 2024 [1], the global urban population facing water scarcity could double from 930 million in 2016 to 2.4 billion people by 2050, with 82% of wastewater released untreated into the environment [1]. Organic traces such as dyes, pesticides, herbicides, pharmaceuticals, and personal care products can be easily released into water through industrial and agricultural activities, which pose significant threats to aquatic ecosystems, the food chain, and human health due to their recalcitrant and toxic nature. Effective treatment of this contaminated water is therefore essential before its environmental release or reuse.
The textile industry alone consumes over 700,000 tons of dyes annually, with 10–15% lost in effluents [2,3]. These complex, water-soluble compounds are resistant to biological degradation [4,5,6]. Their presence in water blocks sunlight, impairing photosynthesis and biodiversity, while long-term accumulation causes serious health hazards, including damage to the kidney, liver, and central nervous system [4,6,7,8].
Similarly, pharmaceutical pollution is a pervasive threat. A global study found caffeine, metformin, and carbamazepine to be the most common pollutants in rivers, with the highest concentrations in South Asia, South America, and sub-Saharan Africa [9,10]. Introduced via industrial emissions and untreated waste, these chemically stable compounds are non-biodegradable and bio-accumulative, contributing to antibiotic resistance and chronic diseases [9,11,12].
Pesticide usage has also surged to boost crop yields, with India’s production, for example, skyrocketing from 5000 metric tons in 1958 to 102,240 tons in 1998 [13]. These persistent compounds can travel long distances through air and water, causing toxicity to both vertebrates and invertebrates, damaging organs and the central nervous system [13,14,15].
Conventional wastewater treatments like biodegradation [16], adsorption [17], membrane filtration [18], anaerobic fermentation [19], coagulation [20], flocculation [21], and electrolysis [22] suffer from drawbacks including high costs due to extensive energy consumption process, the formation of toxic by-products, the fouling of membranes, regeneration issues of the adsorbent, the production of sludge, the construction of large waste particles, incomplete removal, and a short half-life [4,7]. In contrast, advanced oxidation processes, particularly visible light-driven photocatalysis, offer an efficient, eco-friendly alternative by generating reactive oxygen species (ROS) that mineralize pollutants into harmless residues [6,23].
Titanium dioxide (TiO2) is a leading photocatalyst prized for its strong oxidation capability, non-toxicity, superior chemical stability and photostability, non-selective degradation, thorough mineralization, low energy consumption, and low cost [6,23,24,25,26]. Upon UV light excitation, it generates electron–hole pairs that form ROS like hydroxyl radicals (OH) and superoxide anion (O2), which mineralize organic compounds [23]. However, its wide band gap (3.00–3.30 eV for rutile and anatase form) restricts activation to UV light (~5% of sunlight), and rapid charge carrier recombination further limits its efficiency [6,12,25,26].
To harness visible light, researchers have developed TiO2-based composites for the photocatalytic degradation of organic pollutants. Examples include TiO2/g-C3N4 decorated with rGO sheet [27], WO3/TiO2 [28], TiO2/SiO2 [29], Cu2O/WO3/TiO2 [30], Ag3PO4/TiO2-carbon nanofiber [31], and porphyrin polymer/TiO2 composites [32]. These heterostructures narrow the effective band gap and enhance charge separation, significantly boosting photocatalytic performance [7,33].
While conventional methods like hydrothermal [34], sol–gel [35], wet impregnation [36], and chemical precipitation [37] are common for photocatalystd formation, ultrasonication has emerged as a powerful technique. It reduces reaction time, controls morphology, and produces superior photocatalysts when combined with other methods [12,24,38,39,40]. This review focuses on the ultrasonication-assisted fabrication of visible light-driven TiO2-based heterostructures for organic pollutant degradation. It examines the governing parameters of the synthesis process, the factors influencing photocatalytic efficiency, and the underlying mechanisms, followed by a discussion of the photoelectrochemical properties of these composites.

2. Basic Photocatalysis Principles

The photocatalytic process is based on the generation of redox-active species to initiate the degradation of organic contaminants [8]. Photocatalysis has been widely demonstrated to effectively eliminate various organic pollutants from wastewater sources, including pesticides [41], pharmaceuticals [42], heavy metals [43], and microplastics [44]. The fundamental mechanism of photocatalysis using titanium dioxide (TiO2) is illustrated in Figure 1. When TiO2 absorbs light energy equal to or greater than its band gap energy, electron (e) are photoexcited from the valence band (VB) to empty conduction band (CB), leaving positively charged hole (h+) in the VB. The photoexcited e then react with oxygen molecules (O2) adsorbed on the photocatalyst surface or dissolved in water, generating superoxide anion (O2), while the hole (h+) react with water molecules (H2O) and hydroxide anion (OH) to produce hydroxyl radicals (OH) and protons (H+) [3,8], as depicted in Equations (1)–(4) below:
Photocatalyst + light (hv) → e + h+
O2 adsorbed + eO2
H2O adsorbed + h+OH + H+
OH + h+OH
These potent oxidizing agents, H+, OH, and O2, react with organic pollutants to produce carbon dioxide (CO2), H2O, and mineral acid (s) [3], as demonstrated in Equations (5)–(7) below:
O2 + organic pollutants → intermediates → CO2 + H2O + mineral acid (s)
OH + organic pollutants → intermediates → CO2 + H2O + mineral acid (s)
H+ + organic pollutants → intermediates → CO2 + H2O + mineral acid (s)

TiO2

TiO2, also known as titania [45], is a prominent functional photocatalyst widely utilized as a white pigment in various products such as food coloring, paint, toothpaste, and cosmetics [46]. Researchers have extensively explored TiO2 since the initial demonstration of photocatalytic water splitting using TiO2 electrodes under ultraviolet (UV) light in 1972 [46]. TiO2 exists in three crystalline forms: anatase, rutile, and brookite [47]. While anatase and rutile have been extensively studied in photocatalysis, brookite has received comparatively less attention. Anatase TiO2 generally exhibits superior photocatalytic activity compared to rutile, which is often attributed to differences in their predominant crystal facets. The key facets for anatase are (101) and (001), whereas for rutile, they are (110), (100), and (101) [47].
Anatase TiO2 excels as a photocatalyst due to the abundance of under-bonded Ti atoms and the large Ti-O-Ti bond angles on its (001) facet [47]. In its crystal structure, Ti4+ ions are surrounded by six O2− ions, forming an octahedral geometry. Anatase TiO2 has a distorted orthorhombic structure in which each octahedron is connected to ten others, whereas rutile TiO2 has a slightly orthorhombic structure connected to eight neighboring octahedra [47]. These structural differences influence mass density, electronic band structure, and ultimately photocatalytic performance, with anatase TiO2 exhibiting enhanced photoactivity [47]. Furthermore, the typically smaller particle size and larger surface area of the anatase phase TiO2 contribute to its enhanced optical activity in photocatalytic applications compared to the rutile phase [12,48].
TiO2 functions as an n-type semiconductor photocatalyst [23]. Despite its advantages, TiO2 faces several drawbacks such as large band gap energy, poor solar energy conversion efficiency, high charge carrier recombination rates, and inefficient separation and migration rates of photoexcited electrons [8]. To address these challenges, TiO2 heterostructure composites have been developed to enhance the separation of photoexcited electron–hole pairs. These composites facilitate the movement of photoexcited electrons from the conduction band (CB) with a more negative potential to one with a less negative potential, while simultaneously transferring holes from the valence band (VB) with a more positive potential to one with a less positive potential [47]. The lower band gap energies of TiO2 composites also enhance the absorption of visible light, further boosting their photocatalytic efficiency.

3. TiO2-Based Nanocomposite Photocatalysts

3.1. Formation of Heterojunction

The construction of heterostructures by combining two semiconductors has proven to be an effective strategy for enhancing the photocatalytic performance of titanium dioxide (TiO2). This approach involves coupling two or more semiconductor materials to facilitate charge carrier separation, reduce recombination rates, and improve redox capabilities [49]. The interface formed between two distinct metal oxides or semiconductors is known as a heterojunction [50]. Heterojunctions are central to this strategy and are typically built using p-type and n-type semiconductors. These configurations can be homojunctions (formed between the same type, e.g., p-p or n-n) or heterojunctions (formed between different types, e.g., p-n), all aimed at significantly improving photocatalytic efficiency.
A p-n heterojunction enhances photocatalytic performance by leveraging the internal electric field at the semiconductor interface to promote the transfer of photoexcited charge carriers, thereby effectively separating electron–hole pairs. For instance, Xiao et al. [51] synthesized a BiOBr/TiO2 composite with a p-n heterojunction for the photocatalytic degradation of formaldehyde and dyes. High-resolution transmission electron microscopy (HRTEM) analysis revealed lattice fringe spacings corresponding to the (110) plane of BiOBr and the (101) plane of TiO2, confirming the tight attachment and successful formation of a heterojunction between the two components [51]. Such well-defined lattice fringes indicate a highly crystalline structure, which facilitates rapid charge carrier transfer across the heterointerface and reduces recombination rates [52].
The fermi level (EF) represents the highest occupied electronic energy level at absolute zero temperature and the electrochemical potential of electrons in a material [49]. In an n-type semiconductor, EF is situated close to the conduction band (CB) minimum, while in a p-type semiconductor, it lies near the valence band (VB) maximum [49]. When two distinct semiconductors come into contact, the difference in their EF causes electrons to diffuse from the material with higher EF to the one with the lower EF until thermal equilibrium is reached. This charge redistribution creates a built-in electric field at the interface [49]. Under light irradiation, this internal electric field effectively separates the newly generated photoexcited electron–hole pairs, thereby enhancing the photocatalytic performance of the heterostructure composite [50].
The Mott-Schottky (MS) plot is utilized to determine flat band potential ( V F B ) and effective charge carrier density of semiconductors under dark conditions [28]. For instance, Chau et al. [53] investigated heterojunction formation in a Cu2O/WO3/TiO2 composite photocatalyst. Figure 2a illustrates the MS plot for this composite. In such plots, n-type semiconductors like TiO2 and WO3 exhibit a positive slope, whereas p-type semiconductors like Cu2O show a negative slope [54]. The V F B is determined from the intersection point of the tangent to the MS plot with the x-axis (applied potential). A more negative reading indicates stronger reducing power of the photocatalysts [52]. It is important to note that V F B is affected by experimental conditions such as frequency and pH [55]. Chau et al. [53] confirmed the formation of a p-n heterojunction in their Cu2O/WO3/TiO2 composite, as evidenced by the characteristic inverted “V” shape in the MS plot (Figure 2a) [54,56].
Higher charge carrier density generally indicates more efficient charge separation [57]. For example, Tai et al. [56] calculated the acceptor ( N A   ) and donor densities ( N D ) for a GO/TiO2 composite photocatalyst prepared under varying photoreduction times. They noted that a 12 h photoreduction process yielded the highest charge carrier densities for both N A   a n d   N D . The formation of C-O-Ti bonds was identified as a key factor promoting charge carrier transport and increasing charge carrier density. Additionally, the researchers explored the relationship between the built-in potential ( e V b i ) of the p-n heterojunction in the GO/TiO2 composite and the charge carrier density, as depicted in Figure 2b. A larger e V b i helps suppress the recombination of photoexcited charge carriers by strengthening the internal electric field that drives the separation of electrons in the CB and holes in the VB [56]. It is worth noting that heterojunctions in photocatalysis are generally classified into several types, such as Type-II, Z-scheme, and S-scheme.

3.2. Type II Heterojunction

Figure 3a illustrates the formation of a type II heterojunction between two semiconductors, labeled A and B. In such a heterojunction, a space charge region forms at the interface upon contact. The alignment of Fermi levels (EF) causes band bending in the conduction band (CB) and valence band (VB) of both semiconductors, establishing a built-in electric field (IEF). Under light illumination, electrons in the VB of both semiconductors are excited to their respective CBs, generating electron–hole pairs. The IEF then facilitates the transfer of photoexcited electrons from the CB of semiconductor B (with a more negative potential) to the CB of semiconductor A. Simultaneously, holes transfer from the VB of semiconductor A (with a more positive potential) to the VB of semiconductor B [58,59]. This spatial separation leads to the accumulation of electrons in semiconductor A and holes in semiconductor B, thereby effectively reducing the recombination rate of charge carriers.
However, conventional Type II heterojunctions have a significant drawback: the useful electrons and holes accumulate in energy bands with weaker redox potentials. This results in a decrease in the overall redox capability of the system. Furthermore, electrostatic repulsion between electrons can hinder their transfer between semiconductor CBs [59]. To overcome these limitations, researchers have developed alternative charge transfer pathways, such as the Z-scheme heterojunction.

3.3. Z-Scheme Heterojunction

Z-scheme photocatalytic system are generally categorized into three types: traditional Z-scheme, all-solid-state Z-scheme, and direct Z-scheme configurations [59]. The foundational principle was introduced by Bard [60] in 1979, who proposed using a redox electron mediator to shuttle charges between two semiconductors for separate reduction and oxidation reactions. However, recent studies have identified challenges in the charge transfer mechanisms of both traditional and all-solid-state Z-schemes, such as inefficiency and light shielding by the mediator [59,61]. Consequently, this review will focus on the direct Z-scheme heterojunction.
In a direct Z-scheme heterojunction, electrons migrate from a photocatalyst with a higher EF to one with a lower EF to achieve energy level equilibrium, thereby equalizing the EF between the two photocatalysts [62]. Figure 3b illustrates a direct Z-scheme heterojunction where photoexcited electrons transfer from the conduction band (CB) of photocatalyst A to the valence band (VB) of photocatalyst B due to electrostatic attraction between the photoexcited charge carriers. This heterojunction structure facilitates efficient separation and reduces the recombination rate of photoexcited charge carriers, thereby enhancing the photocatalytic performance of the composite photocatalyst. The Z-scheme heterojunction preserves the redox potential of the photocatalysts to a significant degree, owing to its unique mode of photoexcited charge carrier transfer. As depicted in Figure 3b, oxidation reactions occur on the photocatalyst A with a higher oxidation potential, while reduction reactions take place on photocatalyst B with a higher reduction potential.
Li et al. [63] demonstrated this concept by fabricating a direct Z-scheme heterojunction of rutile-TiO2/g-C3N4 using a simple solid-state grinding technique. HRTEM confirmed an intimate interface between TiO2 and g-C3N4, indicating successful heterojunction formation. Photoelectrochemical experiments further verified the effective separation and migration of photoexcited charge carriers, confirming the enhanced photocatalytic activity enabled by the direct Z-scheme mechanism. Heba et al. [64] synthesized a SnS2/SnO2 nanocomposite that forms a Z-scheme heterojunction. This heterojunction, resulting from the SnO2–SnS2 contact, facilitates Z-scheme electron transfer. It leads to a prolonged electron–hole recombination time, thereby improving charge carrier separation.

3.4. S-Scheme Heterojunction

The S-scheme (step-scheme) heterojunction was proposed by Fu et al. [65] to provide a more precise description of the charge transfer mechanism in certain heterojunctions, building upon and refining the direct Z-scheme concept. In an S-scheme system, one semiconductor acts as the reduction photocatalyst (RP) and the other as the oxidation photocatalyst (OP), each with a distinct work function [66]. The RP and OP can be n-type or p-type, but their band structures must be aligned such that the EF, VB, and CB of RP are all higher than those of the OP. This alignment drives electron transfer from the RP to the OP upon contact until their EF equilibrate, creating the necessary internal electric field for S-scheme charge transfer [59].
Figure 3c illustrates a typical S-scheme heterojunction. In this configuration, when the RP and OP come into contact, electrons diffuse from the RP (with a higher EF) to the OP (with a lower EF) until their EF equilibrates. This charge transfer leaves the OP negatively charged and the RP positively charged at the interface, forming a space charge region. Consequently, the energy bands of the OP bend downward (electron accumulation), while those of the RP bend upward (electron depletion) [59]. This band alignment creates an IEF directed from the RP to the OP.
Under light irradiation, the combined effects of band bending and the IEF create a selective charge transfer pathway where they prevent useless electrons in the conduction band (CB) of the RP from migrating to the CB of the OP and useless holes in the valence band (VB) of the OP from migrating to the VB of the RP. Instead, electrostatic attraction facilitates the recombination of these less reactive charges—specifically, the holes in the VB of the RP with the electrons in the CB of the OP. This selective recombination effectively leaves the most reactive electrons in the CB of the OP and the most reactive holes in the VB of the RP, thereby enhancing the overall redox capability of the system and suppressing detrimental bulk recombination within the individual semiconductors [59].
The S-scheme heterojunction significantly enhances photocatalytic performance through several key mechanisms. It preserves the strongest redox potentials by accumulating photoexcited electrons with high reduction potential in the CB of the RP and holes with high oxidation potential in the VB of the OP [58,67]. Furthermore, it improves the separation and transfer efficiency of charge carriers and broadens the light absorption range of the composite [68]. These benefits are driven by the synergistic effects of the IEF, band bending, and Coulombic attraction at the heterojunction interface [59].
Consequently, the S-scheme heterojunction effectively combines the efficient charge separation of a Type II system with the preserved redox power of a Z-scheme pathway. This unique combination results in superior charge carrier separation and strong redox potentials. Due to these advantages, researchers have successfully implemented S-scheme heterojunctions in various applications, including water splitting [69], carbon dioxide reduction [70], and photodegradation of organic pollutants [71]. Rezaei et al. [72] synthesized MoS2 nanoflowers, MoO3 nanobelts, and MoS2@MoO3 heterojunction. The composite demonstrated superior photocatalysis, achieving 96% removal of methylene blue (20 mg/L) in 6 h under visible light. The enhancement is attributed to a synergistic S-scheme mechanism that facilitates charge separation, generating OH, and O2 radicals for degradation.
Experimental techniques like X-ray diffraction (XRD) and X-ray photoelectron spectroscopy (XPS) are vital for confirming S-scheme heterojunction formation. For instance, Salehi et al. [73] observed reduced XRD peak intensities and narrowed peak widths in a TiO2/ZnIn2S4 composite, indicating successful coupling. XPS analysis further confirmed electron transfer from ZnIn2S4 to TiO2, validating the S-scheme mechanism [73,74].
Photoelectrochemical (PEC) studies are crucial for evaluating charge separation. A standard three-electrode setup—with a platinum counter electrode and a reference electrode (e.g., SCE or Ag/AgCl)—is typically used [28]. Electrochemical impedance spectroscopy (EIS) measures charge transfer resistance; a smaller arc radius in the Nyquist plot signifies lower impedance and better charge separation [52]. Feng et al. [75] demonstrated this, where a WO3/TiO2/CS-biochar composite showed the smallest arc radius and a longer electron lifetime in Bode plots, confirming enhanced charge dynamics (Figure 4a,b). Similarly, studies on BiVO4/TiO2 and CuO/TiO2 heterojunctions have shown extended electron lifetimes and reduced recombination [76,77]. Wang et al. [78] used photocurrent measurements to show that a ZrO2/TiO2 composite had a longer photocurrent decay time (6.001 s) than pure TiO2 (2.001 s), proving more effective charge separation (Figure 4c,d). These findings underscore how S-scheme heterojunctions enhance photocatalytic processes by optimizing charge carrier separation and utilization. These findings underscore the significant advancements and potential applications of S-scheme heterojunctions in improving photocatalytic processes through efficient charge carrier utilization and separation.

4. Ultrasonic-Assisted Fabrication of TiO2-Based Nanocomposite Photocatalysts

Numerous methods exist to synthesize TiO2-based composite photocatalysts with a reduced band gap and lower charge carrier recombination rates, including hydrothermal, sol–gel, and co-precipitation techniques [23,80,81,82,83]. Among these, ultrasonication has recently gained significant attention for its substantial benefits in photocatalyst synthesis [84]. Ultrasonication employs high-frequency sound waves (>20 kHz) to trigger cavitation in a liquid, where the rapid formation and collapse of bubbles generate extreme local conditions—exceeding 10,000 K and 1000 atm [85,86]. This energy facilitates nanoparticle formation, creates surface defects on the photocatalyst, and enhances the dispersion of active species on a support material [87,88].
Also known as the sonochemical process, ultrasonication is a simple and environmentally friendly technique. It provides precise control over nanostructure and material dispersion, and its operation at room temperature makes it ideal for using heat-sensitive precursors without risk of decomposition [12,79,85]. Additionally, operating at room temperature eliminates the need for high-temperature conditions that may decompose precursors used in photocatalyst synthesis, making it suitable for heat-sensitive materials.
This is particularly critical for the synthesis of advanced materials such as Metal–Organic Frameworks (MOFs) and Covalent Organic Frameworks (COFs), whose crystalline porous structures would collapse upon calcination. For instance, MIL-125(Ti) and NH2-MIL-125(Ti), are effective photocatalysts synthesized solvothermally below 150 °C; exposure to temperatures over 400 °C would destroy their structure, forming amorphous TiO2 and eliminating their superior adsorption and photocatalytic properties. Similarly, COFs like TFPT-COFs are synthesized at room temperature or with mild heating, and high temperatures would decompose their intricate organic frameworks and active sites [89,90].
Room temperature processing is also essential for photocatalysts using molecular sensitizers, such as Eosin Y-sensitized g-C3N4, where heat would carbonize the dye and nullify its function. Systems dependent on specific 2D morphologies, like BiOX nanosheets, likewise require low temperatures to preserve their structural integrity and activity [91,92].
Ultrasonication prevents particle aggregation and enhances photocatalyst performance. Studies demonstrate that it reduces charge transfer resistance (e.g., TiO2 from 15 kΩ to 7 kΩ) and creates more uniform, dispersed structures with higher surface areas [83].For instance, ultrasonication produced a TiO2/GO composite with a greater surface area (92 m2/g) and enhanced both methyl orange and congo red dyes degradation [93]. Similarly, an rGO/ZnO/TiO2 composite synthesized with ultrasonication had a more even element distribution and a 15% higher degradation rate for crystal violet dye than one made with conventional mixing [87].
Govinda Raj et al. [79] synthesized a novel TiO2/MoS2/BiVO4 composite using an ultrasonic-assisted hydrothermal approach. Their study found that ultrasonication significantly enhanced surface area, pore volume, and photocatalytic activity compared to the standard hydrothermal method. SEM and Brunauer–Emmett–Teller (BET) analyses confirmed the formation of well-dispersed, nanosized particles with reduced charge carrier recombination. As shown in Figure 4e,f, the ultrasonicated TiO2 had a defined nanosized morphology, while the conventional sample was agglomerated. BET analysis measured a higher surface area for both ultrasonicated TiO2 (52.30 m2/g) and the composite (132.70 m2/g), compared to their conventional counterparts (34.80 and 96.70 m2/g, respectively). This increased surface area suggests reduced charge carrier recombination [79]. Larger pore volumes were also found in the ultrasonicated samples (TiO2: 0.15 cm3/g; composite: 0.22 cm3/g) versus the conventional ones (0.13 and 0.18 cm3/g). Consequently, the ultrasonicated composite provides more reactive sites for pollutant adsorption, enhancing its overall photocatalytic performance [79].
Figure 4g shows that the ultrasonic-assisted hydrothermal method produces TiO2 and its composite with lower photoluminescence (PL) intensity, suggesting enhanced charge carrier separation [79]. This is supported by EIS analysis in Figure 4h, where a smaller Nyquist plot arc radius for the sonicated samples indicates improved charge transfer and photocatalytic activity. Additionally, the band gap ( E g ) was reduced for sonicated TiO2 (3.02 eV) and the composite (2.84 eV), compared to their conventionally prepared counterparts (3.20 eV and 2.94 eV, respectively) [79].
Researchers have widely explored ultrasonication for synthesizing various photocatalysts. For instance, studies have successfully produced composites like PTA/TiO2, Fe2O3/TiO2, and ZnO/TiO2 using ultrasonic-assisted methods such as impregnation and spray pyrolysis [12,94,95]. These methods often yield materials with enhanced properties, including improved visible light absorption and reduced crystallite sizes [95,96]. The benefits are clear in direct comparisons. Ultrasonication increased the surface area of TiO2 from 59 to 101 m2/g and reduced its particle size from 70–80 nm to 20–40 nm, leading to better performance [86,97]. It also prevents agglomeration, aids in forming complex structures like core–shell microspheres, and improves heterojunction formation in composites like Bi2WO6/TiO2, resulting in a lower band gap energy [98,99,100]. Table 1 summarizes various TiO2-based composite photocatalysts synthesized using ultrasonication. The cited studies consistently demonstrate that this technique enhances material properties and photocatalytic performance by controlling particle size, morphology, and surface characteristics. In conclusion, ultrasonication is a versatile method for producing efficient TiO2 composites, offering reduced charge carrier recombination and enhanced activity across various applications.

4.1. Parameters of Ultrasonication

In sonochemistry, the effects of ultrasonication parameters including power, frequency, temperature, and duration on photocatalyst size and morphology are actively studied. This review aims to outline the optimization of these variables to achieve materials with highly desirable characteristics.

4.1.1. Time

Ultrasonication duration is a critical parameter in photocatalyst synthesis, directly influencing nanoparticle dispersion and size distribution [143]. While extended times generally promote finer particles and more homogeneous dispersions, excessive exposure can cause aggregation, thermal degradation, and unintended reactions, necessitating careful optimization [143].
The time-dependent effects are evident in specific studies. Rasouli et al. [12] found that 8 min was optimal for dispersing a Fe2O3/TiO2 nanocomposite; longer times introduced excessive energy, promoting agglomeration via processes like Ostwald ripening. Similarly, Karamifar et al. [139] identified 10 min as ideal for dispersing a TiO2/MWCNT/Pani nanocomposite, with benzene removal efficiency improving from 58.50% to 69.53% as time increased. In the synthesis of ZnO/TiO2/ZrO2, Abrinaei and Aghabeygi [144] observed that different durations (5–30 min) yielded platelet-like structures with particle sizes between 30 and 50 nm.
For pure TiO2, Kumar et al. [145] reported a particle size reduction from 982 nm to 560 nm as sonication time increased from 5 to 25 min, attributing this to cavitation effects. Beyond 25 min, no significant further reduction occurred. Tiple et al. [146] also observed a decrease in particle size (554 nm to 316 nm) with time, selecting 45 min as the optimal balance between effectiveness and cost. Estrada-Monje et al. [147] confirmed that even short durations (5–10 min) at high power (400 W) achieved thorough dispersion.
Conversely, very long durations can alter material structure. Elavarasan et al. [148] sonicated TiO2 for 3 to 7 h at 90 °C and 500 W, finding that longer times reduced crystallinity, increased interplanar distance due to oxygen vacancies, and decreased the band gap from 2.97 to 2.90 eV. Collectively, these studies underscore that ultrasonication time must be optimized to achieve desired morphological and structural properties, as its effects are both significant and complex.

4.1.2. Power or Intensity

Ultrasonication intensity, defined as the power per unit area of the ultrasonic wave, is a critical parameter influencing nanoparticle dispersion and composite homogeneity [149]. It directly affects nanoparticle size distribution and suspension stability [144]. While higher intensities effectively break down agglomerates for finer, more uniform dispersions [87,150], excessive intensity can cause particle fragmentation or matrix degradation, necessitating careful optimization [150]. This balance is demonstrated in several studies.
Rasouli et al. [12] found that increasing ultrasonication power from 50 W to 70 W improved the photocatalytic degradation of cefixime from 64.91% to 78.80% for a Fe2O3/TiO2 composite, due to more violent cavitation bubble collapse. However, at 90 W, efficiency dropped to 69.53%, likely due to aggregation from increased thermodynamic stability. Karamifar et al. [139] observed a similar trend, with benzene removal efficiency increasing from 64.91% to 69.53% as power rose from 50 W to 90 W for a TiO2/MWCNT/Pani nanocomposite. Tiple et al. [146] reported a direct correlation between power and size reduction, with TiO2 particle size decreasing from 780 nm to 352.90 nm as power increased from 80 W to 120 W.
Higher power levels generally enhance performance by improving dispersion. Tian et al. [98] achieved a V2O5 separation rate of 99.90% with ultrasonic-assisted alkali leaching, compared to 32.10% without, noting that increased power (20–140 W) significantly boosted separation. Similarly, Abd-Rabboh et al. [106] found that photocatalytic efficiency for rhodamine B degradation increased with ultrasonic intensity up to 300 W for a BiVO4/TiO2 composite. Conversely, Qayyum et al. [151] demonstrated that power must be matched to the desired property; lower power (24 W) produced porous TiO2 with a high surface area (156 m2/g), while higher power (56 W) yielded non-porous, aggregated TiO2 with a low surface area (11 m2/g).

4.1.3. Frequency

Ultrasonic frequency governs cavitation bubble dynamics, including their number, size, distribution, and chemical activity [152]. Higher frequencies (200–500 kHz) produce numerous small, chemically active bubbles ideal for enhancing reactions. In contrast, lower frequencies (20–100 kHz) generate larger, short-lived bubbles that facilitate the rapid synthesis of nano-photocatalysts without complex agents [152,153].
Bandeira et al. [154] synthesized TiO2 using audible sound frequencies (174 Hz, 5 kHz, 15 kHz), finding that all frequencies enhanced crystallinity and produced smaller particles compared to non-sonicated samples. Performance improved with frequency, yielding photocatalytic rates of 0.016 min−1 (174 Hz), 0.050 min−1 (5 kHz), and 0.055 min−1 (15 kHz). Oliveira et al. [155] compared frequencies of 274 Hz and 409 Hz, finding that 409 Hz produced larger crystallites (6.50 nm vs. 5.66 nm) and a lower band gap (e.g., 3.14 eV vs. 3.19 eV at 473 °C), attributed to greater amplitude and spectral homogeneity.
Despite its advantages, ultrasonication also poses certain disadvantages, such as the high energy input required during the process, which can potentially damage photocatalysts and influence their properties. As demonstrated by Dong et al. [156] with egg-white protein-curcumin nanoparticles, an optimal duration (12 min) enhances key characteristics like solubility and encapsulation efficiency. Beyond this threshold, however, excessive sonication induces detrimental effects such as aggregation and increased crystallinity, compromising the product’s integrity.
Rasouli et al. [12] investigated the effect of ultrasonication time (3–8 min) on the efficiency, dimension, and structure of the photocatalyst. They found that as the ultrasonic time increased from 5 to 8 min, the continual energy input inhibited the growth of the α-Fe2O3@TiO2 nanocomposites. Furthermore, the photocatalytic removal efficiency increased from 64.91% to 78.8% as the ultrasonic power was raised from 50 W to 70 W. However, a further increase to 90 W reduced the efficiency to 69.53%. This optimal power at 70 W is critical for forming a regular and homogeneous α-Fe2O3@TiO2 nanocomposite, whereas other powers, such as 50 W, yield irregular, agglomerated structures and result in poorer catalytic performance. The initial improvement is attributed to enhanced cavitation bubble collapse at 70 W, which generates stronger shock waves that prevent agglomeration. Conversely, powers exceeding 70 W promote rapid primary nucleation, increasing thermodynamic stability and causing faster particle aggregation due to an excessive energy input.
Karamifar et al. [139] synthesized a TiO2/MWCNT/Pani nanocomposite via an ultrasonic-assisted method. They determined that a 10 min treatment was optimal, as prolonged ultrasonication caused irregular agglomeration due to the high surface energy of nanostructures driving excessive nucleation. Furthermore, when evaluating benzene degradation, catalyst performance improved from 64.91% to 69.53% as ultrasonic power increased from 50 W to 90 W, establishing 90 W as the optimal power for producing a regular and effective nanocomposite. Thus, while ultrasonication offers significant benefits in enhancing photocatalytic performance, careful parameter optimization is crucial to mitigate potential drawbacks and maximize its effectiveness in material synthesis.

5. Photocatalytic Activity of TiO2-Based Nanocomposite Photocatalysts

Persistent toxic organic pollutants from agricultural, pharmaceutical, and textile industries are prevalent in various water sources, necessitating effective removal strategies. Studying the photocatalytic degradation of such pollutants, Chau et al. [28] used a WO3/TiO2 composite to decompose Reactive Black 5 dye. The degradation was confirmed by the solution’s discoloration and the disappearance of its characteristic UV-Vis absorption peak at 597 nm, which is associated with the dye’s chromophore group. The consistent decrease in this peak indicates that the chromophore was the primary site of oxidation. Similarly, Manivannan et al. [24] observed a gradual reduction in the intensity of the 515 nm peak of Eosin Yellow dye, corresponding to the solution fading over time.
Chemical oxygen demand (COD) is a critical parameter for assessing the complete mineralization of organic compounds and gauging water pollution levels [25]. It measures the oxygen required to chemically oxidize organics in a solution [12]. For instance, Feng et al. [75] used COD to evaluate a WO3@TiO2/CS-biochar composite, achieving 97.56% COD removal after two hours of light irradiation. Total organic carbon (TOC) is another key mineralization indicator [157]. In a study on potato wastewater, Li et al. [158] reported COD and TOC removal rates of 82.34% and 63.36%, respectively, using an ultrasonic-assisted TiO2/Fe3O4 photocatalyst.
The photostability and reusability of photocatalysts are crucial for practical application, as photo-corrosion can lead to material leaching, impacting water safety [159]. Rasouli et al. [12] assessed α-Fe2O3/TiO2 reusability over five cycles, noting a performance decline from 98.04% to 94.10%. This was attributed to material loss during recovery and the blocking of active sites by adsorbed dyes or by-products [12,157]. Importantly, the photocatalyst’s morphology and crystal structure remained stable, confirming its suitability for repeated use in wastewater treatment [12].
Photostability is often assessed through photocurrent retention. For example, Debnath et al. [55] demonstrated high photostability with 99% photocurrent retention after two hours under continuous light. Understanding the photocatalytic mechanism requires identifying the primary reactive oxygen species (ROS) through quenching experiments. Key ROS include the superoxide anion (O2), photogenerated hole (h+), and hydroxyl radical (•OH). Specific scavengers are used to trap these species: p-benzoquinone (BQ) or a nitrogen gas for O2; ammonium oxalate, hydrogen (H2O2), and disodium ethylene diamine tetraacetate (EDTA-2Na) for h+; and tertbutanol or isopropyl alcohol for OH. Silver nitrate (AgNO3) traps electrons (e) and sodium azide is used for singlet oxygen using [23,28,52,125,160,161,162].
Debnath et al. [55] used electron paramagnetic resonance (EPR) with DMPO as a spin-trapping agent to confirm the generation of both O2 and OH radicals [139], which exhibit distinct characteristic peaks [139]. Similarly, Hou et al. [160] detected six distinct peaks of O2 and four characteristic peaks of OH via EPR [75], as shown in Figure 5a,b. Their study compared the formation rates of these radicals for TiO2/rGO and Ag/TiO2/rGO against pure TiO2, correlating the higher radical production in the composites with their enhanced photocatalytic performance. Figure 5c,d show that the EPR signal strength for O2 and OH was consistently higher for the composites under different illumination periods.
Mahamud et al. [81] investigated photocatalytic performance with and without scavengers. Without scavengers, efficiency was 99.90%, but it decreased to 83.20%, 92.00%, and 78.90% with the addition of methanol, sodium bicarbonate, and AgNO3, respectively. These results indicate that O2 and OH play more essential roles in photocatalytic activity than h+. Similarly, Karamifar et al. [139] reported 84% benzene removal without scavengers, which dropped to 73.58%, 58.50%, 20.70%, and 29.71% with AgNO3, EDTA, ethanol, and BQ, respectively. They concluded that OH and O2 are the primary reactive species.
Wafi et al. [140] utilized photoluminescence (PL) with terephthalic acid to trap OH, observing an increase in PL intensity at 430 nm over time, which confirmed OH generation. Correspondingly, Eskandari et al. [161] found that the PL signal for OH increased proportionally with irradiation time (30–240 min), affirming it as the predominant species in the photodegradation of methylene blue and congo red dyes.

Factors Affecting the Photodegradation Performance

The efficiency of photocatalytic degradation is governed by several parameters, including pH, photocatalyst loading, initial contaminant concentration, light source, and temperature. Identifying optimal conditions is essential for maximizing process efficiency. This section discusses key parameters, starting with catalyst dosage, which is critical for cost-effectiveness. Magdalane et al. [163] observed that the photodegradation efficiency of SnO2/TiO2 increased with dosage up to an optimum of 50 mg. Beyond this point, efficiency declined, likely due to light shielding effects that reduce penetration and active radical production [161,163]. Similarly, Yusuff et al. [164] found that methylene blue degradation improved with increased ZnO-TiO2 dosage up to 1.20 g, enhancing the availability of active sites and oxidizing agents. Further increases beyond this dosage yielded no additional improvement.
The initial dye concentration also significantly influences photodegradation efficiency. Mahamud et al. [81] found maximum efficiency at a lower MB concentration (10 mg/L) using a zeolite-supported CdS/TiO2/CeO2 composite, attributing this to enhanced dye adsorption and dispersion on the photocatalyst surface. Conversely, Eskandari et al. [161] observed decreased photodegradation efficiency as the concentrations of Methylene Blue and Congo Red increased from 10 to 30 ppm. This decline is primarily due to increased solution turbidity, which limits light penetration and active site availability on the photocatalyst surface [161], thereby reducing the generation of ROS [81,165].
Photocatalyst activation is dependent on specific light conditions. For instance, TiO2, with a band gap of 3.20 eV, requires UV light for activation. In contrast, composites like Fe3O4@SiO2@g-C3N4/TiO2 can function efficiently under both UV and visible light [166]. Kalikeri et al. [25] demonstrated that BFO/TiO2 performs better under solar light than UV light alone due to this dual activation capability. Light intensity and irradiation time are also critical. Huang et al. [167] showed that TiO2/CoFe2O4 degrades pollutants faster under sunlight than UV light.
Eskandari et al. [161] confirmed that longer irradiation times enhance dye degradation by increasing the production of active species. Duta et al. [168]. investigated CuInS2/TiO2/SnO2 and found that higher visible light intensity (32 W/m2) yielded a rate of 0.0028 s−1, compared to 0.0025 s−1 at a lower intensity (9 W/m2). A mixed UV/visible light source achieved even higher rates (0.0032–0.0035 s−1). Similarly, Moradi et al. [169] showed that increasing light intensity from 10 to 40 W/m2 raised the dye removal percentage to 85.40%, as more energy is delivered to activate the photocatalyst.
Industrial wastewater pH varies significantly and strongly influences photodegradation rates [165]. A photocatalyst’s acid and alkali resistance is therefore crucial for practical application [23]. Gao et al. [23] found that TiO2/rGO/TiN maintained stable performance across a broad pH range (3–11). pH affects the surface charge of both the photocatalyst and the pollutant, governing their electrostatic interaction. Kumar et al. [170] observed that at low pH, the positively charged TiO2 surface effectively adsorbs anionic dyes, enhancing degradation. Conversely, at high pH, the negatively charged surface repels these dyes.
The point of zero charge (pHpzc) is key to understanding this behavior. For Ag/TiO2/rGO, the pH_pzc is between 5 and 6 [160]. At a solution pH below this range, the surface becomes positively charged, attracting anions like Cr2O72−. Above the pHpzc, the negatively charged surface causes repulsion. Moradi et al. [169] illustrated this with AB172 dye (pKa = 6), where the dye’s charge shifts from positive (pH < 6) to negative (pH > 6), altering its interaction with the catalyst.
The molecular structure of the contaminant itself directly impacts degradation efficiency. Nguyen et al. [157] found that the mineralization rate of methyl orange (44.2%) after 3 h was lower than that of Rhodamine B (48%) and Methylene Blue (59%). This was attributed to the resilient azo group in methyl orange, which is difficult to cleave, underscoring how molecular composition dictates degradation kinetics [157].

6. Conclusions and Future Perspectives of Titanium Dioxide (TiO2)-Based Photocatalysts

Researchers continue to develop strategies to overcome the limitations of unmodified TiO2 photocatalysts, focusing on enhancing their activity and stability. This review has summarized the synthesis and performance of TiO2-based composite photocatalysts fabricated via ultrasonication techniques. These composites effectively extend the optical absorption of TiO2 from the UV region to the visible light region by reducing its band gap energy. Furthermore, the heterojunctions formed within these structures significantly suppress the recombination of photoexcited charge carriers, leading to superior photodegradation efficiency.
To ensure the commercial viability of these materials, future research must address key practical challenges. The difficulty in recovering powdered TiO2 photocatalysts from treated wastewater poses a risk of secondary contamination and limits reusability. Developing composite materials that facilitate easy separation and recovery is therefore essential. Concurrently, issues such as optimizing photocatalyst–contaminant contact and maintaining a high active surface area must be resolved to ensure consistently high performance.
Beyond efficiency, more attention must be paid to the potential environmental impact of the photocatalytic process. The toxicity of transformation products formed from incomplete degradation of organic contaminants is often overlooked. Given their nanoscale dimensions (1–100 nm), these photocatalysts have a high potential for subcellular interaction. Future work should prioritize elucidating detailed degradation pathways and identifying intermediate compounds. Comprehensive assessments using cytotoxicity, phytotoxicity, and genotoxicity assays are crucial to evaluate the safety of treated wastewater, particularly if it is intended for reuse in agricultural applications.

Author Contributions

J.H.F.C.: Data curation; Formal analysis; Investigation; Validation; Writing-original draft; Writing—review & editing. E.D.H.K.: Writing—review & editing. J.C.C.: Writing—review & editing. C.W.L.: Conceptualization; Funding acquisition; Project administration; Resources; Validation; Writing—review & editing; Supervision. J.C.J.: Supervision. Y.L.: Writing—review & editing. P.X.: Writing—review & editing. I.A.B.: Resources; Validation. A.K.: Supervision. All authors have read and agreed to the published version of the manuscript.

Funding

This research work was financially supported by the Ministry of Higher Education, Malaysia for niche area research under the Higher Institution Centre of Excellence (HiCoE) program (JPT(BKPI)1000/016/018/28 Jld.3(2) and NANOCAT-2024D). Besides, the authors extend their appreciation to the National Natural Science Foundation of China (42367064), the Yunnan Thousand Youth Talent Program (YNQR-QNRC-2018-049), and Deanship of Research and Graduate Studies at King Khalid University for funding this work through Large Research Project under grant number RGP. 2/622/46.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. UNESCO World Water Assesment Programme; Connor, R. The United Nations World Water Development Report 2024: Water for Prosperity and Peace; SIDALC: Turrialba, Costa Rica, 2024. [Google Scholar]
  2. Dawood, A.F.; Khalil, M.A.A.K. Removal of basic fuchsine dye using (TiO2/MWCNTs) nanomaterial. Mater. Today Proc. 2022, 49, 2888–2897. [Google Scholar] [CrossRef]
  3. Madkhali, N.; Prasad, C.; Malkappa, K.; Choi, H.Y.; Govinda, V.; Bahadur, I.; Abumousa, R.A. Recent update on photocatalytic degradation of pollutants in waste water using TiO2-based heterostructured materials. Results Eng. 2023, 17, 100920. [Google Scholar] [CrossRef]
  4. Vakili, A.; Zinatizadeh, A.A.; Rahimi, Z.; Zinadini, S.; Mohammadi, P.; Azizi, S.; Karami, A.; Abdulgader, M. The impact of activation temperature and time on the characteristics and performance of agricultural waste-based activated carbons for removing dye and residual COD from wastewater. J. Clean. Prod. 2023, 382, 134899. [Google Scholar] [CrossRef]
  5. Tsai, C.-G.; Tseng, W.J. Preparation of TiN–TiO2 composite nanoparticles for organic dye adsorption and photocatalysis. Ceram. Int. 2020, 46, 14529–14535. [Google Scholar] [CrossRef]
  6. Ali, A.; Shoeb, M.; Li, B.; Khan, M.A. Photocatalytic degradation of antibiotic drug and dye pollutants under visible-light irradiation by reduced graphene oxide decorated MoO3/TiO2 nanocomposite. Mater. Sci. Semicond. Process. 2022, 150, 106974. [Google Scholar] [CrossRef]
  7. Ren, X.; Chen, R.; Ding, S.; Fu, N. Preparation and photocatalytic performance of a magnetically recyclable ZnFe2O4@TiO2@Ag2O p-n/Z-type tandem heterojunction photocatalyst: Degradation pathway and mechanism. Colloids Surf. A Physicochem. Eng. Asp. 2023, 658, 130604. [Google Scholar] [CrossRef]
  8. Fawzi Suleiman Khasawneh, O.; Palaniandy, P. Removal of organic pollutants from water by Fe2O3/TiO2 based photocatalytic degradation: A review. Environ. Technol. Innov. 2021, 21, 101230. [Google Scholar] [CrossRef]
  9. Okeke, E.S.; Ezeorba, T.P.C.; Okoye, C.O.; Chen, Y.; Mao, G.; Feng, W.; Wu, X. Environmental and health impact of unrecovered API from pharmaceutical manufacturing wastes: A review of contemporary treatment, recycling and management strategies. Sustain. Chem. Pharm. 2022, 30, 100865. [Google Scholar] [CrossRef]
  10. Wilkinson, J.L.; Boxall, A.B.A.; Kolpin, D.W.; Leung, K.M.Y.; Lai, R.W.S.; Galban-Malag, C.; Adell, A.D.; Mondon, J.; Metian, M.; Marchant, R.A.; et al. Pharmaceutical pollution of the world’s rivers. Proc. Natl. Acad. Sci. USA 2022, 119, e2113947119. [Google Scholar] [CrossRef] [PubMed]
  11. Thakur, A.K.; Kumar, R.; Kumar, A.; Shankar, R.; Khan, N.A.; Gupta, K.N.; Ram, M.; Arya, R.K. Pharmaceutical waste-water treatment via advanced oxidation based integrated processes: An engineering and economic perspective. J. Water Process Eng. 2023, 54, 103977. [Google Scholar] [CrossRef]
  12. Rasouli, K.; Alamdari, A.; Sabbaghi, S. Ultrasonic-assisted synthesis of α-Fe2O3@TiO2 photocatalyst: Optimization of effective factors in the fabrication of photocatalyst and removal of non-biodegradable cefixime via response surface methodology-central composite design. Sep. Purif. Technol. 2023, 307, 122799. [Google Scholar] [CrossRef]
  13. Rani, M.; Shanker, U.; Jassal, V. Recent strategies for removal and degradation of persistent & toxic organochlorine pesticides using nanoparticles: A review. J. Environ. Manag. 2017, 190, 208–222. [Google Scholar] [CrossRef]
  14. Dehghani, M.H.; Ahmadi, S.; Ghosh, S.; Khan, M.S.; Othmani, A.; Khanday, W.A.; Gökkuş, Ö.; Osagie, C.; Ahmaruzzaman, M.; Mishra, S.R.; et al. Sustainable remediation technologies for removal of pesticides as organic micro-pollutants from water environments: A review. Appl. Surf. Sci. Adv. 2024, 19, 100558. [Google Scholar] [CrossRef]
  15. Swathy, K.; Vivekanandhan, P.; Yuvaraj, A.; Sarayut, P.; Kim, J.S.; Krutmuang, P. Biodegradation of pesticide in agricultural soil employing entomopathogenic fungi: Current state of the art and future perspectives. Heliyon 2024, 10, e23406. [Google Scholar] [CrossRef]
  16. Dhiman, N.; Chaudhary, S.; Singh, A.; Chauhan, A.; Kumar, R. Sustainable degradation of pharmaceutical waste using different fungal strains: Enzyme induction, kinetics and isotherm studies. Environ. Technol. Innov. 2022, 25, 102156. [Google Scholar] [CrossRef]
  17. Stylianou, M.; Christou, A.; Michael, C.; Agapiou, A.; Papanastasiou, P.; Fatta-Kassinos, D. Adsorption and removal of seven antibiotic compounds present in water with the use of biochar derived from the pyrolysis of organic waste feedstocks. J. Environ. Chem. Eng. 2021, 9, 105868. [Google Scholar] [CrossRef]
  18. Hu, Y.; Cai, X.; Du, R.; Yang, Y.; Rong, C.; Qin, Y.; Li, Y.-Y. A review on anaerobic membrane bioreactors for enhanced valorization of urban organic wastes: Achievements, limitations, energy balance and future perspectives. Sci. Total Environ. 2022, 820, 153284. [Google Scholar] [CrossRef]
  19. Yang, M.; Watson, J.; Wang, Z.; Si, B.; Jiang, W.; Zhou, B.; Zhang, Y. Understanding and design of two-stage fermentation: A perspective of interspecies electron transfer. Renew. Sustain. Energy Rev. 2022, 168, 112891. [Google Scholar] [CrossRef]
  20. Duan, Y.; Zhao, J.; Qiu, X.; Deng, X.; Ren, X.; Ge, W.; Yuan, H. Coagulation performance and floc properties for synchronous removal of reactive dye and polyethylene terephthalate microplastics. Process Saf. Environ. Prot. 2022, 165, 66–76. [Google Scholar] [CrossRef]
  21. Sun, Y.; Li, D.; Lu, X.; Sheng, J.; Zheng, X.; Xiao, X. Flocculation of combined contaminants of dye and heavy metal by nano-chitosan flocculants. J. Environ. Manag. 2021, 299, 113589. [Google Scholar] [CrossRef]
  22. Peng, C.; Chen, L.; Wu, X.; Wei, X.; Tehrim, A.; Dai, M.; Xu, S. Identification of adsorption or degradation mechanism for the removal of different ionic dyes with iron-carbon micro-electrolysis process. J. Environ. Chem. Eng. 2021, 9, 105690. [Google Scholar] [CrossRef]
  23. Gao, Y.; Wang, Y.; Sun, H.; Feng, Y.; Lu, Y.; Zhao, Y.; Chen, L. A novel macro-meso-micro intimate heterogeneous structure constructed 3DOM TiO2/rGO/TiN composite for efficiently solar photocatalysis degradation of organic dye pollutants. Appl. Surf. Sci. 2023, 621, 156774. [Google Scholar] [CrossRef]
  24. Manivannan, R.; Ryu, J.; Son, Y.-A. Photo discoloration of eosin yellow dye under visible light using TiO2@TPPS nanocomposite synthesized via ultrasonic assisted method. Colloids Surf. A Physicochem. Eng. Asp. 2021, 608, 125601. [Google Scholar] [CrossRef]
  25. Kalikeri, S.; Kodialbail, V.S. Visible light active Bismuth ferrite embedded TiO2 nanocomposite structures for dye mineralization by photocatalysis-A strategy to harness solar energy for remediation of water contaminated with mixture of dyes. Surf. Interfaces 2023, 36, 102492. [Google Scholar] [CrossRef]
  26. Xu, H.; Hao, Z.; Feng, W.; Wang, T.; Fu, X. The floating photocatalytic spheres loaded with weak light-driven TiO2-based catalysts for photodegrading tetracycline in seawater. Mater. Sci. Semicond. Process. 2022, 144, 106610. [Google Scholar] [CrossRef]
  27. Hafeez, H.Y.; Mohammed, J.; Suleiman, A.B.; Ndikilar, C.E.; Sa’id, R.S.; Muhammad, I. Insights into hybrid TiO2-g-C3N4 heterostructure composite decorated with rGO sheet: A highly efficient photocatalyst for boosted solar fuel (hydrogen) generation. Chem. Phys. Impact 2023, 6, 100157. [Google Scholar] [CrossRef]
  28. Chau, J.H.F.; Lee, K.M.; Pang, Y.L.; Abdullah, B.; Juan, J.C.; Leo, B.F.; Lai, C.W. Photodegradation assessment of RB5 dye by utilizing WO3/TiO2 nanocomposite: A cytotoxicity study. Environ. Sci. Pollut. Res. 2022, 29, 22372–22390. [Google Scholar] [CrossRef]
  29. Jiang, Q.; Huang, J.; Ma, B.; Yang, Z.; Zhang, T.; Wang, X. Recyclable, hierarchical hollow photocatalyst TiO2@SiO2 composite microsphere realized by raspberry-like SiO2. Colloids Surf. A Physicochem. Eng. Asp. 2020, 602, 125112. [Google Scholar] [CrossRef]
  30. Chau, J.H.F.; Lai, C.W.; Leo, B.F.; Juan, J.C.; Johan, M.R. Advanced photocatalytic degradation of acetaminophen using Cu2O/WO3/TiO2 ternary composite under solar irradiation. Catal. Commun. 2022, 163, 106396. [Google Scholar] [CrossRef]
  31. Pant, B.; Prasad Ojha, G.; Acharya, J.; Park, M. Ag3PO4-TiO2-Carbon nanofiber Composite: An efficient Visible-light photocatalyst obtained from electrospinning and hydrothermal methods. Sep. Purif. Technol. 2021, 276, 119400. [Google Scholar] [CrossRef]
  32. Zhang, Y.; Zhang, G.-L.; Wang, Y.-T.; Ma, Z.; Yang, T.-Y.; Zhang, T.; Zhang, Y.-H. In-situ synthesized porphyrin polymer/TiO2 composites as high-performance Z-scheme photocatalysts for CO2 conversion. J. Colloid Interface Sci. 2021, 596, 342–351. [Google Scholar] [CrossRef]
  33. Kanakaraju, D.; Chandrasekaran, A. Recent advances in TiO2/ZnS-based binary and ternary photocatalysts for the degradation of organic pollutants. Sci. Total Environ. 2023, 868, 161525. [Google Scholar] [CrossRef]
  34. Aravind, M.; Amalanathan, M.; Aslam, S.; Noor, A.E.; Jini, D.; Majeed, S.; Velusamy, P.; Alothman, A.A.; Alshgari, R.A.; Saleh Mushab, M.S.; et al. Hydrothermally synthesized Ag-TiO2 nanofibers (NFs) for photocatalytic dye degradation and antibacterial activity. Chemosphere 2023, 321, 138077. [Google Scholar] [CrossRef]
  35. Divya, G.; Jaishree, G.; Siva Rao, T.; Prasanna Chippada, M.L.V.; Divya Lakshmi, K.V.; Sai Supriya, S. Improved catalytic efficiency by N-doped TiO2 via sol gel under microwave irradiation: Dual applications in degradation of dye and microbes. Hybrid Adv. 2022, 1, 100010. [Google Scholar] [CrossRef]
  36. Wang, X.; Xu, J.; Liu, S.; Yang, W.; Chen, Y.; Zhang, Y. Synthesis of TiO2/MOF-801(Zr) by a wet impregnation at room temperature for highly efficient photocatalytic reduction of Cr(VI). Solid State Sci. 2022, 129, 106912. [Google Scholar] [CrossRef]
  37. Zikriya, M.; Nadaf, Y.F.; Bharathy, P.V.; Renuka, C.G. Luminescent characterization of rare earth Dy3+ ion doped TiO2 prepared by simple chemical co-precipitation method. J. Rare Earths 2019, 37, 24–31. [Google Scholar] [CrossRef]
  38. Wei, X.; Cai, H.; Feng, Q.; Liu, Z.; Ma, D.; Chen, K.; Huang, Y. Synthesis of co-existing phases Sn-TiO2 aerogel by ultrasonic-assisted sol-gel method without calcination. Mater. Lett. 2018, 228, 379–383. [Google Scholar] [CrossRef]
  39. Patil, S.M.; Deshmukh, S.P.; More, K.V.; Shevale, V.B.; Mullani, S.B.; Dhodamani, A.G.; Delekar, S.D. Sulfated TiO2/WO3 nanocomposite: An efficient photocatalyst for degradation of Congo red and methyl red dyes under visible light irradiation. Mater. Chem. Phys. 2019, 225, 247–255. [Google Scholar] [CrossRef]
  40. Sekar, R.; Sivasamy, R.; Ricardo, B.; Manidurai, P. Ultrasonically synthesized TiO2/ZnS nanocomposites to improve the efficiency of dye sensitized solar cells. Mater. Sci. Semicond. Process. 2021, 132, 105917. [Google Scholar] [CrossRef]
  41. Kajitvichyanukul, P.; Nguyen, V.-H.; Boonupara, T.; Phan Thi, L.-A.; Watcharenwong, A.; Sumitsawan, S.; Udomkun, P. Challenges and effectiveness of nanotechnology-based photocatalysis for pesticides-contaminated water: A review. Environ. Res. 2022, 212, 113336. [Google Scholar] [CrossRef]
  42. Swedha, M.; Balasurya, S.; Syed, A.; Das, A.; Sudheer Khan, S. Continuous photocatalysis via Z-scheme based nanocatalyst system for environmental remediation of pharmaceutically active compound: Modification, reaction site, defect engineering and challenges on the nanocatalyst. J. Mol. Liq. 2022, 353, 118745. [Google Scholar] [CrossRef]
  43. Zhuang, Q.; Chen, H.; Zhang, C.; Cheng, S.; Dong, W.; Xie, A. Rapid chromium reduction by metal-free organic polymer photocatalysis via molecular engineering. J. Hazard. Mater. 2022, 434, 128938. [Google Scholar] [CrossRef] [PubMed]
  44. Ariza-Tarazona, M.C.; Villarreal-Chiu, J.F.; Barbieri, V.; Siligardi, C.; Cedillo-González, E.I. New strategy for microplastic degradation: Green photocatalysis using a protein-based porous N-TiO2 semiconductor. Ceram. Int. 2019, 45, 9618–9624. [Google Scholar] [CrossRef]
  45. Zeshan, M.; Bhatti, I.A.; Mohsin, M.; Iqbal, M.; Amjed, N.; Nisar, J.; AlMasoud, N.; Alomar, T.S. Remediation of pesticides using TiO2 based photocatalytic strategies: A review. Chemosphere 2022, 300, 134525. [Google Scholar] [CrossRef] [PubMed]
  46. Pourmadadi, M.; Rajabzadeh-Khosroshahi, M.; Eshaghi, M.M.; Rahmani, E.; Motasadizadeh, H.; Arshad, R.; Rahdar, A.; Pandey, S. TiO2-based nanocomposites for cancer diagnosis and therapy: A comprehensive review. J. Drug Deliv. Sci. Technol. 2023, 82, 104370. [Google Scholar] [CrossRef]
  47. Acharya, R.; Parida, K. A review on TiO2/g-C3N4 visible-light- responsive photocatalysts for sustainable energy generation and environmental remediation. J. Environ. Chem. Eng. 2020, 8, 103896. [Google Scholar] [CrossRef]
  48. Yu, Z.; Zhu, S.; Zhang, L.; Watanabe, S. Mesoporous single crystal titanium oxide microparticles for enhanced visible light photodegradation. Opt. Mater. 2022, 127, 112297. [Google Scholar] [CrossRef]
  49. Zhao, Y.; Linghu, X.; Shu, Y.; Zhang, J.; Chen, Z.; Wu, Y.; Shan, D.; Wang, B. Classification and catalytic mechanisms of heterojunction photocatalysts and the application of titanium dioxide (TiO2)-based heterojunctions in environmental remediation. J. Environ. Chem. Eng. 2022, 10, 108077. [Google Scholar] [CrossRef]
  50. Yuju, S.; Xiujuan, T.; Dongsheng, S.; Zhiruo, Z.; Meizhen, W. A review of tungsten trioxide (WO3)-based materials for antibiotics removal via photocatalysis. Ecotoxicol. Environ. Saf. 2023, 259, 114988. [Google Scholar] [CrossRef]
  51. Xiao, L.; Yang, Z.; Zhu, H.; Yan, G. Nanoflower-like BiOBr/TiO2 p-n heterojunction composites for enhanced photodegradation of formaldehyde and dyes. Inorg. Chem. Commun. 2022, 146, 110167. [Google Scholar] [CrossRef]
  52. 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]
  53. Chau, J.H.F.; Lai, C.W.; Leo, B.F.; Juan, J.C.; Lee, K.M.; Qian, X.; Badruddin, I.A.; Zai, J. Direct Z-scheme Cu2O/WO3/TiO2 nanocomposite as a potential supercapacitor electrode and an effective visible-light-driven photocatalyst. J. Environ. Manag. 2024, 363, 121332. [Google Scholar] [CrossRef]
  54. Qiu, P.; Xiong, J.; Lu, M.; Liu, L.; Li, W.; Wen, Z.; Li, W.; Chen, R.; Cheng, G. Integrated p-n/Schottky junctions for efficient photocatalytic hydrogen evolution upon Cu@TiO2-Cu2O ternary hybrids with steering charge transfer. J. Colloid Interface Sci. 2022, 622, 924–937. [Google Scholar] [CrossRef] [PubMed]
  55. Debnath, K.; Majumder, T.; Mondal, S.P. Photoelectrochemical study of hydrothermally grown vertically aligned rutile TiO2 nanorods. Chem. Phys. 2022, 561, 111609. [Google Scholar] [CrossRef]
  56. Tai, X.H.; Lai, C.W.; Yang, T.C.K.; Johan, M.R.; Lee, K.M.; Chen, C.-Y.; Juan, J.C. Highly effective removal of volatile organic pollutants with p-n heterojunction photoreduced graphene oxide-TiO2 photocatalyst. J. Environ. Chem. Eng. 2022, 10, 107304. [Google Scholar] [CrossRef]
  57. Chau, J.H.F.; Lai, C.W.; Leo, B.F.; Juan, J.C.; Lee, K.M.; Badruddin, I.A.; Kumar, A.; Sharma, G. Study of Calcination Temperature Influence on Physicochemical Properties and Photodegradation Performance of Cu2O/WO3/TiO2. Catalysts 2025, 15, 601. [Google Scholar] [CrossRef]
  58. Hao, L.; Teng, D.; Guo, X.; Wu, B.; Wan, J.; Zhang, J.; Yang, J.-H. Construction of TiO2/BiOCl S-Scheme heterojunction and photocatalytic degradation of norfloxacin. J. Photochem. Photobiol. A Chem. 2023, 444, 115004. [Google Scholar] [CrossRef]
  59. Deng, A.; Sun, Y.; Gao, Z.; Yang, S.; Liu, Y.; He, H.; Zhang, J.; Liu, S.; Sun, H.; Wang, S. Internal electric field in carbon nitride-based heterojunctions for photocatalysis. Nano Energy 2023, 108, 108228. [Google Scholar] [CrossRef]
  60. Bard, A.J. Photoelectrochemistry and heterogeneous photo-catalysis at semiconductors. J. Photochem. 1979, 10, 59–75. [Google Scholar] [CrossRef]
  61. Liao, G.; Li, C.; Liu, S.-Y.; Fang, B.; Yang, H. Z-scheme systems: From fundamental principles to characterization, synthesis, and photocatalytic fuel-conversion applications. Phys. Rep. 2022, 983, 1–41. [Google Scholar] [CrossRef]
  62. Xie, Z.; Liu, G.; Xie, L.; Wu, P.; Liu, H.; Wang, J.; Xie, Y.; Chen, J.; Lu, C.-Z. Promoting photocatalytic H2 evolution through interfacial charge separation on the direct Z-scheme ZnIn2S4/ZrO2 heterojunction. Int. J. Hydrog. Energy 2023, 48, 32782–32796. [Google Scholar] [CrossRef]
  63. Li, G.; Cai, Y.; Wang, X.; Zhang, L.; Xie, Q.; Chen, P.; Li, C.; Sun, J.; Li, T.; Dong, L. Direct Z-scheme heterojunction rutile-TiO2/g-C3N4 catalyst constructed by solid grinding method for photocatalysis degradation. Chem. Phys. 2022, 559, 111558. [Google Scholar] [CrossRef]
  64. El Sharkawy, H.M.; Abo El-Khair, M.A.; Morshedy, A.S. Construction of SnS2@SnO2 nanocomposite Z-scheme heterojunction for dual-functional photocatalysis: Green hydrogen generation and crystal violet degradation. Int. J. Hydrog. Energy 2025, 137, 471–486. [Google Scholar] [CrossRef]
  65. Fu, J.; Xu, Q.; Low, J.; Jiang, C.; Yu, J. Ultrathin 2D/2D WO3/g-C3N4 step-scheme H2-production photocatalyst. Appl. Catal. B Environ. 2019, 243, 556–565. [Google Scholar] [CrossRef]
  66. Yue, X.; Cheng, L.; Fan, J.; Xiang, Q. 2D/2D BiVO4/CsPbBr3 S-scheme heterojunction for photocatalytic CO2 reduction: Insights into structure regulation and Fermi level modulation. Appl. Catal. B Environ. 2022, 304, 120979. [Google Scholar] [CrossRef]
  67. Nie, C.; Wang, X.; Lu, P.; Zhu, Y.; Li, X.; Tang, H. Advancements in S-scheme heterojunction materials for photocatalytic environmental remediation. J. Mater. Sci. Technol. 2023, 169, 182–198. [Google Scholar] [CrossRef]
  68. Yang, S.; Lu, Q.; Wang, F.; Zhi, Y.; Chen, J.; Wang, Y.; Zhang, H.; Yin, H.; Sun, P.; Cao, W. S-scheme SnO/TiO2 heterojunction with high hole mobility for boosting photocatalytic degradation of gaseous benzene. Chem. Eng. J. 2023, 478, 147345. [Google Scholar] [CrossRef]
  69. Zhu, H.; Zhang, X.; Nie, Y.; Yang, D.; Xiang, G. 2D/2D Janus BiTeCl/GeSe vdW heterostructure as a robust high-performance S-scheme photocatalyst for water splitting. Appl. Surf. Sci. 2023, 635, 157694. [Google Scholar] [CrossRef]
  70. Xiao, Y.; Abulizi, A.; Okitsu, K.; Ren, T. Facile fabrication of SnO2 modified hierarchical BiOI S-scheme heterojunction photocatalyst with efficient activity for carbon dioxide reduction. J. Ind. Eng. Chem. 2023, 125, 317–324. [Google Scholar] [CrossRef]
  71. Huong, V.H.; Nguyen, V.-C.; Ha, M.N.; Pham, D.V.; Nguyen, T.B.; Ma, Y.-R.; Ngac, A.B.; Loan, T.T. A S-scheme heterojunction Fe-doped TiO2/SnO2 with rich oxygen vacancies for photo-Fenton degradation of Rhodamine B under visible light illumination. Opt. Mater. 2023, 140, 113864. [Google Scholar] [CrossRef]
  72. Rezaei, M.M.; Seyed Dorraji, M.S.; Hosseini, S.F.; Rasoulifard, M.H. S-scheme heterojunction of MoO3 nanobelts and MoS2 nanoflowers for photocatalytic degradation. Sci. Rep. 2025, 15, 10789. [Google Scholar] [CrossRef]
  73. Salehi Ghalehsefid, E.; Ghorbani Jahani, Z.; Aliabadi, A.; Ghodrati, M.; Khamesan, A.; Parsaei-Khomami, A.; Mousavi, M.; Hosseini, M.-A.; Ghasemi, J.B.; Li, X. TiO2 nanotube/ZnIn2S4 nanoflower composite with step-scheme heterojunction for efficient photocatalytic H2O2 production and organic dye degradation. J. Environ. Chem. Eng. 2023, 11, 110160. [Google Scholar] [CrossRef]
  74. Zhang, J.; Wang, X.; Dai, J.; Songsiriritthigul, P.; Oo, T.Z.; Zaw, M.; Lwin, N.W.; Aung, S.H.; Chen, F. Defect engineering in 0D/2D TiO2/g-C3N4 heterojunction for boosting photocatalytic degradation of tetracycline in a tetracycline/Cu2+ combined system. Colloids Surf. A Physicochem. Eng. Asp. 2024, 680, 132624. [Google Scholar] [CrossRef]
  75. Feng, X.; Li, X.; Su, B.; Ma, J. Two-step construction of WO3@TiO2/CS-biochar S-scheme heterojunction and its synergic adsorption/photocatalytic removal performance for organic dye and antibiotic. Diam. Relat. Mater. 2023, 131, 109560. [Google Scholar] [CrossRef]
  76. Song, C. Enhancing photocatalytic degradation of hydrolyzed polyacrylamide in oilfield wastewater using BiVO4/TiO2 heterostructure nano-photocatalyst under visible light irradiation. Int. J. Electrochem. Sci. 2023, 18, 100363. [Google Scholar] [CrossRef]
  77. Wang, R.; Cao, J.; Liu, J.; Zhang, Y. Synthesis of CuO@TiO2 nanocomposite and its photocatalytic and electrochemical properties. Application for treatment of azo dyes in industrial wastewater. Int. J. Electrochem. Sci. 2023, 18, 100316. [Google Scholar] [CrossRef]
  78. Wang, F.; Pan, K.; Wei, S.; Ren, Y.; Zhu, H.; Wu, H.-H.; Zhang, Q. Solvothermal preparation and characterization of ordered-mesoporous ZrO2/TiO2 composites for photocatalytic degradation of organic dyes. Ceram. Int. 2021, 47, 7632–7641. [Google Scholar] [CrossRef]
  79. Raj, M.G.; Vijayakumar, E.; Preetha, R.; Narendran, M.G.; Neppolian, B.; Bosco, A.J. Implanting TiO2 @MoS2/BiVO4 nanocomposites via sonochemically assisted photoinduced charge carriers promotes highly efficient photocatalytic removal of tetracycline. J. Alloys Compd. 2022, 929, 167252. [Google Scholar] [CrossRef]
  80. Li, D.; Zhao, Z.; Guo, J.; Shen, R.; Xu, F. Magnetic nanoscaled Fe3O4-CeO2-TiO2 composite: UV-Fenton reaction to degrade AO-7 dye. Inorg. Chem. Commun. 2023, 149, 110389. [Google Scholar] [CrossRef]
  81. Mahamud, M.; Taddesse, A.M.; Bogale, Y.; Bezu, Z. Zeolite supported CdS/TiO2/CeO2 composite: Synthesis, characterization and photocatalytic activity for methylene blue dye degradation. Mater. Res. Bull. 2023, 161, 112176. [Google Scholar] [CrossRef]
  82. Gajera, R.; Patel, R.V.; Yadav, A.; Labhasetwar, P.K. Adsorption of cationic and anionic dyes on photocatalytic flyash/TiO2 modified chitosan biopolymer composite. J. Water Process Eng. 2022, 49, 102993. [Google Scholar] [CrossRef]
  83. Yavuz, C.; Ela, S.E. Fabrication of g-C3N4-reinforced CdS nanosphere-decorated TiO2 nanotablet composite material for photocatalytic hydrogen production and dye-sensitized solar cell application. J. Alloys Compd. 2023, 936, 168209. [Google Scholar] [CrossRef]
  84. Bae, J.-h.; Do, S.-b.; Cho, S.-h.; Lee, K.-m.; Lee, S.-E.; Kim, T.-O. TiO2 treatment using ultrasonication for bubble cavitation generation and efficiency assessment of a dye-sensitized solar cell. Ultrason. Sonochem. 2022, 83, 105933. [Google Scholar] [CrossRef]
  85. Sun, H.; Qin, P.; Liang, Y.; Yang, Y.; Zhang, J.; Guo, J.; Hu, X.; Jiang, Y.; Zhou, Y.; Luo, L.; et al. Sonochemically assisted the synthesis and catalytic application of bismuth-based photocatalyst: A mini review. Ultrason. Sonochem. 2023, 100, 106600. [Google Scholar] [CrossRef]
  86. Kale, D.P.; Deshmukh, S.P.; Shirsath, S.R.; Bhanvase, B.A. Sonochemical preparation of multifunctional rGO-ZnS-TiO2 ternary nanocomposite and its application for CV dye removal. Optik 2020, 208, 164532. [Google Scholar] [CrossRef]
  87. Potle, V.D.; Shirsath, S.R.; Bhanvase, B.A.; Saharan, V.K. Sonochemical preparation of ternary rGO-ZnO-TiO2 nanocomposite photocatalyst for efficient degradation of crystal violet dye. Optik 2020, 208, 164555. [Google Scholar] [CrossRef]
  88. Zhang, W.; Tang, Y.; Lu, C.; Zou, J.; Ruan, M.; Yin, Y.; Qing, M.; Song, Q. Enhancement of catalytic activity in NH3-SCR reaction by promoting dispersibility of CuCe/TiO2-ZrO2 with ultrasonic treatment. Ultrason. Sonochem. 2021, 72, 105466. [Google Scholar] [CrossRef] [PubMed]
  89. Sharifzadeh, Z.; Morsali, A. Amine-Functionalized Metal-Organic Frameworks: From Synthetic Design to Scrutiny in Application. Coord. Chem. Rev. 2022, 459, 214445. [Google Scholar] [CrossRef]
  90. Wang, H.; Qian, C.; Liu, J.; Zeng, Y.; Wang, D.; Zhou, W.; Gu, L.; Wu, H.; Liu, G.; Zhao, Y. Integrating Suitable Linkage of Covalent Organic Frameworks into Covalently Bridged Inorganic/Organic Hybrids toward Efficient Photocatalysis. J. Am. Chem. Soc. 2020, 142, 4862–4871. [Google Scholar] [CrossRef]
  91. Wang, P.; Zong, L.; Guan, Z.; Li, Q.; Yang, J. PtNi Alloy Cocatalyst Modification of Eosin Y-Sensitized g-C3N4/GO Hybrid for Efficient Visible-Light Photocatalytic Hydrogen Evolution. Nanoscale Res. Lett. 2018, 13, 33. [Google Scholar] [CrossRef]
  92. Wang, Z.; Chu, Z.; Dong, C.; Wang, Z.; Yao, S.; Gao, H.; Liu, Z.; Liu, Y.; Yang, B.; Zhang, H. Ultrathin BiOX (X = Cl, Br, I) Nanosheets with Exposed {001} Facets for Photocatalysis. ACS Appl. Nano Mater. 2020, 3, 1981–1991. [Google Scholar] [CrossRef]
  93. Purkayastha, M.D.; Sil, S.; Singh, N.; Ray, P.P.; Darbha, G.K.; Bhattacharyya, S.; Mallick, A.I.; Majumder, T.P. Sonochemical synthesis of nanospherical TiO2 within graphene oxide nanosheets and its application as a photocatalyst and a Schottky diode. FlatChem 2020, 22, 100180. [Google Scholar] [CrossRef]
  94. Huang, Y.; Qiu, S.; Lian, H.; Xu, J. Rearrangement of α-terpineol by using PTA/TiO2 catalyst synthesized by ultrasonic-assisted impregnation method. Tetrahedron 2022, 108, 132659. [Google Scholar] [CrossRef]
  95. Wittawat, R.; Rittipun, R.; Jarasfah, M.; Nattaporn, B. Synthesis of ZnO/TiO2 spherical particles for blue light screening by ultrasonic spray pyrolysis. Mater. Today Commun. 2020, 24, 101126. [Google Scholar] [CrossRef]
  96. Balu, S.; Chen, Y.-L.; Yang, T.C.K.; Chen, J.-N.; Chen, S.-W. Effect of ultrasound-induced hydroxylation and exfoliation on P90–TiO2/g-C3N4 hybrids with enhanced optoelectronic properties for visible-light photocatalysis and electrochemical sensing. Ceram. Int. 2020, 46, 18002–18018. [Google Scholar] [CrossRef]
  97. Seeharaj, P.; Kongmun, P.; Paiplod, P.; Prakobmit, S.; Sriwong, C.; Kim-Lohsoontorn, P.; Vittayakorn, N. Ultrasonically-assisted surface modified TiO2/rGO/CeO2 heterojunction photocatalysts for conversion of CO2 to methanol and ethanol. Ultrason. Sonochem. 2019, 58, 104657. [Google Scholar] [CrossRef]
  98. Tian, W.-J.; Chen, H.-J.; Zhao, L.; Zheng, L.-L.; Hu, Z.-N.; Yang, Y.-L. Preparation of adsorptive TiO2/SiO2 core-shell nanospheres by ultrasonic-coupled alkali leaching process: Synergy of adsorption and photocatalysis towards 2,4-dinitrophenol removal. Colloids Surf. A Physicochem. Eng. Asp. 2023, 679, 132530. [Google Scholar] [CrossRef]
  99. Chen, R.; Ding, S.; Wang, B.; Ren, X. Preparation of ZnFe2O4@TiO2 Novel Core-Shell Photocatalyst by Ultrasonic Method and Its Photocatalytic Degradation Activity. Coatings 2022, 12, 1407. [Google Scholar] [CrossRef]
  100. Mahammed Shaheer, A.R.; Thangavel, N.; Rajan, R.; Abraham, D.A.; Vinoth, R.; Sunaja Devi, K.R.; Shankar, M.V.; Neppolian, B. Sonochemical assisted impregnation of Bi2WO6 on TiO2 nanorod to form Z-scheme heterojunction for enhanced photocatalytic H2 production. Adv. Powder Technol. 2021, 32, 4734–4743. [Google Scholar] [CrossRef]
  101. Yang, H.; Li, W.; Yang, H.; Xiong, Y.; Liu, C.; Han, Y.; Yu, Z.-Z.; Li, X. One-Step In Situ Synthesis of a Reduced Graphene Oxide-Based Hybrid Hydrogel for Highly Efficient Water Evaporation and Comprehensive Wastewater Treatment. ACS Appl. Mater. Interfaces 2025, 17, 46046–46058. [Google Scholar] [CrossRef]
  102. Galedari, M.; Mehdipour Ghazi, M.; Mirmasoomi, S.R. Novel visible-driven Ag2O/Fe2O3/TiO2 nano sized hetero-structured photocatalyst: Synthesis, characterization and photo-degradation of tetracycline. Chem. Eng. Res. Des. 2021, 170, 248–255. [Google Scholar] [CrossRef]
  103. Jonda, E.; Sarman, M.; Bezdicka, P.; Kormunda, M.; Ecorchard, P.; Murafa, N.; Michalska, M. Surface modification of nano-TiO2 with Cu nanoparticles. Ceram. Int. 2025, in press. [Google Scholar] [CrossRef]
  104. Qi, Y.; Shen, Y.; Zhao, S.; Jiang, X.; Ma, R.; Cui, B.; Zhao, Q.; Wei, D. Degradation of multiple xanthates using highly efficient visible light-responsive BiOBr-TiO2 composite photocatalysts. J. Ind. Eng. Chem. 2023, 132, 461–473. [Google Scholar] [CrossRef]
  105. He, Z.; Zhang, M.; Yu, H.; Li, Z.; Deng, B.; Ren, H. Multifunctional BiPO4/TiO2/rGO composite for enhancement removal of malachite green and levofloxacin via adsorption and visible-light photocatalysis synergy. Mater. Sci. Semicond. Process. 2024, 169, 107891. [Google Scholar] [CrossRef]
  106. Abd-Rabboh, H.S.M.; Benaissa, M.; Hamdy, M.S.; Ahmed, M.A.; Glal, M. Synthesis of an efficient, and recyclable mesoporous BiVO4/TiO2 direct Z-scheme heterojunction by sonochemical route for photocatalytic hydrogen production and photodegradation of rhodamine B dye in the visible region. Opt. Mater. 2021, 114, 110761. [Google Scholar] [CrossRef]
  107. Bahal, M.; Kaur, N.; Sharotri, N.; Sud, D. Investigations on Amphoteric Chitosan/TiO2 Bionanocomposites for Application in Visible Light Induced Photocatalytic Degradation. Adv. Polym. Technol. 2019, 2019, 2345631. [Google Scholar] [CrossRef]
  108. Alhaddad, M.; Ismail, A.A.; Alghamdi, Y.G.; Al-Khathami, N.D.; Mohamed, R.M. Co3O4 Nanoparticles Accommodated Mesoporous TiO2 framework as an Excellent Photocatalyst with Enhanced Photocatalytic Properties. Opt. Mater. 2022, 131, 112643. [Google Scholar] [CrossRef]
  109. Kaviyarasan, K.; Vinoth, V.; Sivasankar, T.; Asiri, A.M.; Wu, J.J.; Anandan, S. Photocatalytic and photoelectrocatalytic performance of sonochemically synthesized Cu2O@TiO2 heterojunction nanocomposites. Ultrason. Sonochem. 2019, 51, 223–229. [Google Scholar] [CrossRef]
  110. Xu, T.; Wang, P.; Wang, D.; Zhao, K.; Wei, M.; Liu, X.; Liu, H.; Cao, J.; Chen, Y.; Fan, H.; et al. Ultrasound-assisted synthesis of hyper-dispersed type-II tubular Fe3O4@SiO2@ZnO/ZnS core/shell heterostructure for improved visible-light photocatalysis. J. Alloys Compd. 2020, 838, 155689. [Google Scholar] [CrossRef]
  111. Dobaradaran, S.; Nodehi, R.N.; Yaghmaeian, K.; Jaafari, J.; Niari, M.H.; Bharti, A.K.; Agarwal, S.; Gupta, V.K.; Azari, A.; Shariatifar, N. Catalytic decomposition of 2-chlorophenol using an ultrasonic-assisted Fe3O4–TiO2@MWCNT system: Influence factors, pathway and mechanism study. J. Colloid Interface Sci. 2018, 512, 172–189. [Google Scholar] [CrossRef]
  112. Wang, J.; Ren, P.; Du, Y.; Zhao, X.; Chen, Z.; Pei, L.; Jin, Y. Construction of tubular g-C3N4/TiO2 S-scheme photocatalyst for high-efficiency degradation of organic pollutants under visible light. J. Alloys Compd. 2023, 947, 169659. [Google Scholar] [CrossRef]
  113. Ni, S.; Fu, Z.; Li, L.; Ma, M.; Liu, Y. Step-scheme heterojunction g-C3N4/TiO2 for efficient photocatalytic degradation of tetracycline hydrochloride under UV light. Colloids Surf. A Physicochem. Eng. Asp. 2022, 649, 129475. [Google Scholar] [CrossRef]
  114. Kane, A.; Chafiq, L.; Dalhatou, S.; Bonnet, P.; Nasr, M.; Gaillard, N.; Dikdim, J.M.D.; Monier, G.; Assadi, A.A.; Zeghioud, H. g-C3N4/TiO2 S-scheme heterojunction photocatalyst with enhanced photocatalytic Carbamazepine degradation and mineralization. J. Photochem. Photobiol. A Chem. 2022, 430, 113971. [Google Scholar] [CrossRef]
  115. Alsalme, A.; Galal, A.H.; El-Sherbeny, E.F.; Soltan, A.; Abdel-Messih, M.F.; Ahmed, M.A. Fabrication of S-scheme TiO2/g-C3N4 nanocomposites for generation of hydrogen gas and removal of fluorescein dye. Diam. Relat. Mater. 2022, 122, 108819. [Google Scholar] [CrossRef]
  116. Ahmed, A.S.; Ahamad, T.; Ahmad, N.; Khan, M.Z. Removal enhancement of acid navy blue dye by GO-TiO2 nanocomposites synthesized using sonication method. Mater. Chem. Phys. 2019, 238, 121906. [Google Scholar] [CrossRef]
  117. 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] [PubMed]
  118. Liu, G.-y.; Li, K.-k.; Jia, J.; Zhang, Y.-t. Coal-based graphene as a promoter of TiO2 catalytic activity for the photocatalytic degradation of organic dyes. New Carbon Mater. 2022, 37, 1172–1180. [Google Scholar] [CrossRef]
  119. Lacerda Fernandes, Í.; Pereira Barbosa, D.; Botelho de Oliveira, S.; Antônio da Silva, V.; Henrique Sousa, M.; Montero-Muñoz, M.; Coaquira, J.A.H. Synthesis and characterization of the MNP@SiO2@TiO2 nanocomposite showing strong photocatalytic activity against methylene blue dye. Appl. Surf. Sci. 2022, 580, 152195. [Google Scholar] [CrossRef]
  120. AbdulKareem, E.A.; Mahmoud, Z.H.; Khadom, A.A. Sunlight assisted photocatalytic mineralization of organic pollutants over rGO impregnated TiO2 nanocomposite: Theoretical and experimental study. Case Stud. Chem. Environ. Eng. 2023, 8, 100446. [Google Scholar] [CrossRef]
  121. Deshmukh, S.P.; Kale, D.P.; Kar, S.; Shirsath, S.R.; Bhanvase, B.A.; Saharan, V.K.; Sonawane, S.H. Ultrasound assisted preparation of rGO/TiO2 nanocomposite for effective photocatalytic degradation of methylene blue under sunlight. Nano-Struct. Nano-Objects 2020, 21, 100407. [Google Scholar] [CrossRef]
  122. Balta, Z.; Simsek, E.B. Insights into the photocatalytic behavior of carbon-rich shungite-based WO3/TiO2 catalysts for enhanced dye and pharmaceutical degradation. New Carbon Mater. 2020, 35, 371–383. [Google Scholar] [CrossRef]
  123. Mohammad, A.; Khan, M.E.; Cho, M.H.; Yoon, T. Fabrication of binary SnO2/TiO2 nanocomposites under a sonication-assisted approach: Tuning of band-gap and water depollution applications under visible light irradiation. Ceram. Int. 2021, 47, 15073–15081. [Google Scholar] [CrossRef]
  124. Zhou, Z.; Li, J.; Lü, C.; Yin, C.; Xu, G.; Qin, X.; Wu, W. Dual S-scheme SnO2/TiO2(A)/TiO2(R) nanowires for visible-light-driven naphthalene complete degradation. J. Ind. Eng. Chem. 2024, 129, 474–487. [Google Scholar] [CrossRef]
  125. Chen, J.; Liu, L.; Hu, J.; Zhang, Q.; Gong, L.; Wei, X. Preparation of SnS2/TiO2 by a thermo-solvent ultrasonic method and its high photo-catalytic performance for decontamination under visible light. J. Environ. Chem. Eng. 2020, 8, 104121. [Google Scholar] [CrossRef]
  126. Sedaghati, N.; Habibi-Yangjeh, A.; Pirhashemi, M.; Asadzadeh-Khaneghah, S.; Ghosh, S. Integration of BiOI and Ag3PO4 nanoparticles onto oxygen vacancy rich-TiO2 for efficient visible-light photocatalytic decontaminations. J. Photochem. Photobiol. A Chem. 2020, 400, 112659. [Google Scholar] [CrossRef]
  127. Boudraa, R.; Talantikite-Touati, D.; Souici, A.; Djermoune, A.; Saidani, A.; Fendi, K.; Amrane, A.; Bollinger, J.-C.; Nguyen Tran, H.; Hadadi, A.; et al. Optical and photocatalytic properties of TiO2–Bi2O3–CuO supported on natural zeolite for removing Safranin-O dye from water and wastewater. J. Photochem. Photobiol. A Chem. 2023, 443, 114845. [Google Scholar] [CrossRef]
  128. Bathula, C.; Rabani, I.; Sekar, S.; Youi, H.-K.; Choy, J.-Y.; Kadam, A.; Shretha, N.K.; Seo, Y.-S.; Kim, H.-S. Enhanced removal of organic dye by activated carbon decorated TiO2 nanoparticles from Mentha Aquatica leaves via ultrasonic approach. Ceram. Int. 2021, 47, 8732–8739. [Google Scholar] [CrossRef]
  129. Akhter, P.; Ali, F.; Ali, A.; Hussain, M. TiO2 decorated CNTs nanocomposite for efficient photocatalytic degradation of methylene blue. Diam. Relat. Mater. 2024, 141, 110702. [Google Scholar] [CrossRef]
  130. Manivannan, R.; Ryu, J.; Son, Y.-A. Controlled ultrasonic synthesis of TiO2@C3N4 nanocomposites with porphyrin as a solid-state electron mediator: A promising material for pollutant discoloration under visible light. Ceram. Int. 2021, 47, 14399–14407. [Google Scholar] [CrossRef]
  131. He, X.; Meng, X.; Sun, J.; Yuan, Z.; He, Y.; Chen, S. Synthesis of TiO2@Fe2O3 Nanocomposites as effective Photocatalyst for degradation of p-nitophenol in oilfield wastewater. Int. J. Electrochem. Sci. 2022, 17, 221179. [Google Scholar] [CrossRef]
  132. Du, X.; Bai, X.; Xu, L.; Yang, L.; Jin, P. Visible-light activation of persulfate by TiO2/g-C3N4 photocatalyst toward efficient degradation of micropollutants. Chem. Eng. J. 2020, 384, 123245. [Google Scholar] [CrossRef]
  133. Tian, M.; Wang, J.; Sun, R.; Lu, D.; Li, N.; Liu, T.; Yao, M.; Zhang, G.; Li, L. Facile synthesis of rod-like TiO2-based composite loaded with g-C3N4 for efficient removal of high-chroma organic pollutants based on adsorption-photocatalysis mechanism. Inorg. Chem. Commun. 2022, 141, 109517. [Google Scholar] [CrossRef]
  134. Chi, N.; Xu, W. Synthesis of TiO2/g-C3N4 Hybrid Photocatalyst and its Application for Degradation of Chlorophenol as Organic Water Pollutant. Int. J. Electrochem. Sci. 2022, 17, 220929. [Google Scholar] [CrossRef]
  135. Toghan, A.; Abd El-Lateef, H.M.; Taha, K.K.; Modwi, A. Mesoporous TiO2@g-C3N4 composite: Construction, characterization, and boosting indigo carmine dye destruction. Diam. Relat. Mater. 2021, 118, 108491. [Google Scholar] [CrossRef]
  136. Izzataddini, A.; Romdoni, Y.; Helmiyati; Novi Marantika, R.; Amir, Z.; Kadja, G.T.M.; Jiwanti, P.K.; Khalil, M.; Mohamed Jan, B. Enhancement of visible light organic dyes photodegradation using TiO2 (001)/Graphene oxide nanocomposite. Inorg. Chem. Commun. 2023, 157, 111379. [Google Scholar] [CrossRef]
  137. Uma, K.; Singaravelu, C.M.; Kavinkumar, V.; Jothivenkatachalam, K.; Lin, J.-H. Ultrasonically modified P25-TiO2 /In2O3 heterostructured nanoparticles: An efficient dual- responsive photocatalyst for solution and gas phase reactions. J. Taiwan Inst. Chem. Eng. 2021, 125, 257–266. [Google Scholar] [CrossRef]
  138. Ardani, M.R.; Pang, A.L.; Pal, U.; Haniff, M.A.S.M.; Ismail, A.G.; Hamzah, A.A.; Khanday, W.A.; Ahmadipour, M. Ultrasonic-assisted of TiO2-MWCNT nanocomposite with advanced photocatalytic efficiency for elimination of dye pollutions. Diam. Relat. Mater. 2023, 137, 110066. [Google Scholar] [CrossRef]
  139. Karamifar, M.; Sabbaghi, S.; Mohtaram, M.S.; Rasouli, K.; Mohsenzadeh, M.; Kamyab, H.; Derakhshandeh, A.; Dolatshah, L.; Moradi, H.; Chelliapan, S. Ultrasonic-assisted synthesis of TiO2/MWCNT/Pani nanocomposite: Photocatalyst characterization and optimization of efficient variables in the degradation of benzene via RSM-CCD. Powder Technol. 2024, 432, 119176. [Google Scholar] [CrossRef]
  140. Wafi, M.A.E.; Ahmed, M.A.; Abdel-Samad, H.S.; Medien, H.A.A. Exceptional removal of methylene blue and p-aminophenol dye over novel TiO2/RGO nanocomposites by tandem adsorption-photocatalytic processes. Mater. Sci. Energy Technol. 2022, 5, 217–231. [Google Scholar] [CrossRef]
  141. Moghni, N.; Boutoumi, H.; Khalaf, H.; Makaoui, N.; Colón, G. Enhanced photocatalytic activity of TiO2/WO3 nanocomposite from sonochemical-microwave assisted synthesis for the photodegradation of ciprofloxacin and oxytetracycline antibiotics under UV and sunlight. J. Photochem. Photobiol. A Chem. 2022, 428, 113848. [Google Scholar] [CrossRef]
  142. Liu, Z.; Ren, Y.; Yu, Y.; Shen, K.; Hu, K.; Zhou, J.; Wu, M.; Liu, S.; Wang, Z. Effect of synthesis process steps on the structure and photocatalytic performance of ZnO–TiO2 composites. Ceram. Int. 2024, 50, 3168–3175. [Google Scholar] [CrossRef]
  143. Cheng, Y.; Dai, L.; Zhang, L.; Yu, B.; Yang, C.; Zhou, L.; Lou, B. Optimizing the dewatering performance of zinc smelting iron Slag: Investigating the influence of ultrasonic Time, ultrasonic Power, and Liquid-Solid ratio using response surface methodology. Ultrason. Sonochem. 2024, 103, 106797. [Google Scholar] [CrossRef]
  144. Abrinaei, F.; Aghabeygi, S. Optimization on preparation conditions to improve the nonlinear optical response of ZnO/TiO2/ZrO2 ternary nanocomposites under continuous-wave laser irradiation. Optik 2022, 255, 168720. [Google Scholar] [CrossRef]
  145. Kumar, S.; Sinhmar, P.S.; Gogate, P.R. Ultrasound assisted improved synthesis of TiO2 catalyst and subsequent evaluation for isomerization of alpha pinene. Chem. Eng. Process. Process Intensif. 2021, 169, 108591. [Google Scholar] [CrossRef]
  146. Tiple, A.; Sinhmar, P.S.; Gogate, P.R. Improved direct synthesis of TiO2 catalyst using sonication and its application for the desulfurization of thiophene. Ultrason. Sonochem. 2021, 73, 105547. [Google Scholar] [CrossRef] [PubMed]
  147. Estrada-Monje, A.; Zitzumbo-Guzmán, R.; Bañuelos-Díaz, J.A.; Zaragoza-Contreras, E.A. Ultrasonic dispersion and activation of TiO2 nanoparticles and its effect on bacterial inhibition in EVA films. Mater. Chem. Phys. 2019, 235, 121760. [Google Scholar] [CrossRef]
  148. Elavarasan, M.; Uma, K.; Yang, T.C.K. Photocatalytic oxidation of ethanol using ultrasonic modified TiO2; an in-situ diffuse reflectance infrared spectroscopy study. Results Phys. 2019, 13, 102237. [Google Scholar] [CrossRef]
  149. Bhargava, N.; Mor, R.S.; Kumar, K.; Sharanagat, V.S. Advances in application of ultrasound in food processing: A review. Ultrason. Sonochem. 2021, 70, 105293. [Google Scholar] [CrossRef]
  150. Chen, J.; Yao, Z.; Wang, F.; Chi, Y.; Wang, Z.; Tofil, S.; Yao, J. Grain refinement and Laves phase dispersion by high-intensity ultrasonic vibration in laser cladding of Inconel 718. J. Mater. Res. Technol. 2024, 30, 8563–8575. [Google Scholar] [CrossRef]
  151. Qayyum, A.; Giannakoudakis, D.A.; LaGrow, A.P.; Bondarchuk, O.; Łomot, D.; Colmenares, J.C. High-frequency sonication for the synthesis of nanocluster-decorated titania nanorods: Making a better photocatalyst for the selective oxidation of monoaromatic alcohol. Catal. Commun. 2022, 163, 106406. [Google Scholar] [CrossRef]
  152. Wood, R.J.; Lee, J.; Bussemaker, M.J. A parametric review of sonochemistry: Control and augmentation of sonochemical activity in aqueous solutions. Ultrason. Sonochem. 2017, 38, 351–370. [Google Scholar] [CrossRef] [PubMed]
  153. Harish, V.; Ansari, M.M.; Tewari, D.; Yadav, A.B.; Sharma, N.; Bawarig, S.; García-Betancourt, M.-L.; Karatutlu, A.; Bechelany, M.; Barhoum, A. Cutting-edge advances in tailoring size, shape, and functionality of nanoparticles and nanostructures: A review. J. Taiwan Inst. Chem. Eng. 2023, 149, 105010. [Google Scholar] [CrossRef]
  154. Bandeira, C.R.P.P.; Dória, A.R.; Cruz Ribeiro, J.Y.; Prado, L.R.; Anjos de Jesus, R.; Carvalho Andrade, H.M.; Souza de Santana Castro, R.; Romanholo Ferreira, L.F.; Egues, S.M.S.; Figueiredo, R.T. Tuning the sound frequency in the audible region during the synthesis of the precursor TiO2: Evaluation of the sound effect on the structure and photoactivity relationship. Mater. Chem. Phys. 2021, 265, 124521. [Google Scholar] [CrossRef]
  155. Oliveira, A.C.d.M.; de Jesus, R.A.; Bilal, M.; Iqbal, H.M.N.; Bharagava, R.N.; Yerga, R.M.N.; Ferreira, L.F.R.; Egues, S.M.; Figueiredo, R.T. Influence of sound and calcination temperature on the fabrication of TiO2-based photocatalysts and their photoactivity for H2 production. Mol. Catal. 2022, 529, 112523. [Google Scholar] [CrossRef]
  156. Dong, H.; Wang, P.; Yang, Z.; Li, R.; Xu, X.; Shen, J. Dual improvement in curcumin encapsulation efficiency and lyophilized complex dispersibility through ultrasound regulation of curcumin–protein assembly. Ultrason. Sonochem. 2022, 90, 106188. [Google Scholar] [CrossRef]
  157. Nguyen, C.H.; Tran, M.L.; Tran, T.T.V.; Juang, R.-S. Enhanced removal of various dyes from aqueous solutions by UV and simulated solar photocatalysis over TiO2/ZnO/rGO composites. Sep. Purif. Technol. 2020, 232, 115962. [Google Scholar] [CrossRef]
  158. Li, Z.; Li, L.; Zhang, S.; Zhang, L.; Cui, Y.; Shi, H. Ultrasound assisted TiO2@Fe3O4 nanocomposites photocatalytic degradation of organic pollutants in potato starch processing wastewater. Sep. Purif. Technol. 2024, 332, 125799. [Google Scholar] [CrossRef]
  159. Davies, K.R.; Cherif, Y.; Pazhani, G.P.; Anantharaj, S.; Azzi, H.; Terashima, C.; Fujishima, A.; Pitchaimuthu, S. The upsurge of photocatalysts in antibiotic micropollutants treatment: Materials design, recovery, toxicity and bioanalysis. J. Photochem. Photobiol. C Photochem. Rev. 2021, 48, 100437. [Google Scholar] [CrossRef]
  160. Hou, Y.; Pu, S.; Shi, Q.; Mandal, S.; Ma, H.; Xue, S.; Cai, G.; Bai, Y. Ultrasonic impregnation assisted in-situ photoreduction deposition synthesis of Ag/TiO2/rGO ternary composites with synergistic enhanced photocatalytic activity. J. Taiwan Inst. Chem. Eng. 2019, 104, 139–150. [Google Scholar] [CrossRef]
  161. Eskandari, P.; Amarloo, E.; Zangeneh, H.; Rezakazemi, M.; Zamani, M.R.; Aminabhavi, T.M. Photocatalytic activity of visible-light-driven L-Proline-TiO2/BiOBr nanostructured materials for dyes degradation: The role of generated reactive species. J. Environ. Manag. 2023, 326, 116691. [Google Scholar] [CrossRef]
  162. Kim, J.S.; Kim, Y.B.; Baek, S.K.; Yun, Y.D.; Jung, S.H.; Cho, S.W.; Ahn, C.H.; Cho, H.K. Compositionally graded SnO2/TiO2 bi-layered compounds with dramatically enhanced charge transport efficiency for self-driven water purification applications. J. Alloys Compd. 2019, 776, 839–849. [Google Scholar] [CrossRef]
  163. Magdalane, C.M.; Kanimozhi, K.; Arularasu, M.V.; Ramalingam, G.; Kaviyarasu, K. Self-cleaning mechanism of synthesized SnO2/TiO2 nanostructure for photocatalytic activity application for waste water treatment. Surf. Interfaces 2019, 17, 100346. [Google Scholar] [CrossRef]
  164. Yusuff, A.S.; Thompson-Yusuff, K.A.; Adeniyi, O.D.; Olutoye, M.A. Siliceous termite hill supported ZnO-TiO2 as a solar light responsive photocatalyst: Synthesis, characterization and performance in degradation of methylene blue dye. Surf. Interfaces 2022, 34, 102360. [Google Scholar] [CrossRef]
  165. Rajput, R.B.; Jamble, S.N.; Kale, R.B. A review on TiO2/SnO2 heterostructures as a photocatalyst for the degradation of dyes and organic pollutants. J. Environ. Manag. 2022, 307, 114533. [Google Scholar] [CrossRef]
  166. Narzary, S.; Alamelu, K.; Raja, V.; Jaffar Ali, B.M. Visible light active, magnetically retrievable Fe3O4@SiO2@g-C3N4/TiO2 nanocomposite as efficient photocatalyst for removal of dye pollutants. J. Environ. Chem. Eng. 2020, 8, 104373. [Google Scholar] [CrossRef]
  167. Huang, J.; Jing, H.-X.; Li, N.; Li, L.-X.; Jiao, W.-Z. Fabrication of magnetically recyclable SnO2-TiO2/CoFe2O4 hollow core-shell photocatalyst: Improving photocatalytic efficiency under visible light irradiation. J. Solid State Chem. 2019, 271, 103–109. [Google Scholar] [CrossRef]
  168. Duta, A.; Andronic, L.; Enesca, A. The influence of low irradiance and electrolytes on the mineralization efficiency of organic pollutants using the Vis-active photocatalytic tandem CuInS2/TiO2/SnO2. Catal. Today 2018, 300, 18–27. [Google Scholar] [CrossRef]
  169. Moradi, S.; Farhadian, M.; Nazar, A.R.S.; Moghadam, M. Application of Bi2WO6/N-TiO2 catalyst immobilized on FTO in a tray photoreactor for textile color degradation from aqueous solutions: Effects of mineral salts. J. Mol. Liq. 2023, 377, 121520. [Google Scholar] [CrossRef]
  170. Kumar, A.; Raorane, C.J.; Syed, A.; Bahkali, A.H.; Elgorban, A.M.; Raj, V.; Kim, S.C. Synthesis of TiO2, TiO2/PAni, TiO2/PAni/GO nanocomposites and photodegradation of anionic dyes Rose Bengal and thymol blue in visible light. Environ. Res. 2023, 216, 114741. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Overview of photocatalysis process of TiO2.
Figure 1. Overview of photocatalysis process of TiO2.
Catalysts 15 01010 g001
Figure 2. (a) MS plots of TiO2, Cu2O, WO3, and Cu2O/WO3/TiO2 composites; (b) MS plots of GO/TiO2 composite with the calculated N A ,   N d ,   a n d   e V b i readings. Reprinted from [53,56], with permission from Elsevier.
Figure 2. (a) MS plots of TiO2, Cu2O, WO3, and Cu2O/WO3/TiO2 composites; (b) MS plots of GO/TiO2 composite with the calculated N A ,   N d ,   a n d   e V b i readings. Reprinted from [53,56], with permission from Elsevier.
Catalysts 15 01010 g002
Figure 3. Different types of heterojunctions for composite photocatalyst: (a) Type II; (b) Z-scheme; (c) S-scheme.
Figure 3. Different types of heterojunctions for composite photocatalyst: (a) Type II; (b) Z-scheme; (c) S-scheme.
Catalysts 15 01010 g003
Figure 4. (a) EIS Nyquist plot and (b) Bode EIS plot of WO3, TiO2/CS-biochar, and WO3@TiO2/CS-biochar; The photoelectric response spectra of (c) pure TiO2 and (d) ZrO2/TiO2 composite; SEM image for TiO2 (e) ultrasonication and (f) without ultrasonication; (g) PL spectra and (h) EIS Nyquist plot of TiO2 and TiO2@MoS2/BiVO4 composite prepared with and without ultrasonication. Reprinted from [75,78,79], with permission from Elsevier.
Figure 4. (a) EIS Nyquist plot and (b) Bode EIS plot of WO3, TiO2/CS-biochar, and WO3@TiO2/CS-biochar; The photoelectric response spectra of (c) pure TiO2 and (d) ZrO2/TiO2 composite; SEM image for TiO2 (e) ultrasonication and (f) without ultrasonication; (g) PL spectra and (h) EIS Nyquist plot of TiO2 and TiO2@MoS2/BiVO4 composite prepared with and without ultrasonication. Reprinted from [75,78,79], with permission from Elsevier.
Catalysts 15 01010 g004
Figure 5. ESR spectra of (a) DMPO for O2 and (b) DMPO for OH; Temporal evolution plots of (c) DMPO for O2 and (d) DMPO for OH obtained from second peak intensity. Reprinted from [160], with permission from Elsevier.
Figure 5. ESR spectra of (a) DMPO for O2 and (b) DMPO for OH; Temporal evolution plots of (c) DMPO for O2 and (d) DMPO for OH obtained from second peak intensity. Reprinted from [160], with permission from Elsevier.
Catalysts 15 01010 g005
Table 1. TiO2-based composite photocatalysts prepared through ultrasonication or ultrasonic-assisted method.
Table 1. TiO2-based composite photocatalysts prepared through ultrasonication or ultrasonic-assisted method.
Photocatalyst CompositePreparation TechniqueOptimum
Ultrasonication Parameters
Band Gap (eV) of CompositeSurface Area
(m2/g)
Experimental ConditionsPhotocatalysis Performance
(%; min−1 for k)
Ref.
rGO/MXene-TiO2/polyanilineThermal assisted ultrasonicationTime: 30 min
Power: N/A
Freq.: N/A
2.99TiO2:
N/A
Composite:
N/A
Dosage: 0.20 g/L
[Methyl orange]: 10 mg/L
[Methylene blue]: 10 mg/L
[Rhodamine B]: 10 mg/L
[Norfloxacin]: 10 mg/L
[Tetracycline]: 10 mg/L
Time: 60 min
pH: N/A
Light: Solar
Volume: 50 mL
Methyl orange TiO2:
N/A
Composite:
98.40;
Methylene blue
TiO2:
N/A
Composite:
99.10;
Rhodamine B
TiO2:
N/A
Composite:
99.80;
Norfloxacin
TiO2:
N/A
Composite:
87.80;
Tetracycline
TiO2:
N/A
Composite:
95.10
[101]
Ag2O/TiO2/Fe2O3Impregnation assisted ultrasonicationTime: 45 min
Power: N/A
Freq.: 28 kHz
2.18TiO2:
N/A
Composite:
N/A
Dosage: 0.75 g/L
[Tetracycline]: 15 mg/L
Time: 90 min
pH: N/A
Light: Visible (60 W)
Volume: 300 mL
TiO2:
N/A
Composite:
90
[102]
WO3/TiO2Ultrasonic-assisted solvothermalTime: 30 min
Power: N/A
Freq.: N/A
2.80 to 3.10TiO2:
46.49
Composite:
41.73 to 48.20
Dosage: 0.25 g/L
[Reactive Black 5]: 30 mg/L
Time: 60 min
pH: 3
Light: Visible (150 W)
Volume: 80 mL
TiO2:
N/A
Composite:
92; k = 0.044
[28]
Cu2O/WO3/TiO2Ultrasonic-assisted hydrothermalTime: 60 min
Power: N/A
Freq.: N/A
2.70TiO2:
N/A
Composite:
N/A
Dosage: 0.25 g/L
[Acetaminophen]: 1 mg/L
Time: 60 min
pH: N/A
Light: Visible (150 W)
Volume: 80 mL
TiO2:
k = 0.024
Composite:
k = 0.044
[30]
Cu/TiO2Low temperature chemical process with ultrasonicationTime: 60 min
Power: 280 W
Freq.: N/A
3.11 to 3.23TiO2:
N/A
Composite:
N/A
Dosage: 0.30 g/L
[Acid orange 7]: N/A
[Rhodamine B]: N/A
Time: 120 min
pH: N/A
Light: UV; UV and Visible; Visible
Volume: N/A
Acid orange 7
98
Rhodamine B
99
[103]
Cu2O/WO3/TiO2Ultrasonic-assisted solvothermalTime: 60 min
Power: N/A
Freq.: N/A
2.35 to 2.90TiO2:
N/A
Composite:
35.77
Dosage: 0.13 g/L
[Reactive Black 5]: 30 mg/L
Time: 120 min
pH: 5
Light: Visible (150 W)
Volume: 80 mL
TiO2:
N/A
Composite:
k = 0.023
[53]
BiOBr/TiO2Ultrasonic-assisted hydrothermal and water bath precipitationTime: 5 min
Power: N/A
Freq.: N/A
N/ATiO2:
46.49
Composite:
42.13
Dosage: 0.20 g/L
[Sodium ethyl-xanthate]: 20 mg/L
Time: 40 min
pH: N/A
Light: Visible (400 W)
Volume: 50 mL
TiO2:
N/A
Composite:
93.96; k = 0.065
[104]
BiPO4/TiO2/rGOUltrasonic-assisted hydrothermalTime: N/A
Power: N/A
Freq.: N/A
1.78TiO2:
N/A
Composite:
N/A
Dosage: 0.67 g/L
[Malachite green]: 20 mg/L
[Levofloxacin]: 20 mg/L
Time: Until complete degradation of malachite green and levofloxacin
pH: N/A
Light: Visible (300 W)
Volume: 60 mL
Malachite green
TiO2:
N/A
Composite:
100 (45 min); k = 0.091
Levofloxacin
TiO2:
N/A
Composite:
94 (120 min); k = 0.024
[105]
BiVO4/TiO2Ultrasonic-assisted solvothermalTime: N/A
Power: N/A
Freq.: N/A
2.69TiO2:
N/A
Composite:
N/A
Dosage: 5 g/L
[hydrolyzed polyacrylamide]: 20 mg/L
Time: Until complete degradation of hydrolyzed polyacrylamide
pH: N/A
Light: Visible (35 W)
Volume: 200 mL
TiO2:
100 (115 min)
Composite:
100 (25 min)
[76]
BiVO4/TiO2UltrasonicationTime: 120 min
Power: N/A
Freq.: 80 kHz
2.72TiO2:
97.70
Composite:
59.30
Dosage: 1 g/L
[Rhodamine b]: N/A
Time: 120 min
pH: N/A
Light: Visible (300 W)
Volume: 100 mL
TiO2:
k = 0.002
Composite:
k = 0.021
[106]
chitosan/TiO2UltrasonicationTime: 30 min
Power: N/A
Freq.: 40 kHz
N/ATiO2:
N/A
Composite:
N/A
Dosage: 1 g/L
[Malachite green]: 10 mg/L
Time: 240 min
pH: 7
Light: Visible (15 W)
Volume: 100 mL
TiO2:
N/A
Composite:
91.94; k = 0.007
[107]
Co3O4/TiO2Ultrasonic-assisted impregnationTime: N/A
Power: N/A
Freq.: N/A
2.35TiO2:
200
Composite:
184
Dosage: 1 g/L
[Ciprofloxacin]: 10 mg/L
Time: 60 min
pH: N/A
Light: Visible (300 W)
Volume: 50 mL
TiO2:
k = 0.001
Composite:
100; k = 0.016
[108]
CuO/TiO2UltrasonicationTime: 15 min
Power: N/A
Freq.: N/A
2.81TiO2:
N/A
Composite:
N/A
Dosage: 2.40 g/L
[Acid yellow 36]: 30 mg/L
Time: Until complete degradation of acid yellow 36
pH: N/A
Light: UV-visible (350 W)
Volume: 250 mL
TiO2:
100 (100 min)
Composite:
100 (45 min)
[77]
Cu2O/TiO2UltrasonicationTime: 15 min
Power: 100 W
Freq.: 20 kHz
2.12TiO2:
N/A
Composite:
N/A
Dosage: N/A
[Methyl orange]: N/A
Time: 240 min
pH: N/A
Light: Visible
Volume: N/A
TiO2:
N/A
Composite:
k = 0.223
[109]
α-Fe2O3/TiO2Ultrasonication followed by wet impregnationTime: 8 min
Power: 70 W
Freq.: N/A
2.80TiO2:
N/A
Composite:
106.34
Dosage: 0.01 g/L
[Cefixime]: 20.50 mg/L
Time: 103 min
pH: 4.76
Light: Visible
Volume: 100 mL
TiO2:
N/A
Composite:
98.80; k = 0.057
[12]
Fe3O4/SiO2/ZnO/ZnSUltrasonic-assisted in situ surface sulfidationTime: 120 min
Power: N/A
Freq.: N/A
2.69TiO2:
N/A
Composite:
43.50
Dosage: 0.50 g/L
[Oxytetracycline]: 10 mg/L
[Tetracycline]: 10 mg/L
Time:
90 min (UV light);
180 min (visible light)
pH: N/A
Light:
UV (250 W);
visible (300 W)
Volume: 50 mL
Oxytetracycline (UV light)
TiO2:
N/A
Composite:
99.10
Tetracycline (visible light)
TiO2:
N/A
Composite:
80.90
[110]
Fe3O4/TiO2/MWCNTUltrasonic-assisted wet impregnationTime: 30 min
Power: 300 W
Freq.: 40 kHz
1.75TiO2:
28
Composite:
181.42
Dosage: 0.40 g/L
[2-chlorophenol]: 2 mg/L
Time: 15 min
pH: 5
Light: UV-visible (125 W)
Volume: 200 mL
TiO2:
N/A
Composite:
86
[111]
g-C3N4/TiO2Ultrasonic-assisted hydrothermalTime: N/A
Power: N/A
Freq.: N/A
2.88TiO2:
N/A
Composite:
N/A
Dosage: 0.40 g/L
[Methylene blue]: 20 mg/L
[Tetracycline]: 20 mg/L
[Congo red]: 20 mg/L
[Eosin y]: 20 mg/L
Time: Different reaction time
pH: N/A
Light: Visible
Volume: 50 mL
Methylene blue
TiO2:
68.90; k = 0.019
(60 min)
Composite:
96.60; k = 0.055
(60 min)
Tetracycline
TiO2:
73.60; k = 0.044
(30 min)
Composite:
100; k = 0.315
(30 min)
Congo red
TiO2:
N/A
Composite:
100 (60 min)
Eosin y
TiO2:
N/A
Composite:
87.70 (10 min)
[112]
g-C3N4/TiO2UltrasonicationTime: N/A
Power: N/A
Freq.: N/A
2.98TiO2:
50.19
Composite:
49.37
Dosage: 1 g/L
[Tetracycline]: 50 mg/L
Time: 50 min
pH: N/A
Light: UV (300 W)
Volume: 50 mL
TiO2:
N/A
Composite:
96.53; k = 0.057
[113]
g-C3N4/TiO2Ultrasonic-assisted wet impregnationTime: 30 min
Power: N/A
Freq.: N/A
2.99TiO2:
125.57
Composite:
80.65
Dosage: N/A
[Carbamazepine]: 10 mg/L
Time: 360 min
pH: N/A
Light: UV (24 W)
Volume: N/A
TiO2:
N/A
Composite:
71.41; k = 0.003
[114]
g-C3N4/TiO2UltrasonicationTime: 60 min
Power: 200 W
Freq.: N/A
2.79TiO2:
156
Composite:
43
Dosage: 1 g/L
[Fluorescein dye]: concentration of 5 × 10−5
Time: 120 min
pH: N/A
Light: Visible (350 W)
Volume: 100 mL
TiO2:
45; k = 0.002
Composite:
92; k = 0.025
[115]
GO/TiO2UltrasonicationTime: 60 min
Power: N/A
Freq.: N/A
3.56TiO2:
N/A
Composite:
N/A
Dosage: N/A
[Acid navy blue]: N/A
Time: 90 min
pH: N/A
Light: UV (125 W)
Volume: N/A
TiO2:
N/A
Composite:
95; k = 0.042
[116]
GO/TiO2Ultrasonic-assisted hydrothermalTime: 30 min
Power: 200 W
Freq.: 40 kHz
2.72TiO2:
N/A
Composite:
N/A
Dosage: 4 g/L
[Salicylic acid]: 1 mM
Time: 60 min
pH: N/A
Light: Sunlight
Volume: 250 mL
TiO2:
N/A
Composite:
57; k = 0.001
[117]
Graphene/TiO2Ultrasonic-assisted hydrothermalTime: 30 min
Power: N/A
Freq.: N/A
2.45TiO2:
29.61
Composite:
35.89
Dosage: N/A
[Rhodamine b]: 20 mg/L
[Methyl orange]: 30 mg/L
Time: 120 min
pH: N/A
Light: Visible (300 W)
Volume: 50 mL
Rhodamine b
TiO2:
k = 0.003
Composite:
k = 0.013
Methyl orange
TiO2:
k = 0.003
Composite:
k = 0.015
[118]
MNP/SiO2/TiO2 Ultrasonic-assisted solvothermal followed by sol–gelTime: N/A
Power: N/A
Freq.: N/A
N/ATiO2:
54
Composite:
167
Dosage: 0.56 g/L
[Methylene blue]: 3.50 mg/L
Time: 120 min
pH: 7
Light: UV (80 W)
Volume: 50 mL
TiO2:
k = 0.028
Composite:
k = 0.060
[119]
rGO/TiO2Photolysis followed by ultrasonication Time: 60 min
Power: N/A
Freq.: N/A
2.70TiO2:
N/A
Composite:
N/A
Dosage: 0.10 g/L
[Congo red]: 10 mg/L
[Trichloroacetic acid]: 10 mg/L
Time: 100 min
pH: N/A
Light: Direct sunlight
Volume: 100 mL
Congo red
TiO2:
k = 0.010
Composite:
k = 0.026
Trichloroacetic acid
TiO2:
k = 0.001
Composite:
k = 0.026
[120]
rGO/TiO2UltrasonicationTime: 30 min
Power: N/A
Freq.: N/A
N/ATiO2:
N/A
Composite:
N/A
Dosage: 2 g/L
[Methylene blue]: 20 mg/L
Time: 30 min
pH: 13.20
Light: Sunlight (between 11 am and 3 pm)
Volume: 100 mL
TiO2:
N/A
Composite:
91.30
[121]
rGO/ZnO/TiO2UltrasonicationTime: 60 min
Power: N/A
Freq.: N/A
N/ATiO2:
N/A
Composite:
N/A
Dosage: 0.10 g/L
[Crystal violet]: 50 mg/L
Time: 20 min
pH: 6.50
Light: UV
Volume: 100 mL
TiO2:
N/A
Composite:
87.06
[87]
rGO/ZnS/TiO2UltrasonicationTime: 30 min
Power: 240 W
Freq.: 22 kHz
Composite showed a red shift of absorbance compared to othersTiO2:
N/A
Composite:
N/A
Dosage: 0.40 g/L
[Crystal violet]: 50 mg/L
Time: 50 min
pH: N/A
Light: UV
Volume: 100 mL
TiO2:
N/A
Composite:
97
[86]
Shungite/WO3/TiO2Ultrasonic-solvothermal Time: 60 min
Power: N/A
Freq.: N/A
2.83TiO2:
N/A
Composite:
N/A
Dosage: 0.20 g/L
[Orange II]: 10 mg/L
Time: 120 min
pH: N/A
Light: UV
Volume: 50 mL
TiO2:
N/A
Composite:
93.40; k = 0.036
[122]
SnO2/TiO2Ultrasonic-assisted impregnationTime: N/A
Power: N/A
Freq.: N/A
2.92TiO2:
N/A
Composite:
N/A
Dosage: 0.20 g/L
[Tetracycline]: 10 mg/L
[Methylene blue]: N/A
[Congo red]: N/A
Time: 40 min (tetracycline); 135 min (methylene blue); 100 min (congo red)
pH: N/A
Light: Visible
Volume: 20 mL
Tetracycline
TiO2:
N/A
Composite:
98.76; k = 0.086
Methylene blue
TiO2:
N/A
Composite:
99
Congo red
TiO2:
N/A
Composite:
90
[123]
SnO2/TiO2(A)/TiO2(R)Ultrasonic-assisted alkaline hydrothermalTime: 150 min
Power: N/A
Freq.: N/A
2.72TiO2:
N/A
Composite:
78.50
Dosage: N/A
[Naphthalene]: 30 mg/L
Time: 220 min
pH: N/A
Light: Visible (300 W)
Volume: N/A
TiO2:
24.40; k = 0.001
Composite:
100; k = 0.038
[124]
SnS2/TiO2Thermosolvent ultrasonicationTime: 240 min
Power: N/A
Freq.: 15 kHz
Absorption wavelength of 647 nmTiO2:
N/A
Composite:
5.65
Dosage: 0.50 g/L
[Methyl orange]: 20 mg/L
[Methyl blue]: 5 mg/L
[Phenol]: 10 mg/L
Time: 60 min
pH: N/A
Light: Visible (300 W)
Volume: 400 mL
Methyl orange
TiO2:
N/A
Composite:
k = 0.009
Methyl blue
TiO2:
N/A
Composite:
k = 0.015
Phenol
TiO2:
N/A
Composite:
k = 0.003
[125]
TiO2/BiOI/Ag3PO4UltrasonicationTime: 10 min
Power: N/A
Freq.: N/A
N/ATiO2:
48.10
Composite:
94.20
Dosage: 0.40 g/L
[Rhodamine b]: ~5 mg/L
Time: 60 min
pH: N/A
Light: Visible (50 W)
Volume: 250 mL
TiO2:
N/A
Composite:
99.30; k = 0.061
[126]
TiO2/Bi2O3/CuO/zeoliteUltrasonic-assisted wet impregnationTime: 30 min
Power: N/A
Freq.: 50 kHz
1.73TiO2:
N/A
Composite:
16.85
Dosage: 0.40 g/L
[Safranin-O]: 16 mg/L
Time: 300 min
pH: N/A
Light: Sunlight
Volume: 200 mL
TiO2:
N/A
Composite:
94.10
[127]
TiO2/CarbonUltrasonicationTime: N/A
Power: N/A
Freq.: N/A
2.77TiO2:
27.07
Composite:
49.27
Dosage: 0.026 wt%
[Rhodamine b]: N/A
Time: 75 min
pH: N/A
Light: UV (400 W)
Volume: 100 mL
TiO2:
54
Composite:
94
[128]
TiO2/carbon nanotubesUltrasonic-assisted hydrothermalTime: N/A
Power: N/A
Freq.: N/A
N/ATiO2:
N/A
Composite:
N/A
Dosage: 2 g/L
[Methylene blue]: 30 mg/L
Time: 180 min
pH: 6.30
Light: Visible
Volume: 100 mL
TiO2:
N/A
Composite:
84; k = 0.030
[129]
TiO2/C3N4/TCPUltrasonicationTime: 30 min
Power: 100 W
Freq.: 20 kHz
Absorption wavelength between 400 and 700 nmTiO2:
N/A
Composite:
N/A
Dosage: 1 g/L
[Rhodamine b]: 12 mg/L
Time: 60 min
pH: N/A
Light: Visible (300 W)
Volume: 10 mL
TiO2:
51.70
Composite:
97.70
[130]
TiO2/Fe2O3Ultrasonic-assisted impregnationTime: 60 min
Power: N/A
Freq.: N/A
2.63TiO2:
N/A
Composite:
N/A
Dosage: 0.003 g/L
[p-nitrophenol]: 50 mg/L
Time: Until complete degradation of p-nitrophenol
pH: N/A
Light: Solar (300 W)
Volume: 300 mL
TiO2:
100% (70 min)
Composite:
100% (60 min)
[131]
TiO2/g-C3N4Ultrasonic-assisted calcinationTime: N/A
Power: 500 W
Freq.: 40 kHz
2.34TiO2:
N/A
Composite:
N/A
Dosage: 0.50 g/L
[Acetaminophen]: 5 mg/L
Time: 30 min
pH: N/A
Light: Visible (300 W)
Volume: 100 mL
TiO2:
N/A
Composite:
99.30
[132]
TiO2/g-C3N4Ultrasonic-assisted alkali hydrothermal Time: N/A
Power: N/A
Freq.: N/A
3.00TiO2:
N/A
Composite:
140.73
Dosage: 1 g/L
[Methylene blue]: 60 mg/L
Time: 240 min
pH: N/A
Light: Visible
Volume: 50 mL
TiO2:
N/A
Composite:
97
[133]
TiO2/g-C3N4Ultrasonic-assisted hydrothermal Time: N/A
Power: N/A
Freq.: N/A
2.88TiO2:
N/A
Composite:
N/A
Dosage: 0.20 g/L
[2-Chlorophenol]: 25 mg/L
Time: Until complete degradation of 2-chlorophenol
pH: N/A
Light: UV-visible (250 W)
Volume: 100 mL
TiO2:
100% (110 min)
Composite:
100% (65 min)
[134]
TiO2/g-C3N4UltrasonicationTime: 45 min
Power: N/A
Freq.: N/A
2.97TiO2:
111
Composite:
155
Dosage: 0.50 g/L
[Indigo carmine]: 50 mg/L
Time: 60 min
pH: N/A
Light: Visible
Volume: 100 mL
TiO2:
24; k = 0.004
Composite:
98; k = 0.030
[135]
TiO2/g-C3N4Ultrasonic-assisted hydrothermalTime: N/A
Power: N/A
Freq.: N/A
2.68TiO2:
N/A
Composite:
58.26
Dosage: 0.20 g/L
[Tetracycline]: 50 mg/L
Time: 120 min
pH: N/A
Light: Visible (300 W)
Volume: 100 mL
TiO2:
N/A
Composite:
k = 0.008
[74]
P90-TiO2/g-C3N4UltrasonicationTime: 480 min
Power: 150 W
Freq.: 37 kHz
2.53TiO2:
N/A
Composite:
N/A
Dosage:
0.50 g/L (sunset yellow FCF); 1 g/L (phenol)
[Sunset yellow FCF]: 10 mg/L
[Phenol]: 30 mg/L
Time:
5 min (sunset yellow FCF); 120 min (phenol)
pH: N/A
Light: Visible
Volume: 100 mL
Sunset yellow FCF
TiO2:
k = 0.089
Composite:
98.80, k = 0.837
Phenol
TiO2:
k = 0.004
Composite:
99.35; k = 0.036
[96]
TiO2/GOUltrasonicationTime: 60 min
Power: N/A
Freq.: N/A
Overlapping of the absorption edge of TiO2/GO with GOTiO2:
49
Composite:
92
Dosage: 0.10 g/L
[Congo red]: 5 mg/L
[Methyl orange]: 5 mg/L
Time: 30 min
pH: N/A
Light: UV
Volume: 100 mL
Congo red
TiO2:
N/A
Composite:
k = 0.065
Methyl orange
TiO2:
N/A
Composite:
k = 0.063
[93]
TiO2/GOUltrasonicationTime: N/A
Power: N/A
Freq.: N/A
3.13TiO2:
N/A
Composite:
N/A
Dosage: 0.50 g/L
[Methylene blue]: 0.01 mM
[Crystal violet]: 0.01 mM
Time: 10 min
pH: N/A
Light: Visible (12 W)
Volume: 50 mL
Methylene blue
TiO2:
k = 0.135
Composite:
95; k = 0.325
Crystal violet
TiO2:
k = 0.162
Composite:
99; k = 0.494
[136]
TiO2/In2O3Ultrasonic-assisted sol–gelTime: N/A
Power: 50 W
Freq.: N/A
2.92TiO2:
N/A
Composite:
N/A
Dosage: 0.25 g/L
[Acid blue]: 10 mg/L
Time: 90 min
pH: N/A
Light: UV-visible (350 W)
Volume: 100 mL
TiO2:
k = 0.030
Composite:
98; k = 0.040
[137]
TiO2/MoS2/BiVO4Ultrasonic-assisted hydrothermalTime: 120 min
Power: 750 W
Freq.: 20 kHz
2.84TiO2:
52.30
Composite:
132.70
Dosage: 1 g/L
[Tetracycline]: 20 mg/L
Time: 90 min
pH: N/A
Light: Visible (300 W)
Volume: 40 mL
TiO2:
52.30
Composite:
90.35
[79]
TiO2/MWCNTUltrasonic-assisted sol–gelTime: 60 min
Power: N/A
Freq.: 20 kHz
2.75TiO2:
76.34
Composite:
194.35
Dosage: 0.25 g/L
[Methylene blue]: 5 mg/L
[Rhodamine b]: 5 mg/L
Time: 60 min
pH: N/A
Light: UV
Volume: 200 mL
Methylene blue
TiO2:
62.63; k = 0.019
Composite:
92.36; k = 0.051
Rhodamine b
TiO2:
71.56; k = 0.022
Composite:
94.13; k = 0.053
[138]
TiO2/MWCNT/PaniUltrasonic-assisted in situ polymerizationTime: N/A
Power: N/A
Freq.: N/A
2.90TiO2:
120.05
Composite:
78.69
Dosage: 1.50 g/L
[Benzene]: 700 mg/L
Time: 80 min
pH: 6
Light: Visible (500 W)
Volume: 10 mL
TiO2:
N/A
Composite:
84.92
[139]
TiO2/RGOUltrasonic-assisted hydrothermalTime: 60 min
Power: N/A
Freq.: N/A
3.05TiO2:
41.20
Composite:
77.20
Dosage: 1 g/L
[Methylene blue]: 6.40 mg/L
Time: 60 min
pH: N/A
Light:
UV (6 W); sunlight (1pm and 3 pm in Cairo; 30.39w/m2)
Volume: 50 mL
UV light
TiO2:
28.01
Composite:
85.76
Sunlight
TiO2:
N/A
Composite:
81.00
[140]
TiO2/SiO2Ultrasonication coupled with alkali leachingTime: 120 min
Power: 160 W
Freq.: N/A
Absorption wavelength of 763 nmTiO2:
N/A
Composite:
N/A
Dosage: 0.20 g/L
[2,4-dinitrophenol]: 20 mg/L
Time: 120 min
pH: N/A
Light: Xenon (150 W)
Volume: 100 mL
TiO2:
N/A
Composite:
k = 0.080
[98]
TiO2/TPPSUltrasonicationTime: 30 min
Power: 100 W
Freq.: 20 kHz
2.87TiO2:
N/A
Composite:
N/A
Dosage: 0.10 g/L
[Eosin yellow]: 10 mg/L
Time: 50 min
pH: N/A
Light: Visible
Volume: 100 mL
TiO2:
k = 0.028
Composite:
99; k = 0.090
[24]
TiO2/WO3Ultrasonication and microwave Time: 60 min
Power: 750 W
Freq.: 20 kHz
2.73TiO2:
260
Composite:
235
Dosage: 0.50 g/L
[Ciprofloxacin]: 20 mg/L
[Oxytetracycline]: 20 mg/L
Time:
45 min (UV light); 120 min (sunlight) for ciprofloxacin
60 min (UV light); 120 min (daylight) for oxytetracycline
pH: N/A
Light: UVA (20 W); sunlight
Volume: 200 mL
Ciprofloxacin
TiO2:
100; k = 0.039 (UV light)
45; k = 0.005 (sunlight)
Composite:
100; k = 0.133 (UV light)
96; k = 0.034 (sunlight)
Oxytetracycline
TiO2:
91; k = 0.041 (UV light)
55; k = 0.007 (sunlight)
Composite:
100; k = 0.082 (UV light)
97; k = 0.028 (sunlight)
[141]
TiO2/WO3Ultrasonic-assisted sol–gelTime: 10 min
Power: 22 kHz
Freq.: 190 W
2.60TiO2:
N/A
Composite:
65.39
Dosage: 0.50 g/L
[Congo red]: 20 mg/L
[Methyl red]: 20 mg/L
Time: 120 min
pH: N/A
Light: Visible (250 W)
Volume: 100 mL
Congo red
TiO2:
N/A
Composite:
95
Methyl red
TiO2:
N/A
Composite:
100
[39]
TiO2/ZnIn2S4Ultrasonic-assisted hydrothermalTime: 60 min
Power: N/A
Freq.: N/A
Absorption wavelength of 436 nmTiO2:
N/A
Composite:
N/A
Dosage: N/A
[Rhodamine b]: N/A
Time: 120 min
pH: N/A
Light: Visible (300 W)
Volume: N/A
TiO2:
37.40; k = 0.004
Composite:
90.20; k = 0.033
[73]
TiO2/ZnOUltrasonic-assisted precipitationTime: 10 min
Power: N/A
Freq.: N/A
Around 2.30TiO2:
N/A
Composite:
N/A
Dosage: 4 g/L
[Methyl orange]: 30 mg/L
Time: N/A
pH: N/A
Light: UV (15 W)
Volume: 20 mL
TiO2:
36.20
Composite:
58.70
[142]
Zeolite/CdS/TiO2/CeO2Ultrasonic-assisted impregnationTime: N/A
Power: N/A
Freq.: N/A
2.10TiO2:
N/A
Composite:
N/A
Dosage: 1.50 g/L
[Methylene blue]: 10 mg/L
Time: 120 min
pH: 8
Light: Visible (85 W)
Volume: 100 mL
TiO2:
28
Composite:
99.90
[81]
ZnFe2O4/TiO2Solvothermal followed by ultrasonicationTime: 90 min
Power: N/A
Freq.: N/A
2.03TiO2:
12.73
Composite:
54.81
Dosage: 0.30 g/L
[Rhodamine b]: 10 mg/L
Time: 60 min
pH: N/A
Light: UV (500 W)
Volume: N/A
TiO2:
94.90; k = 0.045
Composite:
97.30; k = 0.062
[99]
ZnFe2O4/TiO2/Ag2OSolvothermal, followed by ultrasonication and chemical precipitationTime: 90 min
Power: N/A
Freq.: N/A
Absorption wavelength of 702 nmTiO2:
N/A
Composite:
120.90
Dosage: 0.30 g/L
[Rhodamine b]: 10 mg/L
Time: 40 min
pH: N/A
Light: UV (500 W)
Volume: 500 mL
TiO2:
N/A
Composite:
98.40; k = 0.094
[7]
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

Foong Chau, J.H.; Kong, E.D.H.; Chia, J.C.; Lai, C.W.; Juan, J.C.; Li, Y.; Xiang, P.; Badruddin, I.A.; Kumar, A. Ultrasonic-Assisted Fabrication of TiO2-Based Composite Photocatalysts for Enhanced Photocatalysis of Organic Pollutants: A Review. Catalysts 2025, 15, 1010. https://doi.org/10.3390/catal15111010

AMA Style

Foong Chau JH, Kong EDH, Chia JC, Lai CW, Juan JC, Li Y, Xiang P, Badruddin IA, Kumar A. Ultrasonic-Assisted Fabrication of TiO2-Based Composite Photocatalysts for Enhanced Photocatalysis of Organic Pollutants: A Review. Catalysts. 2025; 15(11):1010. https://doi.org/10.3390/catal15111010

Chicago/Turabian Style

Foong Chau, Jenny Hui, Ethan Dern Huang Kong, Jing Chang Chia, Chin Wei Lai, Joon Ching Juan, Yue Li, Ping Xiang, Irfan Anjum Badruddin, and Amit Kumar. 2025. "Ultrasonic-Assisted Fabrication of TiO2-Based Composite Photocatalysts for Enhanced Photocatalysis of Organic Pollutants: A Review" Catalysts 15, no. 11: 1010. https://doi.org/10.3390/catal15111010

APA Style

Foong Chau, J. H., Kong, E. D. H., Chia, J. C., Lai, C. W., Juan, J. C., Li, Y., Xiang, P., Badruddin, I. A., & Kumar, A. (2025). Ultrasonic-Assisted Fabrication of TiO2-Based Composite Photocatalysts for Enhanced Photocatalysis of Organic Pollutants: A Review. Catalysts, 15(11), 1010. https://doi.org/10.3390/catal15111010

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