Next Article in Journal
Enhanced Oxygen Reduction Reaction Activity of Carbon-Supported Pt-Co Catalysts Prepared by Electroless Deposition and Galvanic Replacement
Previous Article in Journal
Microbial Spore-Based Biocatalysts: Properties, Applications and New Trends
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Advances in Composite Photocatalysts for Efficient Degradation of Organic Pollutants: Strategies, Challenges, and Future Perspectives

1
Department of Chemical Engineering, Laval University, Quebec, QC G1V 0A6, Canada
2
Department of Chemistry, University of Agriculture Faisalabad, Faisalabad 38040, Pakistan
*
Authors to whom correspondence should be addressed.
Catalysts 2025, 15(9), 893; https://doi.org/10.3390/catal15090893
Submission received: 30 July 2025 / Revised: 8 September 2025 / Accepted: 13 September 2025 / Published: 17 September 2025

Abstract

The persistent release of synthetic dyes such as methylene blue (MB) into aquatic environments poses a significant ecological hazard due to their chemical stability and toxicity. In recent years, the application of engineered composite photocatalysts has emerged as a potent solution for efficient dye degradation under visible and UV light. This review comprehensively summarizes various advanced composites, including carbon-based, metal-doped, and heterojunction materials, tailored for MB degradation. Notably, composites such as TiO2/C-550, WS2/GO/Au, and MOF-derived α-Fe2O3/ZnO achieved near-complete degradation (>99%) within 30–150 min, while others, like ZnO/JSAC-COO and Ag/TiO2/CNT, displayed enhanced charge separation and stability over five consecutive cycles. Band gap engineering (ranging from 1.7 eV to 3.2 eV) and reactive oxygen species (·OH, ·O2) generation were key to their photocatalytic performance. This review compares the structural attributes, synthetic strategies, and degradation kinetics across systems, highlighting the synergistic role of co-catalysts, surface area, and electron mobility. This work offers systematic insight into the state-of-the-art composite photocatalysts and provides a comparative framework to guide future material design for wastewater treatment applications.

Graphical Abstract

1. Introduction

Recently, environmental contamination in the natural water cycle has drawn significant attention [1,2,3]. The production of extremely toxic and carcinogenic wastewater has surged due to the rapid growth of industrialization [4]. In our daily lives, water purification (WP) is crucial for reasons beyond just ensuring access to safe drinking water [5,6]. Wastewater from various industries, households, hospitals, and laboratories contains high concentrations of organic dyes, surfactants, heavy metals, and other harmful compounds [7,8,9]. These pollutants pose serious environmental problems as they are toxic to microorganisms [10], aquatic life [11], and human health [10], and they also degrade the soil quality [12,13]. Dyes are compounds that have a wide range of applications in many sectors. They are widely used in textile, food, cosmetics, leather, paper, plastics, processing, printing, rubber, pharmaceuticals, and tanneries [14]. Konstantinou and Albanis reported that approximately 1–20% of the colors produced worldwide are wasted during the dyeing process and end up in textile effluents [15,16]. As a result, dye degradation in industrial wastewater has gained significant attention, leading to the proposal of various remediation techniques [17,18].
Traditional methods, such as membrane filtration, chemical precipitation, biological treatments, coagulation/flocculation, and adsorption, etc., are commonly used for wastewater treatment. However, these approaches are not always practical or highly effective [19]. Forgacs and coworkers observed that the chemical stability of synthetic textile dyes renders traditional wastewater treatment technologies significantly inadequate in addressing wastewater contaminated with these contaminants [20]. Each of the above-described methods has several drawbacks when it comes to eliminating dyes from wastewater [21]. Since photocatalysis may mineralize the target contaminant, recent research [22,23,24,25,26] has focused on using this technique to degrade the dyes from wastewater. Photocatalytic technology has emerged as a highly effective and widely demonstrated method for decomposing organic dyes. This approach utilizes light, in combination with photocatalysts, to break down harmful organic compounds. Its key advantages include ease of operation, energy efficiency, environmental sustainability, and the prevention of secondary pollution [27]. Due to these attributes, photocatalysis stands out as an effective and sustainable method for wastewater treatment and environmental remediation. Table 1 provides a comparative overview of photocatalysis and other treatment techniques, emphasizing their respective limitations [28].
In recent years, numerous studies have been conducted to remove these pollutants using photocatalysts, exploring various composite materials for dye degradation [36]. Semiconductor composite photocatalysts have gained significant attention due to their enhanced properties, which arise from both individual components and the synergistic effects of composite formation. These materials have been extensively studied and utilized to tackle the urgent challenge of environmental pollution and mitigate the rapid deterioration of the living environment [37,38]. A range of wide and narrow band gap photocatalysts, including titanium [39,40,41] and zinc-based nanocomposites [42,43,44], carbon-based materials [45,46], MOFs [47,48], carbon nanotubes [49,50], and graphene [51,52], have been explored for the degradation of organic pollutants. However, single semiconductors often show low efficiency due to limited solar energy absorption and poor charge separation [53]. The present study aims to review the degradation of organic pollutants, with a specific focus on methylene blue (MB) dye. The primary objective is to assess the effectiveness of composite photocatalysts in enhancing the photocatalytic degradation of MB. It is a widely used heterocyclic aromatic compound, primarily applied in dyeing silk and cotton, and it serves as an indicator of chemical changes. However, its widespread use poses significant environmental concerns due to its persistence and potential toxicity [54].
This study also intends to examine how various composites affect the material’s morphological and structural characteristics, as well as how they affect photocatalytic performance. Therefore, we used a methodical approach to determine the efficiency of various composites in the degradation of organic pollutants to make this review comprehensive. This strategy incorporates the four main processes of planning, searching, screening, and reporting, as suggested by Tranfield and coworkers, as mentioned below [55,56].

2. Methodology

This review was conducted using a systematic framework designed to evaluate the photocatalytic degradation of methylene blue (MB), with particular attention paid to mechanistic insights and practical limitations. The approach followed four main stages, planning, searching, screening, and reporting, ensuring both reliability and comprehensiveness. In the planning stage, the scope of the review was clearly defined. The focus was placed on MB degradation under ultraviolet and visible light, the role of reactive oxygen species (ROS), self-sensitization effects, and the performance of different photocatalysts, including metal oxides, sulfides, and hybrid composites. The searching stage involved gathering relevant studies from reputable scientific databases such as Scopus, Web of Science, ScienceDirect, and Google Scholar. A combination of the keywords “methylene blue photocatalysis,” “visible-light degradation,” “catalyst stability,” “nanocomposites,” “organic pollutants,” and “advanced wastewater treatment” was used to capture a broad yet focused dataset. To ensure updated and credible information, priority was given to peer-reviewed articles, case studies, short communications, and review papers published within the last two decades.
During the screening stage, the collected literature was carefully evaluated for relevance. Duplicate records, studies outside the scope of MB degradation, or papers lacking experimental validation were excluded. Emphasis was placed on experimental investigations that offered insights into photocatalytic mechanisms, degradation efficiencies, and catalyst stability. Finally, in the reporting stage, the selected studies were systematically synthesized. The analysis compared the degradation efficiencies across different catalysts, discussed mechanistic pathways in detail, and highlighted challenges related to catalyst recovery and long-term stability. Furthermore, research gaps were identified, particularly in relation to energy efficiency, large-scale applicability, and sustainability. Following this structured methodology, approximately 200 articles were identified, examined, and consolidated. This comprehensive approach not only summarizes the current understanding of MB photocatalytic degradation but also outlines key directions for future research, thereby enhancing the reliability and significance of the findings.

3. Mechanism of Photocatalysis

The several stages of photocatalysis include the generation of electron–hole pairs through the absorption of solar light, the separation of these charge carriers, the migration of electrons and holes to the surface of the photocatalyst, and the redox processes that are aided by these charge carriers [57,58,59]. Significant electron–hole recombination takes place during electron migration, either in the bulk or on the surface of the photocatalyst. The energy released by this recombination is either heat or light. However, depending on the characteristics of the donor or acceptor species, long-lived electrons and holes on the photocatalyst surface start different redox processes [60]. The most typical photocatalysts are semiconductors, which possess a full electron valence band (VB) with well-defined band gaps and higher-energy free electron conduction bands (CBs) [61]. When the photocatalyst absorbs light energy from the light source, the electrons from the VB will be transferred to the CB to form the corresponding electron–hole (e–h+) pairs (Equation (1)). The superoxide radical anion (O2−) can be produced by the combination of the excited electrons on a photocatalyst’s surface and the dissolved oxygen (electron acceptor) (Equation (2)). Equations (3) and (4) state that free hydroxyl radicals can be produced when H2O and OH molecules react with photogenerated holes on the photocatalyst surface [62]. Additionally, organic contaminants can be directly oxidized by the free pores on photocatalyst surfaces. The strong hydroxyl radicals 2.80 V oxidation potential allows them to effectively break down organic pollutants that have adsorbed on the photocatalyst’s surface, producing H2O and CO2 (Equation (5)) [62,63]. Figure 1 shows a schematic diagram of the photocatalysis mechanism for the photocatalytic degradation of organic pollutants.
P h o t o c a t a l y s t + h γ   ( e C B + h V B + )
e C B + O 2 O 2
h V B + + O H OH
h V B +   H 2 O     H + + OH
( OH ,   O 2 ,   H 2 O 2 ,   h V B + ) + Organic   Pollutant   CO 2 + Water + other   products
Reactive oxygen species (ROS) are classified as primary intermediates during photocatalytic activities. Therefore, in order to understand the mechanism of photodegradation, it is crucial to quantify, identify, and evaluate their kinetics [64]. Recently, there has been significant interest in developing novel photocatalysts with high activity under visible light to enhance photocatalytic efficiency [65,66]. Although they are compelling oxidants, reactive oxygen species produced during photocatalysis, such as hydroxyl radicals (OH), superoxide anions (O2), and singlet oxygen (1O2), do not exhibit selectivity. They efficiently break down target pollutants, but they can also react with background components, such as carbonates, dissolved organic matter, and other non-target substances, altering the water’s chemistry and potentially creating secondary byproducts [1,2]. Degradation efficiency may be decreased, and unexpected environmental effects may become a concern. Therefore, the surface characteristics of the photocatalyst, the reactor conditions, and the complexity of the water matrix all have a significant impact on the selectivity of ROS-driven degradation [3].
Methylene blue’s self-sensitization mechanism entails MB’s interaction with photocatalysts, which accelerates the degradation process via a number of pathways. The primary driving force behind this mechanism is MB’s capacity to absorb light and transfer electrons or energy to the catalyst, thereby promoting its own breakdown. Environmental factors like the pH and light irradiation, the kind of catalyst being used, and the existence of extra sensitizers all affect the process [4,5]. Energy transfer from the excited MB to triplet oxygen creates singlet oxygen, which oxidizes MB, as part of the self-sensitization mechanism of methylene blue degradation. Under visible light, this pathway is important, especially when the photocatalyst is not excited [6]. Porphyrin sensitizers have also been utilized to elucidate the self-sensitization mechanism of methylene blue (MB) degradation, wherein photocatalytic activity is enhanced and the degradation process is facilitated under visible light irradiation through electron transfer from the photoexcited H2TCPP or CuTCPP molecules to the conduction band of TiO2 [7]. Through demethylation during photocatalytic processes, the chromophoric and auxochrome groups of methylene blue (MB) break down, forming a variety of intermediates, including azure A, B, and C, as well as thionin [8].

4. Carbon-Based Composites for Photocatalytic Degradation

Since carbon-based photocatalysts are inexpensive, have good mechanical and chemical durability, and have high electrical conductivity, they have a bright future in photocatalytic applications. Carbon nanotubes (CNTs), graphene, graphite, and porous carbon are examples of carbon-based nanomaterials that have been extensively employed in photocatalysis for dye degradation in recent years [67,68]. Syed et al. found that the incorporation of carbon quantum dots (CQDs) and graphite oxide (GO) into polyacrylonitrile (PAN) nanofibers greatly improves their mechanical stability and photocatalytic activity. PAN/GO, PAN/CQDs, and PAN/GO/CQDs nanofibers were synthesized and evaluated for visible-light-driven photocatalytic degradation of MB. Through electrospinning and chemical cross-linking, the GO/PAN/CQD composite showed 100% MB degradation in 25 min [45]. A new Z-scheme CdS/CQDs/g-C3N4 heterojunction composite was synthesized by Feng et al. through a simple calcination process, showing enhanced photocatalytic activity for MB degradation. The optimized composite exhibited the highest degradation efficiency of 98% after irradiation for 120 min, which greatly exceeded that of pure g-C3N4 (65%), CdS (75%), CdS/g-C3N4 (80%), and CQDs/g-C3N4 (86%). This outstanding performance was due to the effective charge separation and redox capability facilitated by the Z-scheme heterojunction [69].
MWCNTs/TiO2 nanocomposites were prepared by Jiang et al. using the liquid phase deposition method and tested as photocatalysts for MB degradation under UV/visible light irradiation. This composite exhibited higher photocatalytic efficiency than pure TiO2, particularly in neutral and alkaline media, reaching up to 90%. The improved degradation was due to the high surface area of MWCNTs, which enhanced MB adsorption and charge transfer [70]. The addition of H2O2 greatly enhanced the rate of degradation by serving as an electron acceptor, producing the hydroxyl radicals necessary for the breakdown of pollutants. At pH 7.0, over 80% of MB degraded in the presence of H2O2, while just 20% degraded without it. The rate of degradation was the greatest at pH 10.3 (τ1/2 = 7.0 min) and reduced in acidic conditions, where only 42.5% degradation was noted under UV light after 60 min. This phenomenon was attributed to pH-dependent TiO2 surface charge changes affecting MB adsorption. The results demonstrate the promising potential of MWCNTs/TiO2 composites for effective wastewater treatment [71]. In another case, TiO2-functionalized carbon nanotube (TiO2/CNT) composites and Ag-deposited ternary multilayered composites (Ag/TiO2/CNT) were prepared by Ye et al. via chemical deposition and the sonochemical method, respectively, as shown in Figure 2. The incorporation of CNTs and Ag greatly improved the photocatalytic degradation activity of TiO2. The degradation activity of TiO2/CNT composites was found to be about 96%, whereas pure TiO2 was about 80%, and that of Ag/TiO2/CNT composites increased degradation further up to about 98% after 20 min. The improved photocatalysis was due to the synergistic impact of CNTs and Ag within the TiO2 matrix [72]. The high surface area of CNTs was available for dye adsorption, and it stored electrons, which enhanced charge separation [73,74]. The band gap of anatase TiO2 (~3.2 eV) and the work function gaps among TiO2, Ag (~4.24 eV), and CNTs (~4.7 eV) resulted in the formation of Schottky barriers, allowing efficient electron transfer. These results indicate the promising potential of Ag/TiO2/CNT composites for high-performance photocatalytic uses in wastewater treatment and environmental purification [72].
Carbonization and in situ precipitation were employed to create the lignin-based carbon/cadmium sulfide (LCS) composite, resulting in a stable, highly porous substance. The LCS photocatalyst effectively eliminated methylene blue with 91.7% efficiency in two hours when exposed to visible light, with rapid adsorption reaching saturation within approximately 30 min. CdS nanoparticles are uniformly anchored on the carbon scaffold derived from lignin, as seen in SEM micrographs. This promotes dispersion and exposes a large number of active sites, improving adsorption and reaction kinetics (Figure 3). After five consecutive cycles, the catalyst maintained an activity of greater than 80% and demonstrated good resistance to both acidic and alkaline media, indicating its durability. The lignin–carbon domain is believed to enhance performance by serving as an electron sink, facilitating charge separation, and inhibiting electron–hole recombination, thereby strengthening visible-light photocatalysis [75].
The Yb-TiO2/g-C3N5 (YTCN) heterostructured photocatalyst was prepared by Bai et al. through hydrothermal synthesis and calcination and evaluated for the photocatalytic degradation of MB under a 500 W xenon lamp. The optimized YTCN (1:3) composite achieved a 96.57% degradation rate of MB in 105 min, with a decay rate constant of 0.02083 min−1, which was 2.2 times higher than that of pure TiO2. Adding Yb to the TiO2 matrix enhanced charge carrier separation, increasing the MB degradation efficiency by 7.25%, whereas further inclusion of g-C3N5 enhanced it by another 51.86%. The better photocatalytic activity of YTCN is due to heterojunction formation, which prolongs the lifetime of charge carriers and enhances light absorption. Nevertheless, too high a TiO2 ratio may hinder active sites, while a lower proportion of TiO2 will impair the cooperative effect, leading to decreased redox efficiency. The YTCN (1:3) composite exhibited the highest MB degradation performance, affirming the effectiveness of rare earth doping and heterostructure construction in enhancing photocatalytic activity (Figure 4a) [76].
Several characterization methods have been used to confirm the production of a metal-free porphyrin enclosing (2,4-dinitrophenyl)-imino)-methyne chromophore (PcDNPIMC). Because of its suitable band gap and 18πe-aromatic structure, the porphyrin-based photocatalyst demonstrated exceptional optical and photocatalytic characteristics. After 240 min of exposure to 5 W LED light, the ideal conditions for the photodegradation of MB were found to be 20 mg/L of photocatalyst in a 10 ppm solution at pH 5, which led to a 95.5% degradation of the dye in a handmade photoreactor (Figure 4b) [77].
The TiO2/C composites prepared from sawdust were tested for MB removal under simulated sunlight. TiO2/C-550 demonstrated the best photocatalytic efficiency, degrading almost 100% of MB in 30 min. This enhanced activity is linked to the optimized light absorption (band gap ~2.7 eV), minimal charge recombination, and higher adsorption capacity (9.7 mg/g). Superoxide radicals (O2) played a dominant role in the degradation mechanism, as confirmed by trapping studies. Other variants, TiO2/C-600, C-650, and C-700, showed lower efficiencies (71.3%, 83.7%, and 56.8%, respectively), likely due to structural collapse and surface area loss at elevated calcination temperatures. Among all these, TiO2/C-550 also exhibited the highest adsorption and rate constant based on kinetic modeling. Although the material retained ~95% efficiency after five cycles, its durability in real wastewater scenarios needs further exploration due to potential challenges like carbon oxidation and catalyst fouling (Figure 5) [78].
The photocatalytic degradation of MB using TiO2−x-based nanocomposites was investigated, with RGO-TiO2−x exhibiting the highest performance, characterized by a rate constant of 0.075 min−1. This enhanced photocatalytic activity is attributed to the synergistic effect between the oxygen vacancies in TiO2−x, which promote efficient charge carrier separation, and the excellent electrical conductivity of reduced graphene oxide (RGO), which suppresses electron–hole recombination. Compared to pristine TiO2, the RGO-TiO2−x composite achieved nearly double the degradation efficiency, underscoring the critical role of interfacial charge transfer dynamics and improved visible light absorption. The integration of RGO not only facilitates rapid electron transport but also extends light harvesting, thereby significantly boosting the overall photocatalytic performance [79].
Figure 6a illustrates the proposed photodegradation mechanism of TiO2−x and RGO-TiO2−x composites. Under light irradiation, electrons are excited from the valence band (VB) to the conduction band (CB) of TiO2−x. The presence of oxygen vacancies in TiO2−x facilitates the capture of these photoexcited electrons, thereby reducing recombination. Furthermore, due to the favorable alignment of the conduction band energy levels and work functions, efficient electron transfer occurs from TiO2−x to reduced graphene oxide (RGO), significantly enhancing charge separation and transport. Figure 6b demonstrates that the high surface area of RGO serves as an adsorptive platform, concentrating MB molecules near the catalyst surface. This promotes surface diffusion of the pollutants toward TiO2−x, where they interact with oxygen vacancies. These adsorbed species are subsequently oxidized by photogenerated holes or reactive oxygen species (ROS), leading to the formation of intermediate compounds. These intermediates are further mineralized into benign end-products such as CO2, H2O, SO42−, and NO3. The RGO-TiO2−x composite exhibits markedly superior degradation efficiency compared to both pristine TiO2 and TiO2−x alone. This improvement is primarily attributed to its enhanced visible light response, greater charge mobility, and effective suppression of electron–hole recombination [79].
ZnO@CBC demonstrates highly effective removal of MB through a synergistic mechanism involving both adsorption and photocatalytic degradation. The synthesis route of ZnO@CBC is illustrated in Figure 7a. The adsorption process is primarily governed by electrostatic interactions and π–π stacking between MB molecules and the composite surface, both of which are pH-dependent. Fourier-transform infrared (FTIR) spectroscopy confirmed the successful adsorption of MB, as evidenced by changes in characteristic functional group peaks. Scavenger experiments revealed that hydroxyl radicals (·OH) play a dominant role in the photocatalytic degradation pathway. Upon light irradiation, ZnO generates electron–hole (e/h+) pairs. The photoexcited electrons are transferred to the conductive biochar (CBC) component, where they reduce molecular oxygen to form superoxide radicals (·O2), while the photogenerated holes in ZnO oxidize water or hydroxyl groups to produce ·OH radicals. These reactive oxygen species synergistically break down MB molecules into non-toxic end-products. Figure 7b illustrates that the integration of CBC not only facilitates efficient charge separation but also enhances surface adsorption and electron transport, significantly boosting the overall photocatalytic performance of the ZnO@CBC composite [80].
The AC@Fe3O4 composite, with an average crystallite size of 16.73 nm, exhibits excellent photocatalytic performance in the degradation of MB, as illustrated in Figure 8. At an optimal catalyst dosage of 0.50 g/L, AC@Fe3O4 achieves a high photodegradation efficiency of 94.6% for 10 mg/L MB within 140 min, corresponding to a kinetic rate constant of 0.025 min−1. Reusability tests further confirm the material’s stability, maintaining a degradation efficiency of 84% after four successive cycles. These results highlight the robustness and durability of AC@Fe3O4 as a promising photocatalyst for practical wastewater treatment applications [81].
The photodegradation rate of MB dye increased significantly upon forming a ZnO/g-C3N4 composite. The rate constant improved from 0.016 min−1 for pristine ZnO and 0.011 min−1 for g-C3N4 to 0.022 min−1 for the ZnO/g-C3N4 (10 wt%) composite. Notably, further enhancement was observed with the ZnO/g-C3N4 (50 wt%) composite, which achieved a maximum MB removal efficiency of 97%, outperforming the 10 wt% composite. In comparison, the individual components exhibited lower removal efficiencies, with ZnO achieving 90% and g-C3N4 only 73%. These results clearly demonstrate the synergistic effect of the composite structure, which enhances charge separation and light absorption, leading to superior photocatalytic performance [82]. The carboxylate-functionalized activated carbon (JSAC-COO) was obtained by filtration, thorough washing with deionized water, and drying at 60 °C for 24 h. The carboxylation process is depicted in Figure 9. Among the tested materials, the ZnO/JSAC-COO composite exhibited the highest photocatalytic efficiency for methylene blue (MB) degradation, achieving 97.56% removal. This performance significantly surpassed that of pure ZnO (49%) and ZnO-RGO composites (87.40%). The degradation of MB followed pseudo-first-order kinetics, with the ZnO/JSAC-COO composite showing a rate constant of 0.0124 min−1, approximately twofold and fivefold higher than those of ZnO and ZnO-RGO, respectively. This enhancement in both efficiency and reaction rate is attributed to the electron-accepting ability of the carboxylate-functionalized JSAC, which facilitates effective charge separation by capturing photogenerated electrons from ZnO’s conduction band, thereby suppressing electron–hole recombination. Moreover, the composite’s narrow band gap and high surface area further contribute to its superior photocatalytic activity [83].
Under visible light irradiation, the electrons in g-C3N4 are photoexcited and transferred to MnO2 and ZnO, where they participate in the reduction of molecular oxygen (O2) to generate superoxide radicals (O2•−). However, since the conduction band (CB) potential of MnO2 is less negative than the O2/O2•− redox potential, only ZnO and g-C3N4 effectively contribute to the formation of O2•−. The generated superoxide further reacts to form hydrogen peroxide (H2O2), which subsequently decomposes into hydroxyl radicals (OH), potent oxidizing agents responsible for dye degradation. Simultaneously, the photogenerated holes (h+) from MnO2 and ZnO migrate toward g-C3N4, while the h+ in g-C3N4 directly oxidize MB. This directional charge transfer across the heterojunction components significantly suppresses electron–hole recombination, thereby enhancing the overall photocatalytic degradation efficiency of MB (see Figure 10) [84].
The co-doping of sulfur and boron heteroatoms into g-C3N4, forming BSCN, and its integration with Ag2S significantly enhances the visible-light-driven photocatalytic degradation of MB. The Ag2S/BSCN composite exhibits markedly higher photocatalytic performance than individual components, pure g-C3N4, BSCN, or Ag2S, owing to its improved charge carrier separation and reduced electron–hole recombination. Among the prepared composites, the sample containing 5 wt% Ag2S demonstrates the highest photocatalytic activity, outperforming the 1%, 3%, and 10% variants. This optimized composite achieves a remarkable MB degradation efficiency of 98.5% within 120 min under visible light irradiation [85]. Two-dimensional (2D) contour plots and three-dimensional (3D) response surface plots were employed to analyze the influence of key operational parameters on the degradation efficiency of MB. As shown in Figure 11A, the interaction between pH and H2O2 dosage reveals that MB degradation initially increases with the addition of H2O2, reaching an optimum level, beyond which excessive H2O2 leads to a decline in efficiency, likely due to the scavenging effect of hydroxyl radicals. Figure 11B illustrates that lower initial MB concentrations, combined with an optimal H2O2 dose, significantly enhance degradation performance, while higher dye concentrations suppress efficiency due to the limited availability of reactive species. In Figure 11C, it is observed that the MB removal efficiency improves with increasing pH up to a critical point under a fixed H2O2 dose, after which no significant enhancement is observed [86].
Spinel ferrite nanoparticles and their carbon-based composites are widely explored for wastewater treatment due to their stability, adsorption capacity, magnetic behavior, and appropriate band gap energy. These materials are effective in removing both organic and inorganic pollutants through adsorption and photocatalytic degradation. Their adsorption efficiency depends on factors such as the surface area, surface charge, annealing temperature, functional groups, and cation distribution. Owing to their versatility and tunable properties, spinel ferrites hold significant promise for tackling emerging contaminants and addressing novel environmental challenges [46].
The photocatalytic mechanism of CA/ZnS-Ag under visible light irradiation is illustrated in Figure 12. In this system, carbon aerogel (CA) functions as a photosensitizer, facilitating the formation of electron–hole (e-h+) pairs indirectly through interaction with a sacrificial agent, since direct excitation of CA is not feasible. Electrons generated in CA are transferred to the conduction band (CB) of ZnS or reduce the molecular oxygen (O2) to produce superoxide radicals (O2). Meanwhile, the Ag nanoparticles enhance photocatalytic activity via surface plasmon resonance (SPR), injecting additional electrons into the ZnS CB. Additionally, CA can accept electrons from ZnS CB, resulting in the generation of holes. These photogenerated charge carriers migrate to the catalyst surface, where adsorbed water molecules and dye pollutants participate in producing hydroxyl radicals and other reactive species, ultimately leading to effective dye degradation [87].
The photocatalytic mechanism of S-gTAHP is elucidated based on the band structures of TAP and S-gAP, as depicted in Figure 13. Upon exposure to sunlight, both materials generate electron–hole (e–h+) pairs. However, the traditional type II heterojunction mechanism does not apply here because the conduction band (CB) potential of TAP (−0.252 V) is insufficient to reduce O2 to superoxide radicals (O2), and the valence band (VB) potential of S-gAP (1.769 V) is inadequate for oxidizing H2O to hydroxyl radicals (OH). Instead, the system operates via an S-scheme heterojunction, wherein electrons transfer from S-gAP to TAP until their Fermi levels align, resulting in the formation of an internal electric field at the interface. This electric field facilitates the recombination of less active charge carrier electrons in the TAP CB and holes in the S-gAP VB while preserving the more reactive holes in the TAP VB and electrons in the S-gAP CB. These retained charge carriers are responsible for the enhanced photocatalytic activity observed, which aligns with the results of the radical trapping experiments shown in Figure 13 [88].
The AC-ZrO2/CeO2 nanocomposite showed exceptional efficacy in photocatalytic degradation and adsorption, especially when it came to eliminating organic contaminants from wastewater, such as the cationic dye MB. The maximum adsorption capacity of 75.54 mg/g was reached in 60 min, and the adsorption behavior closely matched the Langmuir isotherm model. Additionally, reactive species such as superoxide radicals (O2), photogenerated holes (h+), and hydroxyl radicals (OH) drove the nanocomposite’s improved photocatalytic destruction of MB. The rate constant for MB degradation was found to be 0.05999 min−1, and the degradation of MB followed the Langmuir–Hinshelwood kinetic model [89]. The Cu-g-C3N4/BC (600)/H2O2 system primarily degrades MB through hydroxyl radicals (·OH), as confirmed by quenching experiments and electron paramagnetic resonance (EPR) analysis. The significant decrease in degradation efficiency upon addition of tert-butanol underscores the dominant role of ·OH, while superoxide radicals (·O2) and photogenerated holes (h+) contribute to a lesser extent. EPR spectra verified the in situ generation of ·OH, ·O2, and h+ species under visible light irradiation. The bamboo charcoal (BC) matrix enhances adsorption of both MB and H2O2, improves visible light absorption, and facilitates electron conductivity. Photogenerated electrons originating from Cu2O, CuO, and g-C3N4 were efficiently transferred to BC, suppressing electron–hole recombination and promoting the formation of the reactive radicals ·OH and ·O2. Additionally, the Cu2+/Cu+ redox cycling further accelerates ·OH production through Fenton-like reactions. The overall high photocatalytic efficiency of this composite arises from the synergistic interactions between its components and the stable anchoring of Cu sites on the BC framework (Figure 14b) [90].
Figure 15 illustrates the excellent cyclic stability of the WS2/GO/Au electrocatalyst during the dye removal process, demonstrating high reproducibility over extended use. Catalyst recyclability, alongside photo- and electrochemical stability, is a key factor when evaluating catalytic performance. Notably, the WS2/GO/Au electrocatalyst maintains its activity over an impressive 3000 electrocatalytic cycles in MB degradation. Even at elevated MB concentrations, the catalyst sustains stability and achieves a high conversion efficiency of up to 99%. For a 50 µL MB solution, complete degradation requires approximately 3000 cycles, corresponding to a total duration of 72 min. Continuous electrolysis results in a steady decline in MB absorption, confirming effective dye degradation. Furthermore, complete mineralization of MB is attained within one hour using WS2/GO/Au, whereas WS2/GO and GO alone only achieve 30% and 10% degradation, respectively, within the same timeframe [91].
Incorporation of 10% NiO into g-C3N4 led to a 66% reduction in the photoluminescence peak intensity, indicating suppressed electron–hole recombination. Diffuse reflectance spectroscopy confirmed the enhanced visible light absorption by the NiO/g-C3N4 nanocomposite. Consequently, the photocatalytic degradation efficiency of MB improved significantly, increasing from 33% with pristine g-C3N4 to 91.6% after 90 min at pH 8.0. Scavenger studies revealed that the Z-scheme photocatalytic mechanism played a crucial role in facilitating effective charge separation under visible light irradiation (Figure 16). Among reactive species, superoxide ions (O2•−) were identified as the primary active species responsible for MB degradation. Furthermore, the NiO/g-C3N4 composite exhibited excellent stability, retaining 95% of its degradation efficiency after five consecutive cycles [92].
Sono-photocatalysis combines sunlight, ultrasound, and a catalyst to synergistically enhance the degradation of dyes and pollutants by promoting the generation of reactive radicals. For GO/Fe3O4 nanocomposites, ultrasonic cavitation produces localized high temperatures (~500 K) and pressures (~500 bar), leading to water pyrolysis and the formation of highly reactive OH and H radicals. Additionally, the collapse of cavitation bubbles emits light that excites electrons within the catalyst, generating electron–hole pairs. The catalyst surface provides active sites that facilitate radical formation, while simultaneous sunlight irradiation induces further charge carrier excitation. Electrons are preferentially absorbed by GO, which improves reduction reactions and ultimately boosts dye degradation efficiency (Figure 17) [93].
The Ni/MoS2/MOF-2@g-C3N4 heterostructure demonstrated outstanding photocatalytic performance, achieving up to 91% degradation of MB. This enhanced activity is attributed to the efficient generation of reactive species, specifically superoxide (O2•−) and hydroxyl (OH) radicals. The degradation process followed pseudo-first-order kinetics, completing effectively within 90 min under light irradiation. Moreover, the heterostructure exhibited excellent stability, maintaining high photocatalytic efficiency with only slight performance loss over five consecutive cycles (Figure 18) [94].
When compared to their undoped counterparts, nitrogen-doped multi-walled carbon nanotubes (N-CNTs) demonstrated better adsorption capacity and photocatalytic activity. Stronger π–π electron donor–acceptor interactions between the negatively charged MB molecules and the sp2 lattice of the carbon nanostructure surface are made possible by nitrogen doping, which introduces highly reactive spots and improves performance. A schematic illustration of the surface of the carbonaceous nanostructure is shown in Figure 19a. The following describes the adsorption and photocatalytic mechanisms: the adsorbed MB molecules are efficiently decomposed by the interaction of hydroxyl radicals (OH), superoxide radical anions (O2•−), and hydroperoxyl radicals (HO2) (Figure 19b). In particular, nitrogen doping improves the photocatalytic effectiveness in comparison to graphene nanoribbons (GNRs) and multi-walled carbon nanotubes (MWCNTs) by raising the chemical reactivity of the carbonaceous nanomaterial (CNx), which in turn improves the reaction kinetics during MB photodegradation [95].
Eggshell-based activated carbon was successfully synthesized using a simple chemical activation method with orthophosphoric acid and sodium hydroxide. The carbon showed increasing adsorption efficiency during MB degradation, with reactive species like free radicals and superoxide ions playing a key role in the decolorization process. Activated carbon prepared with orthophosphoric acid outperformed that prepared with sodium hydroxide in terms of photocatalytic activity [96]. Raising the calcination temperature from 650 °C to 800 °C increased the graphitic carbon and rutile TiO2 content but decreased the surface area and TiO2 crystal size. At 800 °C, anatase TiO2 was fully converted to rutile in the TiO2/C composite. This TiO2/C-800 composite followed pseudo-second-order kinetics, indicating chemisorption, with a maximum adsorption capacity of ~15 mg/g. Under UV light (254 nm, 15 W) for 3 h, its degradation efficiency exceeded the adsorption-only removal by 25.1% [97]. Wood-derived carbon (WDC) enhanced MB adsorption and interaction with g-C3N4 (Figure 20). Its porous structure improved light absorption and reduced charge recombination. Electrons transferred from g-C3N4 to WDC formed superoxide radicals (·O2), while valence band holes generated hydroxyl radicals (·OH) or directly oxidized MB. Both radicals contributed to effective dye degradation [98].
By improving charge separation and encouraging the production of reactive radicals, the N-CDs@ZnO composite greatly accelerates the breakdown of MB. Under UV light, nitrogen-doped carbon dots (N-CDs) efficiently transport electrons, preventing photogenerated electron–hole pairs from recombining and producing superoxide radicals (O2). Furthermore, ZnO is activated by the up-conversion luminescence of N-CDs, and dye adsorption is improved by π–π* interactions between the composite and MB molecules. When compared to bare ZnO, the N-CDs@ZnO composite exhibits higher photocatalytic activity due to these synergistic effects (Figure 21) [99].
Figure 22 presents a schematic representation of the enhanced photocatalytic process facilitated by the TiO2@CFs composite. The mechanism involves three primary steps: (a) the initial adsorption of MB molecules onto the carbon fibers (CFs); (b) the absorption of light energy by the TiO2@CFs composite; and (c) the subsequent degradation of MB molecules through photocatalytic reactions enabled by efficient charge transfer. In this system, CFs serve a dual function, providing structural support for TiO2 nanoparticles and accelerating the adsorption of MB molecules by promoting their aggregation on the catalyst surface [100]. Table 2 represents the summary of carbon-based composite photocatalysts used for the degradation of MB.

5. Metal-Based Composites

Metal-based photocatalysts are generally classified into metal halides, metal nitrides, and metal oxides. This section provides a comprehensive review of their applications in photocatalysis, focusing on metal-based composite photocatalysts. Key examples and their corresponding photocatalytic performances are summarized in Table 3. For instance, when MB was treated with 50% CoVO/WxOy under visible light irradiation, a degradation efficiency of 96% was achieved within 80 min, with an observed rate constant of 0.30154 min−1 [109]. CoVO/WxOy nanocomposites are fabricated using a template-free hydrothermal method (Figure 23). Similarly, the effect of different phases in Cu2O/CuO/Cu composites on the degradation of MB under visible light irradiation was investigated. The Cu2O-rich composite exhibited superior photocatalytic performance, achieving a maximum photocurrent density of 13 mA/cm2 along with the highest optimized MB degradation efficiency. In contrast, the Cu2O-poor composite demonstrated significantly lower activity, reaching only 6.1 mA/cm2 [110].
The inorganic halide perovskite Cs3Bi2I9 has garnered significant interest due to its rapid charge carrier mobility, high tolerance to defects, low defect density, excellent water stability, and low toxicity. It exhibits a photoluminescence quantum yield of 2.3%. When exposed to visible light for 50 min, the Cs3Bi2I9/ZIF-8 composite achieved an impressive 98.2% degradation of methylene blue (MB), which is approximately 4.15 times greater than that of pure Cs3Bi2I9. Moreover, the composite retained its high photocatalytic efficiency even after three consecutive cycles, demonstrating strong stability and reusability [111]. The average particle sizes ranged from 15.2 nm for CuCO3 to 87.7 nm for the CuCO3-ZnCO3 composite. Specific surface areas varied from 7.8 m2/g (Ag2CO3) to 55.5 m2/g (CuCO3). Under UV irradiation, carbonate-based nanoparticles effectively degraded MB, with Ag2CO3-ZnCO3 achieving the highest efficiency of 99.88%. Notably, these nanoparticles retained significant photocatalytic activity even when embedded in a polymer matrix; for instance, Ag2CO3 showed a degradation efficiency of 91.67% in the composite [112]. ZnCO3, an n-type semiconductor with a wide band gap of about 3.56 eV, requires UV light for activation. Its conduction and valence band positions are located at 0.43 eV and 3.65 eV, respectively. Doping with silver nanoparticles narrows the band gap, thereby enhancing visible light absorption. In contrast, Ag2CO3 is a p-type semiconductor with a narrower band gap between 2.1 and 2.8 eV; its conduction and valence bands lie at 0.38 eV and 2.66 eV, respectively. In ZnCO3-Ag2CO3 composite systems, electrons transfer from ZnCO3 to Ag2CO3, while Ag nanoparticles serve as electron sinks, promoting charge separation and minimizing recombination. This synergistic effect markedly improves photocatalytic performance. Additionally, ZnCO3 aids in electron transport and stabilizes Ag+ ions, further enhancing the activity and longevity of the composite (Figure 24).
The calcined CuO-WO3 nanocomposite (0.1 M) exhibited superior photocatalytic activity, achieving 70% degradation of MB after 180 min of exposure to sunlight. In comparison, the uncalcined and undoped WO3 nanoparticles demonstrated significantly lower performance, with only 40% degradation under identical conditions. The influence of the photocatalyst dosage was also assessed using 10 mg and 20 mg of the calcined CuO-WO3 nanocomposite. A dosage of 20 mg yielded a maximum degradation efficiency of 92%, whereas 10 mg resulted in 78% degradation after 3 h of sunlight irradiation. The improved performance at higher catalyst loading is attributed to the greater number of accessible active sites, which facilitate enhanced generation of charge carriers and reactive species responsible for MB degradation [113]. Additionally, the optimized PEDOT/Ag2SeO3 nanohybrid with a 3:1:1–4 mL ratio exhibited notable photocatalytic activity, achieving 46.2% degradation of MB under 210 min of incandescent lamp irradiation, as determined at the dye’s absorption maximum of 663 nm. The study also evaluated the effect of pH on photocatalytic efficiency, with the results indicating that maximum degradation occurred under optimal pH conditions. This emphasizes the critical role of pH in modulating photocatalytic performance [114].
Figure 25a presents a proposed mechanism for the photocatalytic degradation of MB using the PEDOT/Ag2SeO3 (3:1:1–4 mL) nanohybrid. Upon illumination, PEDOT, due to its relatively narrow band gap (~1.30 eV), efficiently absorbs light and generates a significant number of electron–hole (e–h+) pairs. In contrast, Ag2SeO3, with its limited light-harvesting capability, produces fewer charge carriers. However, given that the conduction band of PEDOT lies at a lower energy level than that of Ag2SeO3, photogenerated electrons and holes from Ag2SeO3 migrate into the PEDOT matrix. This interfacial charge transfer increases the charge carrier density within PEDOT, thereby enhancing its overall photocatalytic efficiency. In comparison, the reduced graphene oxide (RGO)-supported Zn0.5Cu0.5Fe2O4 nanocatalyst (denoted as MRGO) demonstrated relatively lower efficiency in degrading MB under UV irradiation. Nevertheless, under visible light in aqueous solution, MRGO exhibited excellent activity, achieving 95.2% degradation of a 10 ppm MB solution within 40 min at pH 9 using 20 mg of catalyst. However, its photocatalytic activity decreased significantly after five successive degradation cycles, with the efficiency dropping to 54.6%, indicating limited long-term reusability [115].
Photodegradation experiments using the RGO/Zn0.5Cu0.5Fe2O4 (RGO/ZCF) nanocomposite were conducted under UV light. As shown in Figure 25b, the UV reactor consisted of a 100 mL cylindrical glass chamber (27 cm × 3 cm), lined internally with aluminum foil to enhance light reflection. During the process, 50 mL of MB dye solution was added to the reactor. Figure 25c illustrates the interaction mechanism between the MRGO 20 nanocomposite and MB molecules under visible light. Zn0.5Cu0.5Fe2O4 nanoparticles absorb incident photons, generating e–h+ pairs. The excited electrons are rapidly transferred to the reduced graphene oxide (RGO) layer due to its excellent electrical conductivity, which suppresses recombination and facilitates effective charge separation. The high surface area of RGO promotes strong adsorption of MB molecules onto the catalyst surface. Simultaneously, the photogenerated holes oxidize MB, while the electrons reduce dissolved O2 to generate reactive oxygen species (ROS), including hydroxyl radicals (OH) and superoxide anions (O2), which actively participate in the breakdown of MB into less harmful degradation products [115].
FeSe2 and ZnO exhibited average crystallite sizes of 21 nm and 33 nm, respectively, and individually degraded only 58% and 60% of MB after 300 min of irradiation. In contrast, the FZ-3 composite (75 wt% FeSe2 and 25 wt% ZnO) achieved 97% degradation. The band gaps of FeSe2 (1.3 eV) and ZnO (3.3 eV) were tuned in the composite, enhancing visible light absorption. The improved performance is attributed to a type-II band alignment, where electrons transfer from FeSe2 to ZnO and holes from ZnO to FeSe2, promoting effective charge separation. This facilitates the generation of reactive oxygen species (H2O2 and OH), which synergistically degrade MB into CO2 and H2O, as shown in Figure 26a [116]. The TiO2/Fe3O4 nanocomposite exhibited high visible-light-driven photocatalytic efficiency, degrading MB by 85% under tungsten halogen light and up to 92% under sunlight within 120 min. Its performance declined by only 13.7% after five consecutive cycles, confirming the excellent stability and reusability [117]. As shown in Figure 26b, TiO2 generates electron–hole (e–h+) pairs under light irradiation, while Fe3O4 enhances charge separation and modifies surface properties, improving photocatalytic activity. Scavenger tests using ammonium oxalate (h+), 2-propanol (OH), and benzoquinone (O2•−) were conducted to identify the reactive species involved in the degradation mechanism [117].
As shown in Figure 27, MB solutions treated with alginate/PCL-TiO2 hybrid bio-composites underwent a visible color change from blue to transparent after 24 h of irradiation, indicating effective dye degradation [118]. Additionally, the 4% PVA/TiO2 nanocomposite demonstrated strong antibacterial activity against Staphylococcus aureus, Enterococcus faecalis, Proteus mirabilis, and Pseudomonas aeruginosa, along with high MB degradation efficiency [119]. Thermal treatment of the MB sample at 800 °C for 2 h yielded a material with excellent structural properties, achieving 99.53% degradation of a 100 mL, 20 mg/L MB solution within 60 min. Further analysis highlighted the essential role of oxygen-containing radicals and H2O2 in the Fenton-assisted photodegradation mechanism [120].
The TSFNVMo composite dispersion exhibited a high zeta potential of −35.9 mV, indicating strong colloidal stability and adsorption capability. Despite a reduction in the specific surface area to 76 m2/g due to vanadium and molybdenum incorporation, the TSFNVMo photocatalyst showed excellent performance under UV–visible light, with an adsorption capacity of 10 × 10−3 μmol/g. Under sunlight, the capacity reached 7.2 × 10−3 μmol/g. The band gap narrowed from 3.12 eV to 2.52 eV, enhancing visible light absorption and photocatalytic activity [121]. Under UV radiation, the TiO2-C@N photocatalyst achieved 99.87% MB removal, compared to 28.9% in dark conditions. Kinetic analysis revealed a pseudo-first-order degradation model. As shown in Figure 28a,b, 97.8% degradation was achieved within 30 min at 100 mg/L MB concentration, though efficiency declined to 79% at 500 mg/L. The catalyst retained activity over five adsorption and six degradation cycles (Figure 28c), with only minor losses due to acid stripping with 1 M HNO3, confirming the excellent reusability and stability [122]. In comparison, kaolinite/TiO2 nanocomposites showed enhanced MB degradation under UV light. The band gap increased from 2.93 eV to 3.14 eV with higher kaolin content, leading to improved degradation efficiency from 71% with pure TiO2 to 98% using the composite at 1 g/L dosage and 20 mg/L MB concentration [123].
The diatomite-TiO2 composite, featuring a porous structure and a combination of anatase and rutile phases, achieved 80% MB degradation under natural sunlight within 270 min [124]. In comparison, TiO2/C-550 composites exhibited nearly complete degradation (~100%) of MB in just 30 min. This composite had a direct band gap of 2.7 eV, enabling enhanced visible light absorption and reduced charge recombination. It also retained 95% efficiency after five reuse cycles, confirming the excellent stability [78]. TiO2/C composites displayed mesoporous structures with high BET surface areas (117–138 m2/g) and small TiO2 crystallites (8–27 nm). Higher calcination temperatures (650–800 °C) led to improved adsorption and photocatalytic performance. The TiO2/C-800 composite showed a maximum MB adsorption capacity of ~15 mg/g and followed a pseudo-first-order kinetic model (k = 4.8 × 10−4 min−1). Under 254 nm UV light (15 W) for 3 h, it exhibited 25.1% higher MB removal than by adsorption alone [97]. The 2D/0D g-C3N4/TiO2 nanocomposite outperformed both individual components, achieving 100% MB degradation under visible light within 2 h (Figure 29a). This was attributed to its high surface area (273.32 m2/g), interfacial charge transfer that suppressed electron–hole recombination (Figure 29b), and its layered nanospherical morphology offering abundant active sites (Figure 29c) [125]. Figure 30a illustrates the enhanced MB degradation kinetics of B-doped g-C3N4/TiO2, which followed a pseudo-first-order model. The incorporation of boron into the composite improved the photocatalytic efficiency compared to undoped materials, with performance increasing as the co-catalyst proportion rose (Figure 30b) [126].
The TiO2-carbonized medium-density fiberboard (TiO2-cMDF), synthesized via carbonization of MDF treated with 50% (v/v) titanium tetraisopropoxide (Ti-tip) in isopropyl alcohol, was evaluated for its adsorption and photodegradation of MB under UV-C (254 nm) light. After complete MB adsorption, four consecutive photodegradation cycles were performed, achieving 99% MB removal over 396 h. The degradation followed pseudo-first-order kinetics, with the rate constant decreasing from 11.0 × 10−3/h in the first cycle to 5.9 × 10−3/h [127]. The Cr-doped TiO2/C (0.2%) nanocomposite exhibited enhanced photocatalytic performance, achieving 90% MB degradation in 70 min under natural sunlight, compared to 63% for pure TiO2. The degradation followed pseudo-first-order kinetics, with a regression coefficient (R2) of 0.93, indicating a consistent and efficient mechanism (Figure 31a) [128]. TiO2/CuxO composite films were fabricated by sequentially depositing a TiO2 layer via the doctor blade method, followed by spin coating of a CuxO layer. Among the samples, the TC 3.6 composite (containing 3.6 wt% CuxO) showed the highest photocatalytic activity in degrading MB (Figure 31b) [129].
The photocatalytic degradation of MB and glyphosate was investigated using barium defect-modified graphitic carbon nitride TiO2 at concentrations ranging from 10 to 200 mg L−1 over 105 min. As the catalyst concentration increased, the MB degradation efficiency improved from 56% to 97%. The photocatalytic process initiates when photon energy exceeds the band gap, exciting electrons from the valence band (VB) to the conduction band (CB), leaving behind holes in the VB. These electron–hole pairs generate hydroxyl radicals (OH), which actively degrade persistent organic pollutants into CO2 and H2O. The proposed degradation mechanism is shown in Figure 32 [130]. Doping TiO2 with copper reduced its band gap from 3.0 eV to 2.67 eV. Pure g-C3N4 exhibited a band gap of 2.81 eV, while Cu-TiO2/g-C3N4 composites showed band gaps ranging from 2.82 to 2.88 eV depending on the composition. Among them, Cu-TiO2/50% g-C3N4 demonstrated the highest photocatalytic rate constant (4.4 × 10−3 min−1), which is approximately 5 and 9.8 times greater than that of TiO2 and g-C3N4, respectively. Its adsorption rate constant was calculated as 0.122 g mg−1 min−1. Figure 33 illustrates three mechanisms: (a) a homojunction between the anatase and rutile phases in P25 TiO2, where electrons transfer from anatase to rutile and OH/O2 species primarily form in anatase; (b) a type-I heterojunction in Cu-doped TiO2, where both electrons and holes migrate toward Cu-TiO2, leading to recombination and limited activity; and (c) a Z-scheme in the Cu-TiO2/g-C3N4 system, where charge separation is enhanced, allowing efficient generation of OH and O2 radicals that drive MB degradation [131].
The photocatalytic degradation of MB using CoO/ZnO nanocomposites was optimized at a 1:2 ratio, achieving 67.5% degradation within 3 h, with measurements taken every 30 min. Kinetic analysis using the Langmuir–Hinshelwood model yielded an apparent rate constant (Kapp) of 6.428 × 10−3 min−1 and a half-life (t1/2) of 107.84 ± 0.01 min at the optimal concentration. The broad absorption edge of CoO/ZnO nanoparticles suggests strong absorption in the UV-Vis range (190–350 nm). As illustrated in Figure 33b, UV excitation generates e–h+ pairs in both CoO and ZnO; the photogenerated electrons reduce O2 to form superoxide radicals (O2), while the holes oxidize H2O or OH to generate hydroxyl radicals (OH), both of which contribute to MB degradation [132]. Higher degradation efficiencies were observed with ZnO-based composites: ZnO/activated carbon (ZnO/AC) achieved 85.7%, while magnetically separable ZnO/MNC composites reached 97.14%. The enhanced performance of ZnO/MNC is attributed to the incorporation of Fe3O4, which narrows the band gap, improves photon absorption, and increases the surface area, thereby boosting photocatalytic activity [133].
The Cu/ZnO catalyst demonstrated superior photocatalytic activity for MB degradation, achieving 99% removal at 45 °C within 15 to 75 min. This enhanced performance is attributed to the uniform dispersion of CuO within ZnO and the higher density of surface hydroxyl groups. The degradation rate increased with the temperature (25–45 °C), and the activation energies (Ea) for 5Cu/ZnO were calculated as 18.00 and 12.15 kJ/mol for MB degradation [134]. Similarly, CuO-ZnO nanocomposites with 20 wt% CuO showed significantly improved activity, with a rate constant of 0.017 min−1 for MB degradation, compared to 0.0027 min−1 for pure ZnO nanoparticles. Additionally, a reduction rate constant of 5.925 min−1g−1 was reported for 4-nitrophenol (4-NP) using the same nanocomposites [135]. In another study, g-C3N4/ZnO composites containing 75% ZnO achieved 98% MB degradation within 180 min under UV light, far exceeding the efficiency of individual ZnO or g-C3N4. The proposed charge transfer mechanism responsible for this enhanced photocatalytic activity is illustrated in Figure 34 [136].
Optical analysis revealed that the Fe3O4/ZnO nanocomposite (MZ4) exhibited enhanced visible light absorption and prolonged electron–hole pair lifetimes at the optimal composition, contributing to the improved photocatalytic degradation of MB. MZ4 achieved a maximum degradation efficiency of 88.5% under visible light and maintained excellent stability over five reuse cycles. The optimal pH for MB degradation using MZ4 was determined to be six [137]. Additionally, Ag doping in ZnO was found to reduce the band gap, with the lowest observed value being 3.16 eV. The photocatalytic performance under simulated daylight showed that pure ZnO achieved 93.6% MB degradation within 120 min, while all the Ag/ZnO composites surpassed this, with the highest degradation rate reaching 99.8%. The proposed photocatalytic mechanism for Ag/ZnO systems is illustrated in Figure 35 [138].
The enhanced photocatalytic activity of the α-Fe2O3-ZnO nanocomposite (NC), compared to individual α-Fe2O3 and ZnO nanoparticles (NPs) for MB degradation, is attributed to the improved charge separation within the composite structure. Diffuse reflectance spectroscopy (DRS) analysis revealed a reduced band gap energy of 2.73 eV for the α-Fe2O3-ZnO NC, indicating a red shift relative to ZnO NPs (3.16 eV). The rate constant for MB photodegradation at extended reaction times was approximately 5.34 times greater than that observed at earlier stages. Moreover, a comparison between the mineralization rate (k ≈ 0.0192 min−1; t1/2 = 36.1 min) and the degradation rate (k = 0.0119 min−1) over the same time interval suggests that the mineralization process proceeds 1.6 times faster than photodegradation, indicating rapid conversion of MB and its intermediates into inorganic end-products [139]. Graphene oxide (GO) exhibited limited photocatalytic activity, achieving 32% MB degradation within 1 h. This was significantly improved to 82% with the incorporation of ZnO, forming GO@ZnO composites. Further enhancement was observed with Fe doping, achieving up to 99% degradation. The catalytic efficiency increased with the Fe content, though the relationship was nonlinear. The optimum performance was attained at 5% Fe doping, with nearly 96% MB degradation in real river water samples. Under sunlight, the GO@ZnO catalyst showed 90% degradation within 60 min, outperforming pristine GO. When doped with 5% Fe, the composite (GO-ZnO@5Fe) reached complete degradation (100%) within the same period, confirming the critical influence of the Fe doping concentration on photocatalytic performance [140]. Diffuse reflectance UV-Vis showed band gaps of 2.7 eV (IRMOF-3) and 2.8 eV (IRMOF-3/ZnO), with enhanced dye degradation due to the dual energy transfer pathways. IRMOF-3/ZnO offers a cost-effective, visible-light-active photocatalyst. A BiOI/biocarbon composite (2:3 ratio) showed superior MB degradation over pure BiOI (62%) due to the improved charge separation via the biocarbon matrix (Figure 36) [141]. Table 3 shows a summary of metal-based composite photocatalysts used for MB degradation.
Table 3. Summary of metal-based composite photocatalysts for the degradation of MB.
Table 3. Summary of metal-based composite photocatalysts for the degradation of MB.
Metal-Based CompositesSynthetic MethodDegradation EfficiencyBand GapRef.
CoVO/WxOyHydrothermal method96%-[109]
Cu2O-rich and Cu2O-poorElectrodeposition48% for Cu2O-rich and 46% Cu2O-poor~2.0 eV[110]
Cs3Bi2I9Hot-injection method98.2%4.9 eV,[111]
Ag2CO3-ZnCO3Precipitation99.88%0.38 eV[112]
CuO-WO3Co-precipitation92%2.1 to 2.4 eV[113]
PEDOT/Ag2SeO3In situ synthesis46.2%2.41 eV[114]
RGO photocatalyst loaded with Zn0.5Cu0.5Fe2O4Co-precipitation95.2%~1.7 eV[115]
MgFe2O4-TiO2 (MFO-TiO2Hydrothermal approach, followed by a calcination process99.53%3.2 and 3.0 eV[120]
Diatomite-TiO2 compositeImpregnation method80%3 eV[124]
FeSe2–ZnOCost-effective chemical method97%1.91 eV[116]
TiO2/Fe3O4Sol-gel assisted method92%-[117]
TiO2-F-V-MoSol-gel method0.0363%2.52 eV[121]
TiO2-C@NSol hydrothermal method99.87%-[122]
kaolinite/TiO2Sol-gel method98%3.14 eV[123]
TiO2/C-550Sol-gel method100%2.7 eV[78]
B-doped g-C3N4/TiO2Co-precipitation-1.5678 eV[126]
(DM g. C3N4) TiO2Solvothermal97%2.23 eV[130]
Cu-TiO2/g-C3N4Hydrothermal method-2.81 eV[131]
Chromium-TiO2/carbonHydrothermal method90%3.2 eV[128]
CoO/ZnOHydrothermal method67.5%3.44–3.14 eV[132]
GO-ZnO@5FeHydrothermal method100%2.58[140]
2D/0D g-C3N4/TiO2Thermal polycondensation method100%~ 2.7 eV[125]
ZnO/MNCCo-precipitation method97.14%-[133]
Ag/ZnOHydrothermal method99.8%.3.16 eV[138]
TiO2-cMDFCarbonization method99%3.2 eV[127]
TiO2/CuxOSpin-coating technique-1.70 eV[129]
Fe3O4/ZnOSolid state method88.5%3.39 eV[137]
TiO2/C-800Calcination with a one-pot liquid phase reaction25.1%-[97]
Cu/ZnOIncipient wetness impregnation99%3.37 eV[134]
g-CN/ZnOPyrolysis method98%-[136]
α-Fe2O3-ZnO NCPrecipitation method56.9%2.73 eV[139]
BiOI/CHydrothermal method79.6%~ 2.51 eV[141]

6. Metal–Organic-Framework-Based Composites

Metal–organic frameworks (MOFs) are porous crystalline materials composed of metal ions (e.g., Zn2+, Cu2+, Mg2+, Cr3+, Al3+, Fe3+) coordinated to organic ligands (linkers), forming 3D frameworks. Their inherent porosity allows the adsorption of external molecules, making them effective in photocatalysis and water treatment [142]. Most studies on dye degradation using MOFs (Table 4) focus on catalyst performance and dye removal under varying pH [143], catalyst/dye dosage [144], and recyclability [145,146]. HKUST-1, a well-studied Cu-based MOF, consists of Cu2+ paddle-wheel units coordinated with 1,3,5-benzenetricarboxylate (BTC) linkers, forming a face-centered cubic porous structure with pore sizes of 1.4, 1.1, and 0.5 nm (Figure 37a). Each Cu center is bonded to four oxygen atoms from BTC and one water molecule in the hydrated state. Post-synthetic dehydration generates coordinatively unsaturated sites, enhancing HKUST-1’s catalytic activity [147].
TiO2 doped with Bi2O3 and ZnO showed significantly enhanced visible light photocatalytic activity. Bi2O3-ZnO/TiO2 nanofibers achieved the highest MB degradation rate constant of 0.0234 min−1, 4.97 and 16.7 times higher than ZnO/TiO2 and pure TiO2, respectively, due to the improved charge separation and increased active sites [148]. The photocatalytic efficiency was highly pH-dependent (Figure 37b,c), reaching 98.94% at pH 10.2 within 90 min but only 31.15% at pH 4.4 after 180 min. HKUST-1 exhibited a band gap of 3.60 eV, a crystallite size of 19 nm, a surface area of 315 m2/g, and followed first-order degradation kinetics [149]. A novel heterogeneous photosensitizer, MB@UC, was prepared by immobilizing MB onto UiO-66-(COOH)2 (UC) MOFs, as shown in Figure 37d. The MB@UC composite efficiently generated singlet oxygen (1O2) under visible light, enabling selective photodegradation of the sulfur mustard simulant CEES (2-chloroethyl ethyl sulfide) into less toxic sulfoxide within 6 min, avoiding overoxidation to toxic sulfone. The half-life for CEES degradation was only 1.8 min under ambient conditions, demonstrating the strong 1O2 generation and high selectivity of MB@UC [150].
Figure 37. (a) Illustrates the coordination between the copper dimer and the linker, which is shown in the center, the cubic framework of HKUST-1. Reproduced with permission from [147] (copyright 2025, Elsevier). The photocatalytic degradation of MB dye using HKUST-1 is being conducted under different pH conditions: (b) acidic [149], (c) basic [149], and (d) formation of the MB@UC composite. Reproduced with permission from [150] (copyright 2025, Royal Society of Chemistry).
Figure 37. (a) Illustrates the coordination between the copper dimer and the linker, which is shown in the center, the cubic framework of HKUST-1. Reproduced with permission from [147] (copyright 2025, Elsevier). The photocatalytic degradation of MB dye using HKUST-1 is being conducted under different pH conditions: (b) acidic [149], (c) basic [149], and (d) formation of the MB@UC composite. Reproduced with permission from [150] (copyright 2025, Royal Society of Chemistry).
Catalysts 15 00893 g037
The OM-PE@PbBrOH⊂ZIF-67 photocatalyst demonstrated excellent water stability for up to two weeks, retaining high photocatalytic efficiency in aqueous media. Strong N–Pb interfacial coordination facilitated enhanced electron transfer, while the staggered-gap heterostructure enabled high charge mobility (3.51 × 108 s−1), robust redox activity, and prolonged operational stability, contributing to its superior photocatalytic performance [151]. Similarly, the Cu2O(TC)@NH2-MIL-125(Ti) core–shell heterostructure exhibited effective charge recombination suppression and improved visible-light-driven MB degradation. After 120 min of light exposure, the MB removal efficiencies were 38.79% for the core–shell composite versus 74.88% residual MB for Cu2O(TC) alone, indicating higher photocatalytic efficiency and stability (Figure 38a–d). The degradation kinetics followed a pseudo-first-order model [152]. The heterojunctions between MOFs and inorganic semiconductors effectively improve photocatalytic activity and charge separation. Using a water-bath-prepared Ce-MOF as the precursor, a microwave-assisted hydrothermal method was employed to synthesize a Ce-MOF/CdIn2S4/CdS multicomponent heterojunction. The final composite consisted of central CdS spheres surrounded by rod-like Ce-MOF structures, interlaced with CdIn2S4 nanosheets (Figure 38e). With a kinetic rate constant of 0.0089 min−1, the 10% Ce-MOF/CdIn2S4/CdS composite demonstrated a photocatalytic degradation efficiency of 76.7% for ciprofloxacin (CIP) under full-spectrum light in 120 min [153].
The Mo0.09@Ni-MOF photocatalyst demonstrated superior catalytic activity compared to pristine Ni-MOF, achieving 84% MB degradation efficiency versus 68% for Ni-MOF. Mo0.09@Ni-MOF also showed high stability and reusability over four successive cycles. The reaction followed pseudo-first-order kinetics, with rate constants of 0.01703 min−1 for Ni-MOF and 0.01666 min−1 for Mo@Ni-MOF, the latter indicating faster dye degradation kinetics and a strong correlation (R2 = 0.95864) [154]. The Cr-PTC-HIna/TiO2 composite exhibited a band gap of 2.02 eV, rod-shaped morphology, and a particle size of 239.16 nm. Using the Box–Behnken design, optimal MB degradation of 88.55% was achieved at 35.7 ppm MB, pH 6.6, 65.7 mg composite dosage, and 0.13% photocatalyst under 250 W mercury lamp irradiation for 2 h [155]. The CuWO4@MIL-101(Fe) composite, with enhanced active sites and a Z-scheme heterojunction to minimize electron–hole recombination, achieved 96.92% MB degradation in 120 min. It also showed excellent OER performance, with an overpotential of 188 mV and an onset potential of 1.27 V to reach 10 mA cm−2 current density [156]. Bimetallic MOFs were developed by doping Ni into ZIF-8 to enhance MB degradation. The Ni(20)-ZIF-8 composite achieved 93.22% degradation in 150 min using 30 mg of photocatalyst in 50 mL of 30 mg/L MB at neutral pH. This performance was 3.18 times higher than pristine ZIF-8. The composite also retained 76.43% efficiency after five cycles, confirming the excellent reusability [157]. Because of its abundant active sites before UV LED irradiation, Ni-ZIF-8 efficiently adsorbs MB in the dark (Figure 39a). In a model ship test, a hydrophobic composite coating also decreased the drag by 36% thanks to the creation of a gas cavity upon contact with water. The coating demonstrated its dual functionality for marine applications by enabling MB degradation in seawater and displaying self-cleaning qualities (Figure 39b) [158].
The photo-Fenton activity of MnMg-MOF heterojunctions was evaluated for MB degradation under simulated sunlight. Using 0.3 g/L catalyst and 10 mg/L H2O2, an 87.79% decolorization rate was achieved after 180 min, with a TOC removal rate of 88% after three cycles (Figure 40a,b) [159]. Ti3C2 incorporation at 1.5 wt% enhanced the photocatalytic activity of UNiMOF fourfold due to its excellent conductivity and layered structure, which promoted effective charge separation [160]. ZIF-8/Ti3C2Tx, synthesized via in situ growth (Figure 40c), exhibited superior MB adsorption capacity (107 mg/g) compared to Ti3C2Tx (9 mg/g) and ZIF-8 (3 mg/g), attributed to synergistic effects and optimized interlayer spacing (Figure 40d) [161]. Additionally, the UiO-66/MXene composite demonstrated a maximum MB adsorption capacity of 312 mg/g, achieving over 97% removal with just 0.02 g of sorbent from a 20 mg/L MB solution [161]. Copper mixed-triazolate MOFs (CuMtz-3a and CuMtz-3b), synthesized using mixed ligands (3,5-diphenyl-1H-1,2,4-triazole and 1H-1,2,4-triazole), demonstrated superior MB adsorption performance compared to the mono-ligand MOF (CuTz-1), as shown in Figure 41a. CuMtz-3b exhibited an adsorption capacity of 152 mg/g, 182% higher than CuTz-1, and a notable photocatalytic rate constant of 0.031 min−1 for MB degradation. This enhanced activity was attributed to improved visible light absorption and efficient charge separation. The photocatalyst maintained high degradation efficiency and structural stability over four cycles, indicating excellent reusability [162].
UIO-66 and hydroxyl-modified UIO-66, referred to as UIO-66-2OH (2,3), were synthesized via a solvothermal method by Li et al. BET-BJH analysis revealed surface areas of 1050 m2/g for UIO-66 and 570 m2/g for UIO-66-2OH (2,3). The band gap of UIO-66-2OH (2,3) was determined to be 2.6 eV. Under visible light, UIO-66-2OH (2,3) achieved complete degradation of MB within 100 min and maintained 99.5% photocatalytic activity after five cycles [164]. Under simulated sunlight, 10 mg of MOF/CdS-6 (mass ratio of MOF to thioacetamide 6:1) achieved 91.9% MB degradation in 100 min, outperforming pure Cd-MOF (62.3%) and CdS (67.5%). This improvement is attributed to the composite’s larger surface area, increased active sites, and extended visible light response due to CdS incorporation. Mott–Schottky analysis confirmed the formation of a type-II heterojunction between Cd-MOF and CdS, which reduced electron–hole recombination. The photocurrent intensity of MOF/CdS-6 was 8× and 2.5× higher than that of Cd-MOF and CdS, respectively, with excellent stability over five cycles [165]. A photocatalytic degradation rate of 95.87% was observed in a ZIF-8/ZnO composite (1:2 mass ratio) after 40 min of UV exposure. A sponge-like porous C-ZnO/PVDF membrane was fabricated by incorporating C-ZnO into PVDF using extended conversion technology (Figure 41b). At 15 wt% C-ZnO loading, optimal performance was recorded, with 95.01% degradation in 270 min, 83.94% porosity, and a pure water flux of 280.56 L m−2·h−1, showing high potential for wastewater treatment applications [166]. MOF/polymer composites also demonstrated high MB removal efficiencies 96 96% after 30–45 min of UV irradiation using 2% MIL-53(Cr)/polymer and 2% HKUST-1(Cu)/polymer. These catalysts remained effective over 10 cycles, with reduced activity after the seventh cycle. However, their photocatalytic activity was limited to UV light (λ < 365 nm), highlighting the need for MOFs responsive to visible light for broader photocatalytic applications [167,168].
Figure 41. (a) Preparation of CuMtz and Cutz-1 MOFs T. Reproduced with permission from [162]. (b) The synthesis of ZIF-8/ZnO/PVDF. Reproduced with permission from [166] (copyright Elsevier).
Figure 41. (a) Preparation of CuMtz and Cutz-1 MOFs T. Reproduced with permission from [162]. (b) The synthesis of ZIF-8/ZnO/PVDF. Reproduced with permission from [166] (copyright Elsevier).
Catalysts 15 00893 g041
The formation of a heterojunction between the incorporated photocatalytic degradation of MB demonstrated that TiO2/Al2O3@Cu(BDC) achieves superior degradation ae compared to other compounds, and the recombination of photogenerated electron–hole pairs has been reduced. Under visible light illumination, photogenerated electrons are transferred from the conduction bands of Cu(BDC) and Al2O3 to the conduction band of TiO2, resulting in the presence of positive holes (h+) in the valence bands of Cu(BDC) and Al2O3. The photocatalytic degradation of MB has demonstrated that TiO2/Al2O3@Cu(BDC) achieves superior degradation when compared to other compounds [169]. Porous cobalt-based metal–organic frameworks (Co-MOFs) were synthesized and functionalized with amino groups using 1,4-diazabicyclo [2.2.2]octane (DABCO) as a linker. These Co-MOFs were further encapsulated with photosensitive dyes, MB and methyl orange (MO), to form MB@Co-MOF and MO@Co-MOF composites. Their photocatalytic activities under visible light were evaluated for the degradation of representative azo dyes, with the results compared to unmodified Co-MOF. The photodegradation efficiencies for Eriochrome Black T (EBT) reached 78.9%, 92.0%, and 99.7% using Co-MOF, MO@Co-MOF, and MB@Co-MOF, respectively, at 100 mg/L concentration. The corresponding rate constants (K1) were 0.02, 0.05, and 0.08 min−1 over 240 min of visible light exposure [170]. Morphological analysis confirmed uniform CuO dispersion on the CuCr2O4 matrix. The CuCr2O4/CuO nanocomposite achieved ~90% MB degradation in the presence of H2O2, with complete degradation within 35 min [171].
BET analysis showed high surface areas of 1322 m2/g for ZIF-8 and ~1507 m2/g for NDCQDs/ZIF-8. The incorporation of nitrogen-doped carbon quantum dots (NDCQDs) improved dye adsorption and enhanced photocatalytic efficiency under visible light (Figure 42a). This improvement was attributed to the synergistic effects of ZnO in ZIF-8, the up-conversion fluorescence of NDCQDs, and efficient charge separation. Scavenging experiments identified hydroxyl radicals (OH) as the primary reactive species. The MB degradation followed a pseudo-first-order kinetic model, with rate constants of 0.00559 min−1 for ZIF-8 and 0.0103 min−1 for NDCQDs/ZIF-8 (Figure 42b,c) [172].
The CoFe2O4/SiO2/Cu core was functionalized with glutaric anhydride and 3-(triethoxysilyl)propylamine, followed by in situ self-assembly of Cu-MOF nanostructures. The composite exhibited enhanced MB degradation (98% within 30 min) at alkaline pH (pH = 10), with degradation kinetics best described by a pseudo-second-order model. The high activity was attributed to the synergistic effects of multiple transition metals and the mesoporous structure [173]. Mesoporous Fe(25)ZnO, containing 25 wt% α-Fe2O3, achieved nearly complete MB degradation after 150 min of irradiation, outperforming pristine ZnO and α-Fe2O3. It maintained 95.42% efficiency after three reuse cycles, attributed to the formation of an n–n heterojunction enhancing charge separation [174]. The MOF-5/GO10 composite demonstrated 92% MB removal efficiency and followed pseudo-first-order kinetics, with rate constants of 0.0369 and 0.0396 min−1 for MOF-5/GO5 and MOF-5/GO10, respectively (Figure 43) [175].
Table 4. Summary of the photocatalytic performance of various catalysts for MB dye degradation under different experimental conditions, highlighting the effects of the pH, catalyst/dye dosage, and recyclability.
Table 4. Summary of the photocatalytic performance of various catalysts for MB dye degradation under different experimental conditions, highlighting the effects of the pH, catalyst/dye dosage, and recyclability.
MOF-Based CompositesSynthetic MethodDegradation EfficiencyBand GapRef.
Bi2O3-ZnO/TiO2 MOFElectrospinning98%2.9 eV[148]
HKUST-1, a Cu-based MOFSolvothermal method98.94%3.60 eV[149]
MB-modified UiO-66-(COOH)2 MOFAdsorption method--[150]
OM-PE@PbBrOH⊂ZIF-67Solvothermal method72%0.4 eV[151]
Cu2O(TC)@NH2-MIL-125(Ti)Solvothermal method74.88%1.69 eV[152]
Ce-MOF/CdIn2S4/CdSHydrothermal method76.7%2.23 eV[153]
Mo0.09@Ni-MOFSolvothermal method84%1.83 eV[154]
Cr-PTC single bond HIna/TiO2Solvothermal method88.55%2.02 eV[155]
CuWO4@MIL-101(Fe)Solvothermal method96.92%-[156]
Ni(20)-ZIF-8One-step room-temperature method93.22%4.40 eV[157]
ZIF-8/POTSSuperhydrophobic method93.85%4.9 eV[158]
MnMg-MOFOne-step cationic membrane electro-conversion88%0.37 eV[159]
UNiMOF/Ti3C2Electrostatic self-assembly99.49%3.43 eV[160]
ZIF-8/Ti3C2TxIn situ growth--[163]
UiO-66/MXeneSolvothermal method>97%-[161]
CuTz-1One-pot economic synthesis73%1.72 eV[162]
UIO-66-2OH (2,3)Solvothermal method99.5%2.6 eV[164]
Cd-MOF/CdSIn situ sulfurization91.9%2.9 eV[165]
ZIF-8/ZnOPhase inversion method95.01%3.40 eV[166]
TiO2/Al2O3@Cu(BDC)In situ incorporation of pre-synthesized precursors-3.29 eV[169]
MO@Co-MOFSolvothermal method99.7%1.71 eV[170]
NDCQDs/ZIF-8Hydrothermal method28%4.961 eV[172]
CoFe2O4/SiO2/Cu-MOFSol-gel method98%-[173]
MOF-derived α-Fe2O3/ZnOCalcination100%2.34 eV[174]
MOF-5/GOHummer’s method92%3.5 eV[175]

6.1. Recovery and Reusability of Photocatalysts

Methylene blue (MB) photodegradation with different catalysts has been thoroughly investigated, with an emphasis on catalyst efficiency and recovery. Utilizing magnetic nanoparticles and nanocomposites has demonstrated promising outcomes in terms of enhancing the degradation process and facilitating catalyst recovery. The Fe3O4@SiO2@ZnMn2O4@CuIITHPP nanocomposite showed high photocatalytic activity by degrading MB by 98% in 180 min when exposed to visible–LED light. An external magnetic field made it easy to recover the catalyst, demonstrating its usefulness in wastewater treatment [176,177]. Methylene blue degradation is effectively catalyzed by the Fe3O4@SiO2@CeO2 core–shell magnetic nanostructure. An external magnetic field makes it simple to recover the magnetic catalyst following the photocatalytic process. Even after five uses, the catalyst showed over 92% efficiency in removing methylene blue under ideal conditions, suggesting that it could be used repeatedly in wastewater treatment. The study emphasizes how well the catalyst works with photocatalysis to break down dye pollutants [176]. The ZnO/Fe3O4/SiO2 hybrid microbeads are a powerful photocatalyst for methylene blue degradation. After the degradation process, the microbeads can be fully recovered and reused because they contain magnetic nanoparticles. This characteristic makes the catalyst more useful because it makes it simple to remove it from the solution after the reaction. The study showed that these microbeads outperformed conventional ZnO powder in completely degrading methylene blue when exposed to UV light [178]. The effectiveness and reusability of the nanocomposite were demonstrated in the study, which showed that it achieves over 90% photocatalytic efficiency for MB degradation over five cycles. Both UV and sunlight radiation can cause degradation, but UV light accelerates the process. For the removal of MB from aqueous solutions, the Ru/Fe3O4 nanocomposite is therefore a promising catalyst [179].

6.2. Mineralization and Byproduct Formation

TOC analysis is crucial for evaluating the degree of mineralization in photocatalytic processes. Using gadolinium-oxide-decorated multi-walled carbon nanotube/titania nanocomposites, the mineralization of methylene blue (MB) during photocatalytic degradation was evaluated by total organic carbon (TOC) analysis. After five hours, the study found that 80.0% of the TOC had been removed, indicating that MB had effectively broken down into inorganic species and minimized harmful byproducts. Comparing the photocatalytic activity to neat composites and commercial titania, the improved charge separation and electron transfer made possible by the gadolinium oxide nanoparticles resulted in higher degradation efficiencies [180]. Using ZrO2/MWCNT nanocomposites, the study concentrated on the photocatalytic degradation of methylene blue. Analysis of the total organic carbon (TOC) revealed a notable decrease from 148.00 mg/L to 1.26 mg/L, suggesting efficient degradation. After 25–60 min of UV irradiation, the ZrO2/MWCNT catalyst showed high efficiency, achieving 90–94% degradation. This demonstrates the potential of the produced nanocomposite for photocatalytic treatment of industrial wastewater contaminated with dyes, such as methylene blue [181]. Au-TiO2 catalysts were used to assess the photocatalytic degradation of methylene blue (MB) using total organic carbon (TOC) analysis. UV-Vis spectroscopy and TOC measurements were used to track the degradation process, and the results showed that MB had been converted by more than 80%. The findings indicated that different organic intermediate products were formed during the photocatalytic degradation, which followed a first-order kinetic model. The study demonstrated how well photocatalysis works to mineralize organic pollutants into non-toxic byproducts [182].

6.3. Challenges and Limitations

While photocatalysis is showing promising results in the degradation of dyes in lab settings, a number of real-world obstacles limit its widespread use. Salts, naturally occurring organic matter, and competing pollutants frequently reduce photocatalytic efficiency and even deactivate the catalyst surface in actual wastewater, which lowers the stability and reusability over the long term [183,184]. Additionally, uneven light penetration, problems with catalyst recovery, and the requirement for optimized reactor designs make it challenging to scale laboratory results to pilot or industrial scales [185,186]. Finally, because the production of advanced nanomaterials and the energy requirements of UV–visible light sources raise operating costs and make widespread adoption difficult, the economic viability of photocatalysis is a significant concern [187,188].
The structural and electronic characteristics of CuS nanostructures are examined in a study by Sudhaik and his colleague, which also emphasizes the difficulties in maintaining their stability when exposed to extended light. The authors point out that extended exposure may cause CuS to decompose, which would impair its photocatalytic activity [189]. Similarly, Zhang and his colleagues stress that because of their broadband absorption characteristics, binary metal sulfides, such as CdS, can only absorb ultraviolet light for photocatalytic reactions. They add that these sulfides may change structurally when exposed to light, which would reduce their photocatalytic effectiveness [190].
Due to their large band gaps, the most extensively researched photocatalysts, including TiO2, ZnO, and CeO2, are only active in UV light, which makes up less than 5% of the solar spectrum. Additionally, the production of reactive oxygen species is greatly decreased by rapid electron–hole recombination, which lowers the degradation efficiency [191]. Since laboratory research is frequently carried out in simplified systems, actual industrial wastewater contains natural organic matter and inorganic ions (such as carbonates and chlorides) that scavenge reactive radicals or compete for active sites. In real-world situations, these interferences decrease the effectiveness of methylene blue degradation. As the majority of the research literature assesses catalyst performance in controlled environments with ideal pH, dye concentration, and light intensity, these settings rarely accurately represent the makeup of actual wastewater. The complex mixture of ions, suspended solids, and other organic and inorganic components found in actual wastewater can affect the kinetics of pollutant degradation and catalyst behavior [192].
Consequently, factors like the removal efficiency, degradation duration, ideal pH, and catalyst dosage that are determined in laboratory systems might not be directly applicable to wastewater applications in practical wastewater applications [193]. Depending on the catalyst type and ion concentration, the presence of inorganic anions (such as bicarbonate, sulfate, and chloride) has demonstrated both inhibitory and enhancing effects on dye degradation [26]. Synthetic setup studies rarely capture the interactions between the numerous concurrent organic and inorganic pollutants found in wastewater. These results show that to close the gap between laboratory research and field applications, catalyst performance evaluation in realistic matrices is crucial [194,195].

7. Conclusions

This review highlights the remarkable progress made in the development of composite photocatalysts for the degradation of MB dye using photocatalytic phenomena. It covers mainly carbon-based, metal-based based and MOF-based composites and some bio-derived nanocomposites for high MB degradation. The photocatalytic efficiency, in some cases, exceeds 99% within 30–90 min. The photo-composites such as ZnO@g-C3N4, TiO2/C-550, and Ag/TiO2/CNT showed superior activity due to enhanced light harvesting, reduced electron–hole recombination, and high surface area. Notably, Z-scheme and S-scheme architectures proved particularly effective by maintaining strong redox potential while promoting charge carrier separation. Additionally, certain composites like RGO-TiO2−x and NiO/g-C3N4 exhibited long-term stability, maintaining over 90% efficiency across multiple reuse cycles. Despite these advancements, several challenges remain. Many catalysts demonstrate limited performance under real wastewater conditions due to the presence of competing ions, turbidity, or fouling agents. The reproducibility of results across large-scale systems is still uncertain, and the synthesis of some high-performance catalysts involves expensive precursors or complex fabrication steps. Addressing these limitations requires a shift toward more sustainable, cost-effective, and scalable catalyst designs, supported by comprehensive mechanistic studies

Future Perspectives

To bridge the gap between laboratory success and industrial-scale application, several avenues must be pursued. The use of artificial intelligence and machine learning to model degradation kinetics and predict photocatalyst behavior under varying conditions represents an emerging frontier. A shift toward low-cost, earth-abundant, and environmentally benign materials is also essential to meet industrial and sustainability demands. Importantly, organocatalysts offer a promising new direction in this field. Organic acids and small molecules, acting as sustainable catalytic agents, can facilitate degradation of organic pollutants through photooxidative or redox pathways without relying on heavy metals or expensive semiconductors. Future developments of sustainable photocatalytic technologies will be guided by the practical relevance of mechanistic insights combined with pilot-scale studies and industrial case studies. Future research should assess the stability and long-term performance of photocatalysts in continuous-flow reactors, as the majority of current studies are based on batch experiments. These investigations will shed light on the catalyst durability, sustained degradation efficiency, and usefulness for treating industrial wastewater.

Author Contributions

A.M.: Writing the original draft. M.A.I. and T.-O.D.: Conceptualization, resources, supervision, and overall guidance. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Dong, S.; Feng, J.; Fan, M.; Pi, Y.; Hu, L.; Han, X.; Liu, M.; Sun, J.; Sun, J. Recent developments in heterogeneous photocatalytic water treatment using visible light-responsive photocatalysts: A review. RSC Adv. 2015, 5, 14610–14630. [Google Scholar] [CrossRef]
  2. Stojadinović, S.; Radić, N.; Tadić, N.; Vasilić, R.; Grbić, B. Enhanced ultraviolet light driven photocatalytic activity of ZnO particles incorporated by plasma electrolytic oxidation into Al2O3 coatings co-doped with Ce3+. Opt. Mater. 2020, 101, 109768. [Google Scholar] [CrossRef]
  3. Nasrabadi, T. An indexapproach tometallic pollution in riverwaters. Int. J. Environ. Res. 2015, 9, 385–394. [Google Scholar]
  4. Borker, P.; Salker, A. Photocatalytic degradation of textile azo dye over Ce1−xSnxO2 series. Mater. Sci. Eng. B 2006, 133, 55–60. [Google Scholar]
  5. Gupta, V.K.; Ali, I.; Saleh, T.A.; Nayak, A.; Agarwal, S. Chemical treatment technologies for waste-water recycling—An overview. RSC Adv. 2012, 2, 6380–6388. [Google Scholar] [CrossRef]
  6. Verma, N.; Chundawat, T.S.; Chandra, H.; Vaya, D. An efficient time reductive photocatalytic degradation of carcinogenic dyes by TiO2-GO nanocomposite. Mater. Res. Bull. 2023, 158, 112043. [Google Scholar] [CrossRef]
  7. Zhang, Q.-Q.; Ying, G.-G.; Pan, C.-G.; Liu, Y.-S.; Zhao, J.-L. Comprehensive evaluation of antibiotics emission and fate in the river basins of China: Source analysis, multimedia modeling, and linkage to bacterial resistance. Environ. Sci. Technol. 2015, 49, 6772–6782. [Google Scholar] [CrossRef]
  8. Rizzo, L.; Sannino, D.; Vaiano, V.; Sacco, O.; Scarpa, A.; Pietrogiacomi, D. Effect of solar simulated N-doped TiO2 photoca-talysis on the inactivation and antibiotic resistance of an E. coli strain in biologically treated urban wastewater. Appl. Catal. B Environ. 2014, 144, 369–378. [Google Scholar]
  9. Mudhoo, A.; Ramasamy, D.L.; Bhatnagar, A.; Usman, M.; Sillanpää, M. An analysis of the versatility and effectiveness of composts for sequestering heavy metal ions, dyes and xenobiotics from soils and aqueous milieus. Ecotoxicol. Environ. Saf. 2020, 197, 110587. [Google Scholar] [CrossRef]
  10. Al-Tohamy, R.; Ali, S.S.; Li, F.; Okasha, K.M.; Mahmoud, Y.A.-G.; Elsamahy, T.; Jiao, H.; Fu, Y.; Sun, J. A critical review on the treatment of dye-containing wastewater: Ecotoxicological and health concerns of textile dyes and possible remediation ap-proaches for environmental safety. Ecotoxicol. Environ. Saf. 2022, 231, 113160. [Google Scholar]
  11. Gita, S.; Hussan, A.; Choudhury, T. Impact of textile dyes waste on aquatic environments and its treatment. Environ. Ecol. 2017, 35, 2349–2353. [Google Scholar]
  12. Padhi, B. Pollution due to synthetic dyes toxicity & carcinogenicity studies and remediation. Int. J. Environ. Sci. 2012, 3, 940. [Google Scholar]
  13. Mohan, D.; Shukla, S.P. Hazardous consequences of textile mill effluents on soil and their remediation approaches. Clean. Eng. Technol. 2022, 7, 100434. [Google Scholar] [CrossRef]
  14. Muraro, P.C.L.; Mortari, S.R.; Vizzotto, B.S.; Chuy, G.; Dos Santos, C.; Brum, L.F.W.; da Silva, W.L. Iron oxide nanocatalyst with titanium and silver nanoparticles: Synthesis, characterization and photocatalytic activity on the degradation of Rhodamine B dye. Sci. Rep. 2020, 10, 3055. [Google Scholar] [CrossRef]
  15. Zollinger, H. Color Chemistry: Syntheses, Properties, and Applications of Organic Dyes and Pigments; John Wiley & Sons: Hoboken, NJ, USA, 2003. [Google Scholar]
  16. Konstantinou, I.K.; Albanis, T.A. TiO2-assisted photocatalytic degradation of azo dyes in aqueous solution: Kinetic and mechanistic investigations: A review. Appl. Catal. B Environ. 2004, 49, 1–14. [Google Scholar] [CrossRef]
  17. Tang, W.Z.; An, H. UV/TiO2 photocatalytic oxidation of commercial dyes in aqueous solutions. Chemosphere 1995, 31, 4157–4170. [Google Scholar] [CrossRef]
  18. Prado, A.G.; Bolzon, L.B.; Pedroso, C.P.; Moura, A.O.; Costa, L.L. Nb2O5 as efficient and recyclable photocatalyst for indigo carmine degradation. Appl. Catal. B Environ. 2008, 82, 219–224. [Google Scholar] [CrossRef]
  19. Zinicovscaia, I. Conventional methods of wastewater treatment. In Cyanobacteria for Bioremediation of Wastewaters; Springer International Publishing: Berlin/Heidelberg, Germany, 2016; pp. 17–25. [Google Scholar]
  20. Forgacs, E.; Cserháti, T.; Oros, G. Removal of synthetic dyes from wastewaters: A review. Environ. Int. 2004, 30, 953–971. [Google Scholar] [CrossRef] [PubMed]
  21. Arslan, I.; Balcioglu, I.A. Advanced oxidation of raw and biotreated textile industry wastewater with O3, H2O2/UV-C and their sequential application. J. Chem. Technol. Biotechnol. 2001, 76, 53–60. [Google Scholar]
  22. Chakrabarti, S.; Dutta, B.K. Photocatalytic degradation of model textile dyes in wastewater using ZnO as semiconductor cat-alyst. J. Hazard. Mater. 2004, 112, 269–278. [Google Scholar] [CrossRef]
  23. Reddy, M.P.; Venugopal, A.; Subrahmanyam, M. Hydroxyapatite photocatalytic degradation of calmagite (an azo dye) in aqueous suspension. Appl. Catal. B Environ. 2007, 69, 164–170. [Google Scholar] [CrossRef]
  24. Stylidi, M.; Kondarides, D.I.; Verykios, X.E. Visible light-induced photocatalytic degradation of Acid Orange 7 in aqueous TiO2 suspensions. Appl. Catal. B Environ. 2004, 47, 189–201. [Google Scholar]
  25. Wang, F.; Min, S.; Han, Y.; Feng, L. Visible-light-induced photocatalytic degradation of methylene blue with polyani-line-sensitized TiO2 composite photocatalysts. Superlattices Microstruct. 2010, 48, 170–180. [Google Scholar] [CrossRef]
  26. Chahar, D.; Kumar, D.; Thakur, P.; Thakur, A. Visible light induced photocatalytic degradation of methylene blue dye by using Mg doped Co-Zn nanoferrites. Mater. Res. Bull. 2023, 162, 112205. [Google Scholar] [CrossRef]
  27. Jiang, G.; Lin, Z.; Chen, C.; Zhu, L.; Chang, Q.; Wang, N.; Wei, W.; Tang, H. TiO2 nanoparticles assembled on graphene oxide nanosheets with high photocatalytic activity for removal of pollutants. Carbon 2011, 49, 2693–2701. [Google Scholar] [CrossRef]
  28. Fernández, C.; Larrechi, M.S.; Callao, M.P. An analytical overview of processes for removing organic dyes from wastewater effluents. TrAC Trends Anal. Chem. 2010, 29, 1202–1211. [Google Scholar] [CrossRef]
  29. Fouad, K.; Bassyouni, M.; Alalm, M.G.; Saleh, M.Y. Recent developments in recalcitrant organic pollutants degradation using immobilized photocatalysts. Appl. Phys. A 2021, 127, 612. [Google Scholar] [CrossRef]
  30. Bhatia, D.; Sharma, N.R.; Singh, J.; Kanwar, R.S. Biological methods for textile dye removal from wastewater: A review. Crit. Rev. Environ. Sci. Technol. 2017, 47, 1836–1876. [Google Scholar] [CrossRef]
  31. Chen, B.; Bu, Y.; Yang, J.; Nian, W.; Hao, S. Methods for total organic halogen (TOX) analysis in water: Past, present, and future. Chem. Eng. J. 2020, 399, 125675. [Google Scholar] [CrossRef]
  32. Cairone, S.; Hegab, H.M.; Khalil, H.; Nassar, L.; Wadi, V.S.; Naddeo, V.; Hasan, S.W. Novel eco-friendly polylactic acid nanocomposite integrated membrane system for sustainable wastewater treatment: Performance evaluation and antifouling analysis. Sci. Total Environ. 2024, 912, 168715. [Google Scholar] [CrossRef]
  33. El Messaoudi, N.; El Khomri, M.; El Mouden, A.; Bouich, A.; Jada, A.; Lacherai, A.; Iqbal, H.M.; Mulla, S.I.; Kumar, V.; Amé-rico-Pinheiro, J.H.P. Regeneration and reusability of non-conventional low-cost adsorbents to remove dyes from wastewaters in multiple consecutive adsorption–desorption cycles: A review. Biomass Convers. Biorefinery 2024, 14, 11739–11756. [Google Scholar] [CrossRef]
  34. Ang, W.L.; Mohammad, A.W. State of the art and sustainability of natural coagulants in water and wastewater treatment. J. Clean. Prod. 2020, 262, 121267. [Google Scholar] [CrossRef]
  35. dos Santos, J.D.; Veit, M.T.; Juchen, P.T.; da Cunha Gonçalves, G.; Palacio, S.M.; Fagundes-Klen, M. Use of different coagulants for cassava processing wastewater treatment. J. Environ. Chem. Eng. 2018, 6, 1821–1827. [Google Scholar] [CrossRef]
  36. Moghaddas, S.M.T.H.; Elahi, B.; Darroudi, M.; Javanbakht, V. Green synthesis of hexagonal-shaped zinc oxide nanosheets using mucilage from flaxseed for removal of methylene blue from aqueous solution. J. Mol. Liq. 2019, 296, 111834. [Google Scholar] [CrossRef]
  37. Qi, Q.; Zhang, T.; Yu, Q.; Wang, R.; Zeng, Y.; Liu, L.; Yang, H. Properties of humidity sensing ZnO nanorods-base sensor fab-ricated by screen-printing. Sens. Actuators B Chem. 2008, 133, 638–643. [Google Scholar] [CrossRef]
  38. Wang, C.; Liu, H.; Qu, Y. TiO2-based photocatalytic process for purification of polluted water: Bridging fundamentals to applications. J. Nanomater. 2013, 2013, 319637. [Google Scholar] [CrossRef]
  39. Anwar, D.I.; Mulyadi, D. Synthesis of Fe-TiO2 composite as a photocatalyst for degradation of methylene blue. Procedia Chem. 2015, 17, 49–54. [Google Scholar]
  40. Xia, S.-J.; Liu, F.-X.; Ni, Z.-M.; Shi, W.; Xue, J.-L.; Qian, P.-P. Ti-based layered double hydroxides: Efficient photocatalysts for azo dyes degradation under visible light. Appl. Catal. B Environ. 2014, 144, 570–579. [Google Scholar] [CrossRef]
  41. Yang, Y.; Xu, L.; Wang, H.; Wang, W.; Zhang, L. TiO2/graphene porous composite and its photocatalytic degradation of methylene blue. Mater. Des. 2016, 108, 632–639. [Google Scholar] [CrossRef]
  42. Atchudan, R.; Edison, T.N.J.I.; Perumal, S.; Karthikeyan, D.; Lee, Y.R. Facile synthesis of zinc oxide nanoparticles decorated graphene oxide composite via simple solvothermal route and their photocatalytic activity on methylene blue degradation. J. Photochem. Photobiol. B Biol. 2016, 162, 500–510. [Google Scholar] [CrossRef]
  43. Lv, T.; Pan, L.; Liu, X.; Lu, T.; Zhu, G.; Sun, Z. Enhanced photocatalytic degradation of methylene blue by ZnO-reduced graphene oxide composite synthesized via microwave-assisted reaction. J. Alloys Compd. 2011, 509, 10086–10091. [Google Scholar] [CrossRef]
  44. Saleh, R.; Taufik, A. Degradation of methylene blue and congo-red dyes using Fenton, photo-Fenton, sono-Fenton, and sonophoto-Fenton methods in the presence of iron (II, III) oxide/zinc oxide/graphene (Fe3O4/ZnO/graphene) composites. Sep. Purif. Technol. 2019, 210, 563–573. [Google Scholar] [CrossRef]
  45. Syed, N.; Feng, Y.; Fahad, R.; Huang, J.; Mahar, F.K. Carbon-based composite nanofibers for photocatalytic degradation of methylene blue dye under visible light. Polym. Adv. Technol. 2023, 34, 2286–2297. [Google Scholar] [CrossRef]
  46. Asokan, J.; Kumar, P.; Arjunan, G.; Shalini, M.G. Photocatalytic Performance of Spinel Ferrites and their Carbon-Based Composites for Environmental Pollutant Degradation. J. Clust. Sci. 2025, 36, 42. [Google Scholar] [CrossRef]
  47. Ma, B.; Hu, W.; Zhou, Y.; Li, Y.; Zhang, Y.; Zhou, M. Synthesis of MOF@ COF composites and study on dye adsorption prop-erties. Inorg. Chem. Commun. 2025, 172, 113475. [Google Scholar] [CrossRef]
  48. Liu, Y.; Yang, J.; Wu, J.; Jiang, Z.; Zhang, X.; Meng, F. The Application of Multifunctional Metal–Organic Frameworks for the Detection, Adsorption, and Degradation of Contaminants in an Aquatic Environment. Molecules 2025, 30, 1336. [Google Scholar] [CrossRef] [PubMed]
  49. Ganesan, M.; Chinnuraj, I.P.; Rajendran, R.; Rojviroon, T.; Rojviroon, O.; Thangavelu, P.; Sirivithayapakorn, S. Development of eco-friendly SrTiO3/multiwalled carbon nanotube (STO/MWCNT) composite with enhanced performance for photocatalytic applications in environment remediation and energy storage. Diam. Relat. Mater. 2025, 155, 112254. [Google Scholar] [CrossRef]
  50. Sirusy, S.; Ashrafi, H.; Akhond, M. Synthesis and Utilization of rGO/Ultrathin Nanotube Bi5O7I for Photodegradation of Methylene Blue and Photoreduction of Cr6+ to Cr3+ toward Detoxification of Water. ACS Appl. Nano Mater. 2025, 8, 3927–3941. [Google Scholar] [CrossRef]
  51. Arya, K.; Kumar, A.; Sharma, A.; Thakur, K.; Kumar, R.; Mehta, S.K.; Singh, S.; Kumar, V.; Kataria, R. Light-assisted synergistic effect of Zn-MOF@ rGO nanocomposite for methylene blue degradation and toxicity analysis to water reclamation. Inorg. Chem. Commun. 2025, 173, 113768. [Google Scholar] [CrossRef]
  52. Mazarji, M.; Mahmoodi, N.M.; Bidhendi, G.N.; Li, A.; Li, M.; James, A.; Mahmoodi, B.; Pan, J. Synthesis, characterization, and enhanced photocatalytic dye degradation: Optimizing graphene-based ZnO-CdSe nanocomposites via response surface methodology. J. Alloys Compd. 2025, 1010, 177999. [Google Scholar] [CrossRef]
  53. Wang, F.; Li, Q.; Xu, D. Recent progress in semiconductor-based nanocomposite photocatalysts for solar-to-chemical energy conversion. Adv. Energy Mater. 2017, 7, 1700529. [Google Scholar] [CrossRef]
  54. Majeed, A.; Ibrahim, A.H.; Al-Rawi, S.S.; Iqbal, M.A.; Kashif, M.; Yousif, M.; Abidin, Z.U.; Ali, S.; Arbaz, M.; Hussain, S.A. Green organo-photooxidative method for the degradation of methylene blue dye. ACS Omega 2024, 9, 12069–12083. [Google Scholar] [CrossRef] [PubMed]
  55. Tranfield, D.; Denyer, D.; Smart, P. Towards a methodology for developing evidence-informed management knowledge by means of systematic review. Br. J. Manag. 2003, 14, 207–222. [Google Scholar] [CrossRef]
  56. Centobelli, P.; Cerchione, R.; Esposito, E. Environmental sustainability in the service industry of transportation and logistics service providers: Systematic literature review and research directions. Transp. Res. Part D Transp. Environ. 2017, 53, 454–470. [Google Scholar] [CrossRef]
  57. Humayun, M.; Ullah, H.; Usman, M.; Habibi-Yangjeh, A.; Tahir, A.A.; Wang, C.; Luo, W. Perovskite-type lanthanum ferrite based photocatalysts: Preparation, properties, and applications. J. Energy Chem. 2022, 66, 314–338. [Google Scholar] [CrossRef]
  58. Zhang, F.; Wang, X.; Liu, H.; Liu, C.; Wan, Y.; Long, Y.; Cai, Z. Recent advances and applications of semiconductor photo-catalytic technology. Appl. Sci. 2019, 9, 2489. [Google Scholar] [CrossRef]
  59. Lettieri, S.; Pavone, M.; Fioravanti, A.; Santamaria Amato, L.; Maddalena, P. Charge carrier processes and optical properties in TiO2 and TiO2-based heterojunction photocatalysts: A review. Materials 2021, 14, 1645. [Google Scholar] [CrossRef]
  60. Kumar, S.G.; Devi, L.G. Review on modified TiO2 photocatalysis under UV/visible light: Selected results and related mecha-nisms on interfacial charge carrier transfer dynamics. J. Phys. Chem. A 2011, 115, 13211–13241. [Google Scholar] [CrossRef]
  61. Zhang, L.; Ran, J.; Qiao, S.-Z.; Jaroniec, M. Characterization of semiconductor photocatalysts. Chem. Soc. Rev. 2019, 48, 5184–5206. [Google Scholar] [CrossRef]
  62. Sharma, E.; Thakur, V.; Sangar, S.; Singh, K. Recent progress on heterostructures of photocatalysts for environmental reme-diation. Mater. Today Proc. 2020, 32, 584–593. [Google Scholar] [CrossRef]
  63. Sheng, H.; Li, Q.; Ma, W.; Ji, H.; Chen, C.; Zhao, J. Photocatalytic degradation of organic pollutants on surface anionized TiO2: Common effect of anions for high hole-availability by water. Appl. Catal. B Environ. 2013, 138, 212–218. [Google Scholar] [CrossRef]
  64. Nosaka, Y.; Nosaka, A.Y. Generation and detection of reactive oxygen species in photocatalysis. Chem. Rev. 2017, 117, 11302–11336. [Google Scholar] [CrossRef]
  65. Arora, I.; Chawla, H.; Chandra, A.; Sagadevan, S.; Garg, S. Advances in the strategies for enhancing the photocatalytic activity of TiO2: Conversion from UV-light active to visible-light active photocatalyst. Inorg. Chem. Commun. 2022, 143, 109700. [Google Scholar] [CrossRef]
  66. Djurišić, A.B.; He, Y.; Ng, A. Visible-light photocatalysts: Prospects and challenges. Apl. Mater. 2020, 8, 030903. [Google Scholar] [CrossRef]
  67. Nasrollahzadeh, M.; Sajjadi, M.; Iravani, S.; Varma, R.S. Carbon-based sustainable nanomaterials for water treatment: State-of-art and future perspectives. Chemosphere 2021, 263, 128005. [Google Scholar] [CrossRef] [PubMed]
  68. Ayanda, O.S.; Mmuoegbulam, A.O.; Okezie, O.; Durumin Iya, N.I.; Mohammed, S.a.E.; James, P.H.; Muhammad, A.B.; Unimke, A.A.; Alim, S.A.; Yahaya, S.M. Recent progress in carbon-based nanomaterials: Critical review. J. Nanoparticle Res. 2024, 26, 106. [Google Scholar] [CrossRef]
  69. Feng, S.; Chen, T.; Liu, Z.; Shi, J.; Yue, X.; Li, Y. Z-scheme CdS/CQDs/g-C3N4 composites with visible-near-infrared light re-sponse for efficient photocatalytic organic pollutant degradation. Sci. Total Environ. 2020, 704, 135404. [Google Scholar] [CrossRef] [PubMed]
  70. Abimannan, G.; Sengodan, P.; Ravichandran, S.; Mary Anjalin, F.; Kumar, K.R.; Maadeswaran, P. Structural, optical, mor-phological and charge transfer properties of CeO2/MWCNTs nanocomposite and their photocatalytic activity of organic dye degradation. J. Sol-Gel Sci. Technol. 2023, 105, 625–636. [Google Scholar]
  71. Jiang, G.; Zheng, X.; Wang, Y.; Li, T.; Sun, X. Photo-degradation of methylene blue by multi-walled carbon nanotubes/TiO2 composites. Powder Technol. 2011, 207, 465–469. [Google Scholar] [CrossRef]
  72. Ye, X.; Wang, Z.; Wang, Q.; Chen, D.; Lin, Y.; Liu, S. Enhanced photocatalytic activity of ternary multilayered Ag/TiO2/CNT composites for methylene blue degradation. Micro Nano Lett. 2019, 14, 771–776. [Google Scholar] [CrossRef]
  73. Yu, Y.; Jimmy, C.Y.; Chan, C.-Y.; Che, Y.-K.; Zhao, J.-C.; Ding, L.; Ge, W.-K.; Wong, P.-K. Enhancement of adsorption and photocatalytic activity of TiO2 by using carbon nanotubes for the treatment of azo dye. Appl. Catal. B Environ. 2005, 61, 1–11. [Google Scholar] [CrossRef]
  74. Mashkoor, F.; Nasar, A. Inamuddin Carbon nanotube-based adsorbents for the removal of dyes from waters: A review. Environ. Chem. Lett. 2020, 18, 605–629. [Google Scholar] [CrossRef]
  75. Pan, H.; Tianyang, L.; Keting, L.; Xu, L. Preparation and properties of high-efficiency lignin-carbon based photocatalytic composites for the degradation of dye wastewater under visible light. J. Wood Chem. Technol. 2025, 45, 11–22. [Google Scholar] [CrossRef]
  76. Bai, H.; Xiong, R.; Wang, N.; Tian, M.; Zhao, J.; Tang, F.; Jiang, J. Synergistic effects of rare-metal ytterbium doping on TiO2/g-C3N5 heterostructures for enhanced photocatalytic degradation of methylene blue. Inorg. Chem. Commun. 2025, 175, 114159. [Google Scholar] [CrossRef]
  77. Yadav, P.R.; Bhagat, P.R. Metal Free Innovative Approach for the Removal of Methylene Blue from Simulated Water Sample by Novel Porphyrin Based Photocatalyst. Opt. Mater. 2025, 160, 116757. [Google Scholar] [CrossRef]
  78. Song, Y.-J.; Li, H.-C.; Xiong, Z.-W.; Cheng, L.; Du, M.; Liu, Z.-Q.; Li, J.; Li, D.-Q. TiO2/carbon composites from waste sawdust for methylene blue photodegradation. Diam. Relat. Mater. 2023, 136, 109918. [Google Scholar] [CrossRef]
  79. Sriramoju, J.B.; Muniyappa, M.; Marilingaiah, N.R.; Sabbanahalli, C.; Shetty, M.; Mudike, R.; Chitrabanu, C.P.; Shivaramu, P.D.; Nagaraju, G.; Rangappa, K.S.; et al. Carbon-based TiO2−x heterostructure nanocomposites for enhanced photocatalytic degradation of dye molecules. Ceram. Int. 2021, 47, 10314–10321. [Google Scholar] [CrossRef]
  80. Ma, B.; Zhu, J.; Xu, Y.; Zhang, L.; Liu, D.; Chen, C.; Sun, B. Photocatalytic degradation of methylene blue using ZnO modified with nano-biochar derived from bacterial cellulose. Ceram. Int. 2025, 51, 14948–14956. [Google Scholar] [CrossRef]
  81. Ullah, S.; Kumar, O.P.; Ali, R.; Riaz, N.N.; Ahmad, A.; Tufail, M.K.; Muhammad, N.; Rehman, A.u.; Shah, S.S.A.; Nazir, M.A. Biomass Derived Hybrid Activated Carbon/Iron Oxide Composite for Photodegradation of Methylene Blue Upon Visible Light. ChemistrySelect 2025, 10, e202405028. [Google Scholar] [CrossRef]
  82. Pourali, S.; Amrollahi, R.; Alamolhoda, S.; Masoudpanah, S. In situ synthesis of ZnO/g-C3N4 based composites for photo-degradation of methylene blue under visible light. Sci. Rep. 2025, 15, 462. [Google Scholar] [CrossRef]
  83. Elias, M.; Alam, R.; Khatun, S.; Hossain, M.S.; Shah, S.S.; Aziz, M.A.; Uddin, M.N.; Hossain, M.A. Hydrothermal synthesis of carboxylated functionalized jute stick carbon and reduced graphene oxide based ZnO nanocomposite photocatalysts: A comparative study. J. Ind. Eng. Chem. 2025, 143, 424–436. [Google Scholar] [CrossRef]
  84. Gindose, T.G.; Atisme, T.B.; Hailegebreal, T.D.; Zereffa, E.A. ZnO Modified g-C3N4–MnO2 Composite for Photodegradation of Methylene Blue. ChemistrySelect 2025, 10, e202403418. [Google Scholar] [CrossRef]
  85. Shanmugam, P.; Parasuraman, B.; Mandlimath, T.R.; Smith, S.M.; Thangavelu, P.; Tangjaideborisu, Y.; Nakorn, P.N.; Boonyuen, S. Synergistic effect of boron and sulphur co-doping g-C3N4 nanosheet/Ag2S heterojunctions for high-performance visible light-driven photocatalytic methylene blue. Inorg. Chem. Commun. 2025, 174, 113912. [Google Scholar] [CrossRef]
  86. Tahir, S.; Zahid, M.; Hanif, M.; Bhatti, I.; Naqvi, S.; Bhatti, H.; Jilani, A.; Alshareef, S.; El-Sharnouby, M.; Shahid, I. The syner-gistic effect of g-C3N4/GO/CuFe2O4 for efficient sunlight-driven photocatalytic degradation of methylene blue. Int. J. Environ. Sci. Technol. 2025, 22, 4829–4846. [Google Scholar] [CrossRef]
  87. Shanmugam, P.; Parasuraman, B.; Boonyuen, S.; Thangavelu, P.; AlSalhi, M.S.; Zheng, A.L.T.; Viji, A. Hydrothermal synthesis and photocatalytic application of ZnS-Ag composites based on biomass-derived carbon aerogel for the visible light degradation of methylene blue. Environ. Geochem. Health 2024, 46, 92. [Google Scholar] [CrossRef] [PubMed]
  88. Sun, F.; Xu, D.; Xie, Y.; Liu, F.; Wang, W.; Shao, H.; Ma, Q.; Yu, H.; Yu, W.; Dong, X. Tri-functional aerogel photocatalyst with an S-scheme heterojunction for the efficient removal of dyes and antibiotic and hydrogen generation. J. Colloid Interface Sci. 2022, 628, 614–626. [Google Scholar] [CrossRef]
  89. Zhang, X.; Sathiyaseelan, A.; Zhang, L.; Lu, Y.; Jin, T.; Wang, M.-H. Zirconium and cerium dioxide fabricated activated car-bon-based nanocomposites for enhanced adsorption and photocatalytic removal of methylene blue and tetracycline hydro-chloride. Environ. Res. 2024, 261, 119720. [Google Scholar]
  90. Ji, X.; Tan, Z.; Yang, H.; Shi, Z.; Yang, J.; Alhadhrami, A.; Zhang, J.; Mersal, G.A.; El-Bahy, Z.M.; Guo, Z. Sustainable bamboo charcoal based nanocomposite catalysts for rapid adsorption and photo-Fenton degradation of toxic dyes. Sustain. Mater. Technol. 2024, 41, e01080. [Google Scholar] [CrossRef]
  91. Shinde, V.; Tanwade, P.; Katayama, T.; Furube, A.; Sathe, B.; Koinkar, P. Ternary composite WS2/GO/Au synthesized from laser ablation and hydrothermal method for photo-and electro-chemical degradation of methylene blue dye. Surf. Interfaces 2024, 46, 104067. [Google Scholar] [CrossRef]
  92. Ahmed, M.A.; Ahmed, M.A.; Mohamed, A.A. Fabrication of NiO/g-C3N4 Z-scheme heterojunction for enhanced photocata-lytic degradation of methylene blue dye. Opt. Mater. 2024, 151, 115339. [Google Scholar] [CrossRef]
  93. Ragupathi, H.; Mariadhas, J.; Venkatesh, K.; Inbanathan, S.; Choe, Y. Decorating 2D graphene oxides sheets with spherical shaped Fe3O4 for the applications of supercapacitors and sunlight induced sonophotocatalytic degradation of methylene blue dye. Colloids Surf. A Physicochem. Eng. Asp. 2024, 683, 132927. [Google Scholar] [CrossRef]
  94. Anum, A.; Nazir, M.A.; Shah, S.S.A.; Elnaggar, A.Y.; Mahmoud, M.; El-Bahy, S.M.; Malik, M.; Wattoo, M.A.; Rehman, A.u. Advanced Nix/MoSx/MOF-2@ g-C3N4 carbon nanostructures for the effective eradication of the Methylene blue dye. Fuller. Nanotub. Carbon Nanostructures 2024, 32, 1103–1114. [Google Scholar] [CrossRef]
  95. Ramírez-Aparicio, J.; Samaniego-Benítez, J.E.; Murillo-Tovar, M.A.; Benítez-Benítez, J.L.; Muñoz-Sandoval, E.; Gar-cía-Betancourt, M.L. Removal and surface photocatalytic degradation of methylene blue on carbon nanostructures. Diam. Relat. Mater. 2021, 119, 108544. [Google Scholar] [CrossRef]
  96. Ahmad, A.; Jini, D.; Aravind, M.; Parvathiraja, C.; Ali, R.; Kiyani, M.Z.; Alothman, A. A novel study on synthesis of egg shell based activated carbon for degradation of methylene blue via photocatalysis. Arab. J. Chem. 2020, 13, 8717–8722. [Google Scholar] [CrossRef]
  97. Cai, J.; Hu, S.; Xiang, J.; Zhang, H.; Men, D. The effect of graphitized carbon on the adsorption and photocatalytic degradation of methylene blue over TiO2/C composites. RSC Adv. 2020, 10, 40830–40842. [Google Scholar] [CrossRef]
  98. Wu, L.; Chen, Y.; Li, Y.; Meng, Q.; Duan, T. Functionally integrated g-C3N4@ wood-derived carbon with an orderly inter-connected porous structure. Appl. Surf. Sci. 2021, 540, 148440. [Google Scholar] [CrossRef]
  99. Atchudan, R.; Edison, T.N.J.I.; Mani, S.; Perumal, S.; Vinodh, R.; Thirunavukkarasu, S.; Lee, Y.R. Facile synthesis of a novel nitrogen-doped carbon dot adorned zinc oxide composite for photodegradation of methylene blue. Dalton Trans. 2020, 49, 17725–17736. [Google Scholar] [CrossRef]
  100. Dai, Z.; Ren, P.; Cao, Q.; Gao, X.; He, W.; Xiao, Y.; Jin, Y.; Ren, F. Synthesis of TiO2@ lignin based carbon nanofibers composite materials with highly efficient photocatalytic to methylene blue dye. J. Polym. Res. 2020, 27, 1–12. [Google Scholar] [CrossRef]
  101. Nawaz, A.; Goudarzi, S.; Saravanan, P.; Zarrin, H. Z-scheme induced g-C3N4/WS2 heterojunction photocatalyst with im-proved electron mobility for enhanced solar photocatalysis. Sol. Energy 2021, 228, 53–67. [Google Scholar] [CrossRef]
  102. Karimi, M.A.; Atashkadi, M.; Ranjbar, M.; Habibi-Yangjeh, A. Novel visible-light-driven photocatalyst of NiO/Cd/g-C3N4 for enhanced degradation of methylene blue. Arab. J. Chem. 2020, 13, 5810–5820. [Google Scholar] [CrossRef]
  103. Jayaprakash, R.; Dineshbabu, N.; Selvaraj, S.; Vignesh, S.; Arun, T.; Ravichandran, K. Hydrothermally constructed and visi-ble-light activated efficient NiO/ZnO/g-C3N4 ternary nanocomposites for methylene blue dye degradation and antibacterial applications. Inorg. Chem. Commun. 2024, 159, 111643. [Google Scholar] [CrossRef]
  104. Utomo, W.P.; Afifah, P.A.I.; Rozafia, A.I.; Mahardika, A.A.; Santoso, E.; Liu, R.; Hartanto, D. Modulation of particle size and morphology of zinc oxide in graphitic carbon nitride/zinc oxide composites for enhanced photocatalytic degradation of methylene blue. Surf. Interfaces 2024, 46, 104017. [Google Scholar] [CrossRef]
  105. Zhu, Z.; Yang, P.; Li, X.; Luo, M.; Zhang, W.; Chen, M.; Zhou, X. Green preparation of palm powder-derived carbon dots co-doped with sulfur/chlorine and their application in visible-light photocatalysis. Spectrochim. Acta Part A Mol. Biomol. Spec-trosc. 2020, 227, 117659. [Google Scholar] [CrossRef]
  106. Son, B.T.; Long, N.V.; Hang, N.T.N. The development of biomass-derived carbon-based photocatalysts for the visi-ble-light-driven photodegradation of pollutants: A comprehensive review. RSC Adv. 2021, 11, 30574–30596. [Google Scholar] [CrossRef]
  107. Guo, M.; Hu, Y.; Wang, R.; Yu, H.; Sun, L. Molecularly imprinted polymer-based photocatalyst for highly selective degradation of methylene blue. Environ. Res. 2021, 194, 110684. [Google Scholar] [CrossRef]
  108. Ghorai, K.; Panda, A.; Bhattacharjee, M.; Mandal, D.; Hossain, A.; Bera, P.; Seikh, M.M.; Gayen, A. Facile synthesis of CuCr2O4/CeO2 nanocomposite: A new Fenton like catalyst with domestic LED light assisted improved photocatalytic activity for the degradation of RhB, MB and MO dyes. Appl. Surf. Sci. 2021, 536, 147604. [Google Scholar] [CrossRef]
  109. Zubair, A.; Nawaz, F.; Fareed, I.; Khan, M.D.; Ali, Z.; Nawaz, M.; Anam, H.S.; Tahir, M.; Butt, F.K. Novel CoVO/WxOy composites for methylene blue photodegradation and electrocatalytic applications. Mater. Sci. Semicond. Process. 2025, 186, 109104. [Google Scholar] [CrossRef]
  110. Maulana, M.A.R.; TriGoutomo, B.; Mahmud, A. The effect of Phase Composition on the Photocatalytic Activity of Cu2O/CuO/Cu Composites for Methylene Blue Photodegradation. Chem. Mater. 2025, 4, 1–8. [Google Scholar] [CrossRef]
  111. Tang, T.; Zhang, H.; Wang, H.; Dou, X.; Wen, J.; Jiang, L. Composite ZIF-8 with Cs3Bi2I9 to Enhance the Photodegradation Ability on Methylene Blue. Molecules 2025, 30, 1413. [Google Scholar] [CrossRef] [PubMed]
  112. Długosz, O.; Chlebowska, Z.; Banach, M. Preparation of Materials Based on Metal Carbonate Nanoparticles for Photodegra-dation of Organic Pollutants. J. Clust. Sci. 2025, 36, 55. [Google Scholar] [CrossRef]
  113. Urooj, N.; Arshad, M.; Arshad, R.; Intisar, A.; Batool, M.; Bashir, F.; Kousar, R. Remarkable visible light actuated photodeg-radation of methylene blue using copper oxide/tungsten oxide heterojunction composites. Mater. Sci. Eng. B 2025, 315, 118079. [Google Scholar] [CrossRef]
  114. Rethnakumaran, A.V.; Menamparambath, M.M. In Situ Generation of Poly (3, 4-ethylenedioxythiophene)/Ag2SeO3 Nano-hybrids at Hexane/Water Interface for Photodegradation of Organic Dyes. Macromol. Mater. Eng. 2025, 310, 2400409. [Google Scholar] [CrossRef]
  115. Abuzeyad, O.H.; El-Khawaga, A.M.; Tantawy, H.; Gobara, M.; Elsayed, M.A. Merits photocatalytic activity of rGO/zinc copper ferrite magnetic nanocatalyst for photodegradation of methylene blue (MB) dye. Discov. Nano 2025, 20, 1–16. [Google Scholar] [CrossRef] [PubMed]
  116. Awais, M.; Hussain, R.; Shah, A.; Alajmi, M.F.; Maryam, R.; Hussain, A.; Khan, S.U.; ur Rahman, S. Design of novel FeSe2/ZnO composites and their application in photodegradation of methylene blue. Phys. B Condens. Matter 2024, 685, 416048. [Google Scholar] [CrossRef]
  117. Borhani, M.N.; Tavakoli, A.; Mollaei, A.M.; Borhani, T.N. Visible light photodegradation of methylene blue by ionic liquid based TiO2/Fe3O4 nanophotocatalysts. Opt. Mater. 2024, 154, 115627. [Google Scholar] [CrossRef]
  118. Bukit, B.F.; Pratama, A.W.; Frida, E.; Sedayu, B.B.; Fransiska, D.; Purnomo, D.; Rochima, E.; Rahmawati, I.; Suhartana, S.; Syamani, F.A. Eco-friendly alginate/PCL-TiO2 hybrid biocomposites: Preparation, properties, and methylene blue photodeg-radation. South Afr. J. Chem. Eng. 2025, 51, 254–264. [Google Scholar]
  119. Agalya, K.; Vijayakumar, S.; Vidhya, E.; Prathipkumar, S.; Mythili, R.; Devanesan, S.; AlSalhi, M.S.; Kim, W. Fabrication of PVA/TiO2 Composites Via Green Synthesis and Assessment of their Photodegradation and Anti-Germ Capabilities. Waste Biomass Valorization 2024, 15, 6441–6451. [Google Scholar] [CrossRef]
  120. Hieu, N.H.; Hai, N.D.; Dat, N.M.; Nam, N.T.H.; Giang, N.T.H.; Cong, C.Q.; Huong, L.M. Chitosan-mediated coupling of MgFe2O4–TiO2 nanocomposites for efficient methylene blue photodegradation. ACS Appl. Nano Mater. 2024, 7, 6583–6595. [Google Scholar] [CrossRef]
  121. Kassas, A.; Dhaini, B.; Zahwa, I.; Zayyat, R.; Shaito, A.; Hussein, B.; Mouyane, M.; Bernard, J.; Houivet, D.; Toufaily, J. Sun-light-activated composite TiO2-FV-Mo materials for photodegradation of the organic pollutant methylene blue. Heliyon 2024, 10, e40489. [Google Scholar] [CrossRef]
  122. Onwubiko, V.; Matsushita, Y.; Elshehy, E.A.; El-Khouly, M.E. Facile synthesis of TiO2–carbon composite doped nitrogen for efficient photodegradation of noxious methylene blue dye. RSC Adv. 2024, 14, 34298–34310. [Google Scholar] [CrossRef]
  123. Fendi, K.; Bouzidi, N.; Boudraa, R.; Saidani, A.; Manseri, A.; Quesada, D.E.; Hai, T.N.; Bollinger, J.-C.; Salvestrini, S.; Kebir, M.; et al. Testing of kaolinite/TiO2 nanocomposites for methylene blue removal: Photodegradation and mechanism. Int. J. Chem. React. Eng. 2024, 22, 1493–1508. [Google Scholar] [CrossRef]
  124. Coba, A.J.O.; Briceño, S.; Vizuete, K.; Debut, A.; González, G. Diatomite with TiO2 nanoparticles for the photocatalytic deg-radation of methylene blue. Carbon Trends 2025, 19, 100488. [Google Scholar] [CrossRef]
  125. Razali, M.H.; Md Fauzi, M.A.F.; Mohd Azam, B.; Yusoff, M. g-C3N4/TiO2 nanocomposite photocatalyst for methylene blue photodegradation under visible light. Appl. Nanosci. 2022, 12, 3197–3206. [Google Scholar] [CrossRef]
  126. Faryad, S.; Azhar, U.; Tahir, M.B.; Ali, W.; Arif, M.; Sagir, M. Spinach-derived boron-doped g-C3N4/TiO2 composites for efficient photo-degradation of methylene blue dye. Chemosphere 2023, 320, 138002. [Google Scholar] [CrossRef] [PubMed]
  127. Pe III, J.A.; Mun, S.P.; Lee, M. TiO2-carbonized medium-density fiberboard for the photodegradation of methylene blue. Wood Sci. Technol. 2021, 55, 1109–1122. [Google Scholar] [CrossRef]
  128. Padwal, Y.; Panchang, R.; Fouad, H.; Terashima, C.; Chauhan, R.; Gosavi, S.W. Unveiling the mechanism of enhanced meth-ylene blue degradation using chromium-TiO2/carbon nanocomposite photocatalyst. J. Solid State Electrochem. 2023, 27, 3557–3568. [Google Scholar] [CrossRef]
  129. Yoo, H.; Kim, J.H. Photoactive TiO2/CuxO composite films for photocatalytic degradation of methylene blue pollutant mole-cules. Adv. Powder Technol. 2021, 32, 1287–1293. [Google Scholar] [CrossRef]
  130. Aziz, K.; Naz, A.; Manzoor, S.; Khan, M.I.; Shanableh, A.; Fernandez Garcia, J. Visible light photodegradation of glyphosate and methylene blue using defect-modified graphitic carbon nitride decorated with Ag/TiO2. Catalysts 2023, 13, 1087. [Google Scholar] [CrossRef]
  131. Liyanaarachchi, H.; Thambiliyagodage, C.; Liyanaarachchi, C.; Samarakoon, U. Efficient photocatalysis of Cu doped TiO2/g-C3N4 for the photodegradation of methylene blue. Arab. J. Chem. 2023, 16, 104749. [Google Scholar] [CrossRef]
  132. Rini, N.P.; Istiqomah, N.I.; Suharyadi, E. Enhancing photodegradation of methylene blue and reusability using CoO/ZnO composite nanoparticles. Case Stud. Chem. Environ. Eng. 2023, 7, 100301. [Google Scholar] [CrossRef]
  133. Ngullie, R.C.; Bhuvaneswari, K.; Shanmugam, P.; Boonyuen, S.; Smith, S.M.; Sathishkumar, M. Magnetically recoverable bi-omass-derived carbon-aerogel supported ZnO (ZnO/MNC) composites for the photodegradation of methylene blue. Catalysts 2022, 12, 1073. [Google Scholar] [CrossRef]
  134. Acedo-Mendoza, A.; Infantes-Molina, A.; Vargas-Hernández, D.; Chávez-Sánchez, C.; Rodríguez-Castellón, E.; Tánori-Córdova, J. Photodegradation of methylene blue and methyl orange with CuO supported on ZnO photocatalysts: The effect of copper loading and reaction temperature. Mater. Sci. Semicond. Process. 2020, 119, 105257. [Google Scholar] [CrossRef]
  135. Bekru, A.G.; Tufa, L.T.; Zelekew, O.A.; Goddati, M.; Lee, J.; Sabir, F.K. Green synthesis of a CuO–ZnO nanocomposite for efficient photodegradation of methylene blue and reduction of 4-nitrophenol. ACS Omega 2022, 7, 30908–30919. [Google Scholar] [CrossRef]
  136. Ngullie, R.C.; Alaswad, S.O.; Bhuvaneswari, K.; Shanmugam, P.; Pazhanivel, T.; Arunachalam, P. Synthesis and characteri-zation of efficient ZnO/g-C3N4 nanocomposites photocatalyst for photocatalytic degradation of methylene blue. Coatings 2020, 10, 500. [Google Scholar] [CrossRef]
  137. Elshypany, R.; Selim, H.; Zakaria, K.; Moustafa, A.H.; Sadeek, S.A.; Sharaa, S.; Raynaud, P.; Nada, A.A. Elaboration of Fe3O4/ZnO nanocomposite with highly performance photocatalytic activity for degradation methylene blue under visible light irradiation. Environ. Technol. Innov. 2021, 23, 101710. [Google Scholar] [CrossRef]
  138. Guo, Y.; Fu, X.; Liu, R.; Chu, M.; Tian, W. Efficient green photocatalyst of Ag/ZnO nanoparticles for methylene blue photo-degradation. J. Mater. Sci. Mater. Electron. 2022, 33, 1–13. [Google Scholar] [CrossRef]
  139. Noruozi, A.; Nezamzadeh-Ejhieh, A. Preparation, characterization, and investigation of the catalytic property of α-Fe2O3-ZnO nanoparticles in the photodegradation and mineralization of methylene blue. Chem. Phys. Lett. 2020, 752, 137587. [Google Scholar] [CrossRef]
  140. Sahoo, S.; Bhuyan, M.; kumar Sahu, A.; Alagarsamy, P.; Sahoo, D. Photodegradation of methylene blue by met-al-nanoparticles-modulated-graphene-based composites: An efficient way of sewage water management. Solid State Sci. 2023, 142, 107255. [Google Scholar] [CrossRef]
  141. Yang, J.; Yao, X.; Wu, H.; Li, D.; Liang, J.; Tang, X.; Zhu, Z. Fabrication of biocarbon modified BiOI heterostructures for enhanced organic pollutants photodegradation activity. J. Mater. Sci. Mater. Electron. 2025, 36, 1–12. [Google Scholar] [CrossRef]
  142. Saini, I.; Singh, V.; Hamad, S.; Ram, S. Recent development in bimetallic metal organic frameworks as photocatalytic material. Inorg. Chem. Commun. 2024, 160, 111897. [Google Scholar] [CrossRef]
  143. Mohammadnejad, M.; Nekoo, N.M.; Alizadeh, S.; Sadeghi, S.; Geranmayeh, S. Enhanced removal of organic dyes from aqueous solutions by new magnetic HKUST-1: Facile strategy for synthesis. Sci. Rep. 2023, 13, 17981. [Google Scholar] [CrossRef]
  144. Zhang, J.; Su, C.; Xie, X.; Liu, P.; Huq, M.E. Enhanced visible light photocatalytic degradation of dyes in aqueous solution activated by HKUST-1: Performance and mechanism. RSC Adv. 2020, 10, 37028–37034. [Google Scholar] [CrossRef]
  145. Gupta, H.; Saini, I.; Singh, V.; Singh, V.; Yarramaneni, S.; Grover, P. Fast decomposition of organic contaminant in wastewater using Zn and Mn bimetallic metal organic frameworks. Polyhedron 2024, 261, 117116. [Google Scholar] [CrossRef]
  146. Gupta, H.; Saini, I.; Singh, V.; Singh, V.; Mishra, B. Enhancing photocatalytic activity via formation of heterojunctions intro-duced through postmetalation of metal organic frameworks with silver ions. Phys. Scr. 2024, 99, 1059a1053. [Google Scholar] [CrossRef]
  147. Van Tran, T.; Nguyen, D.T.C.; Nguyen, T.T.; Le, H.T.; Van Nguyen, C.; Nguyen, T.D. Metal-organic framework HKUST-1-based Cu/Cu2O/CuO@ C porous composite: Rapid synthesis and uptake application in antibiotics remediation. J. Water Process Eng. 2020, 36, 101319. [Google Scholar] [CrossRef]
  148. Saffari, M.; Ghorbanloo, M.; Morsali, A.; Gan, Y.; Su, Y. Metal Organic Framework-Derived Bi2O3-ZnO/TiO2 Nanofibers Catalysts for Enhanced Photodegradation of Methylene Blue. Catal. Lett. 2025, 155, 132. [Google Scholar] [CrossRef]
  149. Singh, V.; Singh, V.; Thakur, O.; Kumar, A.; Yarramaneni, S. Study of Structural Changes Induced in Copper-based Metal Organic Framework during Photocatalytic Dye Degradation of Methylene Blue using Vibrational Spectroscopy. J. Alloys Compd. 2025, 1022, 180073. [Google Scholar]
  150. Zhou, J.; Zhou, Q.; Sun, H.; Li, X.; Chen, A.; Chen, J.; Chu, C. Selective detoxification of a sulfur mustard simulant in air by a methylene blue-functionalized metal–organic framework. Dalton Trans. 2025, 54, 1827–1837. [Google Scholar] [CrossRef]
  151. Wang, Y.; Wang, Z.; Luo, L.; Tawiah, B.; Liu, C.; Hua, J.; Ming, Y.; Xin, J.H.; Wong, W.-Y.; Fei, B. Water-Stable Mapbbr3@ Pbbroh Qds Confined by Metal–Organic Framework for Photodegradation of Wastewater. J. Colloid Interface Sci. 2025, 700, 138581. [Google Scholar] [CrossRef]
  152. Sakeerali, C.; Heo, S.M.; Kim, C.W. Preparation of core–shell structured Cu2O@ NH2-MIL-125 (Ti) MOF and efficient photo-catalytic degradation of methylene blue. Chem. Phys. 2025, 591, 112602. [Google Scholar] [CrossRef]
  153. Zheng, H.; Ma, F.; Ling, M.; Yang, Z.; Ji, R.; Jiang, Y.; Li, L.; Zhang, W. The multicomponent heterostructure flower spherical Ce-MOF/CdIn2S4/CdS for ciprofloxacin photodegradation via effective electron transfer. J. Solid State Chem. 2025, 344, 125210. [Google Scholar] [CrossRef]
  154. Liaqat, R.; Jamshaid, M.; Abo-Dief, H.M.; Ali, S.E.; El-Bahy, Z.M.; Fiaz, M.; Wattoo, M.A.; ur Rehman, A. Mo@ Ni-MOF nanocomposite: A promising photocatalyst for photodegradation of Methylene blue. J. Mol. Struct. 2024, 1315, 139011. [Google Scholar] [CrossRef]
  155. Adawiah, A.; Fitriani, S.; Andriyani, L.; Azizah, Y.N.; Wahyudin, W.; Sukandar, D.; Zulys, A. Optimization of methylene blue photodegradation by Cr-PTCHIna/TiO2 composite using box-behnken design application. Sustain. Chem. Clim. Action 2024, 5, 100047. [Google Scholar] [CrossRef]
  156. Khan, M.; Ahmed, M.M.; Akhtar, M.N.; Sajid, M.; Riaz, N.N.; Asif, M.; Kashif, M.; Shabbir, B.; Ahmad, K.; Saeed, M. Fabri-cation of CuWO4@ MIL-101 (Fe) nanocomposite for efficient OER and photodegradation of methylene blue. Heliyon 2024, 10, e40546. [Google Scholar] [CrossRef] [PubMed]
  157. Zulfa, L.L.; Hidayat, A.R.P.; Utomo, W.P.; Subagyo, R.; Kusumawati, E.N.; Kusumawati, Y.; Hartanto, D.; Widyastuti, W.; Ediati, R. Facile synthesis of Ni-ZIF-8 with improved photodegradation performance for methylene blue. Case Stud. Chem. Environ. Eng. 2024, 10, 100828. [Google Scholar] [CrossRef]
  158. Li, M.; Xiao, W.; Yin, Z.; Chen, Y.; Luo, Y.; Hong, Z.; Xue, M. Construction of a robust MOF-based superhydrophobic composite coating with the excellent performance in antifouling, drag reduction, and organic photodegradation. Prog. Org. Coat. 2024, 186, 108086. [Google Scholar] [CrossRef]
  159. Lin, S.; Zhang, T.; Dou, Z.; Yang, H. MnMg-MOF material photo-Fenton reaction degradation of methylene blue. Mater. Sci. Semicond. Process. 2024, 171, 108021. [Google Scholar] [CrossRef]
  160. Cheng, L.; Tang, Y.; Xie, M.; Sun, Y.; Liu, H. 2D ultrathin NiMOF decorated by Ti3C2 MXene for highly improved photocatalytic performance. J. Alloys Compd. 2021, 864, 158913. [Google Scholar] [CrossRef]
  161. Kashif, S.; Akram, S.; Murtaza, M.; Amjad, A.; Shah, S.S.A.; Waseem, A. Development of MOF-MXene composite for the removal of dyes and antibiotic. Diam. Relat. Mater. 2023, 136, 110023. [Google Scholar] [CrossRef]
  162. Zhang, W.; Chen, T.; Guo, P.; Zhang, W.; Yang, G. High removal of methyl blue over copper mixed-triazolate MOF by both adsorption and photodegradation from aqueous solution. Microporous Mesoporous Mater. 2024, 369, 113053. [Google Scholar] [CrossRef]
  163. Gu, C.; Weng, W.; Lu, C.; Tan, P.; Jiang, Y.; Zhang, Q.; Liu, X.; Sun, L. Decorating MXene with tiny ZIF-8 nanoparticles: An effective approach to construct composites for water pollutant removal. Chin. J. Chem. Eng. 2022, 42, 42–48. [Google Scholar] [CrossRef]
  164. Li, S.; Yang, S.; Liang, G.; Yan, M.; Wei, C.; Lu, Y. Regulation and photocatalytic degradation mechanism of a hydroxyl modified UiO-66 type metal organic framework. RSC Adv. 2023, 13, 5273–5282. [Google Scholar] [CrossRef] [PubMed]
  165. Jing, C.; Zhang, Y.; Zheng, J.; Ge, S.; Lin, J.; Pan, D.; Naik, N.; Guo, Z. In-situ constructing visible light CdS/Cd-MOF photo-catalyst with enhanced photodegradation of methylene blue. Particuology 2022, 69, 111–122. [Google Scholar] [CrossRef]
  166. Tang, T.; Li, C.; He, W.; Hong, W.; Zhu, H.; Liu, G.; Yu, Y.; Lei, C. Preparation of MOF-derived C-ZnO/PVDF composites membrane for the degradation of methylene blue under UV-light irradiation. J. Alloys Compd. 2022, 894, 162559. [Google Scholar] [CrossRef]
  167. Wang, Q.; Gao, Q.; Al-Enizi, A.M.; Nafady, A.; Ma, S. Recent advances in MOF-based photocatalysis: Environmental remedi-ation under visible light. Inorg. Chem. Front. 2020, 7, 300–339. [Google Scholar] [CrossRef]
  168. Brahmi, C.; Benltifa, M.; Vaulot, C.; Michelin, L.; Dumur, F.; Millange, F.; Frigoli, M.; Airoudj, A.; Morlet-Savary, F.; Bousselmi, L. New hybrid MOF/polymer composites for the photodegradation of organic dyes. Eur. Polym. J. 2021, 154, 110560. [Google Scholar] [CrossRef]
  169. Jatoi, Y.F.; Fiaz, M.; Athar, M. Synthesis of efficient TiO2/Al2O3@ Cu (BDC) composite for water splitting and photodegradation of methylene blue. J. Aust. Ceram. Soc. 2021, 57, 489–496. [Google Scholar] [CrossRef]
  170. El-Fawal, E.M.; Abd El Salam, H. Deposition of dyes on Cobalt-based metal-organic framework (Co-MOF) composites with promoted achievement photocatalytic degradation of an anionic dye (EBT) under visible light irradiation. Int. J. Environ. Anal. Chem. 2023, 103, 106–122. [Google Scholar] [CrossRef]
  171. Hariganesh, S.; Vadivel, S.; Maruthamani, D.; Kumaravel, M.; Paul, B.; Balasubramanian, N.; Vijayaraghavan, T. Facile large scale synthesis of CuCr2O4/CuO nanocomposite using MOF route for photocatalytic degradation of methylene blue and tet-racycline under visible light. Appl. Organomet. Chem. 2020, 34, e5365. [Google Scholar] [CrossRef]
  172. Si, Y.; Li, X.; Yang, G.; Mie, X.; Ge, L. Fabrication of a novel core–shell CQDs@ ZIF-8 composite with enhanced photocatalytic activity. J. Mater. Sci. 2020, 55, 13049–13061. [Google Scholar] [CrossRef]
  173. Saemian, T.; Sadr, M.H.; Yaraki, M.T.; Gharagozlou, M.; Soltani, B. Synthesis and characterization of CoFe2O4/SiO2/Cu-MOF for degradation of methylene blue through catalytic sono-Fenton-like reaction. Inorg. Chem. Commun. 2022, 138, 109305. [Google Scholar] [CrossRef]
  174. Zulfa, L.L.; Ediati, R.; Hidayat, A.R.P.; Subagyo, R.; Faaizatunnisa, N.; Kusumawati, Y.; Hartanto, D.; Widiastuti, N.; Utomo, W.P.; Santoso, M. Synergistic effect of modified pore and heterojunction of MOF-derived α-Fe2O3/ZnO for superior photo-catalytic degradation of methylene blue. RSC Adv. 2023, 13, 3818–3834. [Google Scholar] [CrossRef]
  175. Bouider, B.; Haffad, S.; Bouakaz, B.S.; Berd, M.; Ouhnia, S.; Habi, A. MOF-5/Graphene oxide composite photocatalyst for enhanced photocatalytic activity of methylene blue degradation under solar light. J. Inorg. Organomet. Polym. Mater. 2023, 33, 4001–4011. [Google Scholar] [CrossRef]
  176. Gholamrezapor, E.; Eslami, A. Photocatalytic degradation of methylene blue in aqueous solution by magnetic ZnMn2O4@ copper porphyrin nanocomposite. J. Iran. Chem. Soc. 2022, 19, 1071–1080. [Google Scholar] [CrossRef]
  177. Ziaadini, F.; Mostafavi, A.; Shamspur, T.; Fathirad, F. Photocatalytic degradation of methylene blue from aqueous solution using Fe3O4@ SiO2@ CeO2 core-shell magnetic nanostructure as an effective catalyst. Adv. Environ. Technol. 2019, 5, 127–132. [Google Scholar]
  178. Al-Sharabi, M.; Baiocco, D.; Lobel, B.T.; Cayre, O.J.; Zhang, Z.; Routh, A.F. Magnetic zinc oxide/silica microbeads for the photocatalytic degradation of azo dyes. Colloids Surf. A Physicochem. Eng. Asp. 2024, 695, 134169. [Google Scholar] [CrossRef]
  179. Kumar, A.P.; Ahmed, F.; Kumar, S.; Anuradha, G.; Harish, K.; Kumar, B.P. Synthesis of magnetically recoverable Ru/Fe3O4 nanocomposite for efficient photocatalytic degradation of methylene blue. J. Clust. Sci. 2022, 33, 853–865. [Google Scholar] [CrossRef]
  180. Light, S. Gadolinium Nanoparticle Decorated Multiwalled Carbon. Ph.D. Thesis, University of Johannesburg, Johannesburg, South Africa, 2012. [Google Scholar]
  181. Gangadhar, A.; Abhilash Mavinakere, R.; Jagadish, K.; Srikantaswamy, S. Photo-catalytic dye degradation of methylene blue by using ZrO2/MWCNT nanocomposites. Water Pract. Technol. 2021, 16, 1265–1276. [Google Scholar] [CrossRef]
  182. Arias, M.-C.; Aguilar, C.; Piza, M.; Zarazua, E.; Anguebes, F.; Anguebes, F.; Anguebes, F.; Anguebes, F.; Cordova, V. Removal of the methylene blue dye (MB) with catalysts of Au-TiO2: Kinetic and degradation pathway. Mod. Res. Catal. 2021, 10, 1–14. [Google Scholar] [CrossRef]
  183. Chong, M.N.; Jin, B.; Chow, C.; Sain, C. Recent developments in photocatalytic water treatment technology: A review. Water Res. 2010, 44, 2997–3027. [Google Scholar] [CrossRef]
  184. Iervolino, G.; Zammit, I.; Vaiano, V.; Rizzo, L. Limitations and prospects for wastewater treatment by UV and visible-light-active heterogeneous photocatalysis: A critical review. Heterog. Photocatal. Recent Adv. 2019, 378, 225–264. [Google Scholar]
  185. Younis, S.A.; Kim, K.-H. Heterogeneous photocatalysis scalability for environmental remediation: Opportunities and challenges. Catalysts 2020, 10, 1109. [Google Scholar] [CrossRef]
  186. Alalm, M.G.; Djellabi, R.; Meroni, D.; Pirola, C.; Bianchi, C.L.; Boffito, D.C. Toward scaling-up photocatalytic process for multiphase environmental applications. Catalysts 2021, 11, 562. [Google Scholar] [CrossRef]
  187. Kanakaraju, D.; Glass, B.D.; Oelgemöller, M. Advanced oxidation process-mediated removal of pharmaceuticals from water: A review. J. Environ. Manag. 2018, 219, 189–207. [Google Scholar] [CrossRef]
  188. Sudhaik, A.; Raizada, P.; Rangabhashiyam, S.; Singh, A.; Nguyen, V.-H.; Van Le, Q.; Khan, A.A.P.; Hu, C.; Huang, C.-W.; Ahamad, T.; et al. Copper sulfides based photocatalysts for degradation of environmental pollution hazards: A review on the recent catalyst design concepts and future perspectives. Surf. Interfaces 2022, 33, 102182. [Google Scholar] [CrossRef]
  189. Zhang, N.; Xing, Z.; Li, Z.; Zhou, W. Sulfur vacancy engineering of metal sulfide photocatalysts for solar energy conversion. Chem Catal. 2023, 3, 100375. [Google Scholar] [CrossRef]
  190. Kumari, H.; Sonia; Suman; Ranga, R.; Chahal, S.; Devi, S.; Sharma, S.; Kumar, S.; Kumar, P.; Kumar, S.; et al. A review on photocatalysis used for wastewater treatment: Dye degradation. Water Air Soil Pollut. 2023, 234, 349. [Google Scholar] [CrossRef]
  191. Ahuja, T.; Brighu, U.; Saxena, K. Recent advances in photocatalytic materials and their applications for treatment of wastewater: A review. J. Water Process Eng. 2023, 53, 103759. [Google Scholar] [CrossRef]
  192. Chalatsi-Diamanti, P.; Isari, E.A.; Grilla, E.; Kokkinos, P.; Kalavrouziotis, I.K. Recent prospects, challenges and advancements of photocatalysis as a wastewater treatment method. Water Emerg. Contam. Nanoplastics 2025, 4, 12. [Google Scholar] [CrossRef]
  193. Qudsieh, I.Y.; Ali, M.A.; Maafa, I.M. Effect of water matrix on photocatalytic degradation of organic pollutants in water: A literature review. Rev. Chem. Eng. 2025, 41, 539–573. [Google Scholar] [CrossRef]
  194. Adenuga, D.O.; Tichapondwa, S.M.; Chirwa, E.M. Influence of wastewater matrix on the visible light degradation of phenol using AgCl/Bi24O31Cl10 photocatalyst. Environ. Sci. Pollut. Res. 2023, 30, 98922–98933. [Google Scholar] [CrossRef] [PubMed]
  195. Azzaz, A.A.; Jellali, S.; Hamed, N.B.H.; El Jery, A.; Khezami, L.; Assadi, A.A.; Amrane, A. Photocatalytic treatment of wastewater containing simultaneous organic and inorganic pollution: Competition and operating parameters effects. Catalysts 2021, 11, 855. [Google Scholar] [CrossRef]
Figure 1. Illustration of the mechanism involved in photocatalytic degradation. When exposed to light irradiation, electrons (e) transition from the valence band (VB) to the conduction band (CB), resulting in the formation of holes (h+) in the VB. The photogenerated charge carriers engage in redox reactions, resulting in the formation of reactive oxygen species (ROS), including H2O2, OH, and h+. These species then facilitate the mineralization of organic pollutants into CO2 and water. The diagram illustrates the critical function of band gap excitation and charge separation in facilitating pollutant degradation. It highlights the need for more effective strategies to reduce electron–hole recombination and enhance the efficiency of reactive oxygen species (ROS) generation.
Figure 1. Illustration of the mechanism involved in photocatalytic degradation. When exposed to light irradiation, electrons (e) transition from the valence band (VB) to the conduction band (CB), resulting in the formation of holes (h+) in the VB. The photogenerated charge carriers engage in redox reactions, resulting in the formation of reactive oxygen species (ROS), including H2O2, OH, and h+. These species then facilitate the mineralization of organic pollutants into CO2 and water. The diagram illustrates the critical function of band gap excitation and charge separation in facilitating pollutant degradation. It highlights the need for more effective strategies to reduce electron–hole recombination and enhance the efficiency of reactive oxygen species (ROS) generation.
Catalysts 15 00893 g001
Figure 2. The schematic illustrates a semiconductor’s photocatalytic mechanism, most likely Ag/TiO2/CNT composites. It demonstrates how electrons are excited by light and subsequently separated from holes using carbon nanotubes (CNTs) and silver (Ag) nanoparticles. By preventing recombination, this charge separation increases the generation of reactive species that break down pollutants.
Figure 2. The schematic illustrates a semiconductor’s photocatalytic mechanism, most likely Ag/TiO2/CNT composites. It demonstrates how electrons are excited by light and subsequently separated from holes using carbon nanotubes (CNTs) and silver (Ag) nanoparticles. By preventing recombination, this charge separation increases the generation of reactive species that break down pollutants.
Catalysts 15 00893 g002
Figure 3. (a) MB photodegradation by the LCS composite at varying pH levels; and (b) MB photocatalytic degradation by the LCS composite in a cyclic manner. SEM images (ce) of the LCS composite. Reproduced with permission from [75]. Copyright 2025, Taylor & Francis.
Figure 3. (a) MB photodegradation by the LCS composite at varying pH levels; and (b) MB photocatalytic degradation by the LCS composite in a cyclic manner. SEM images (ce) of the LCS composite. Reproduced with permission from [75]. Copyright 2025, Taylor & Francis.
Catalysts 15 00893 g003
Figure 4. (a) Synthesis process for the YTCN photocatalyst [76]; and (b) schematic representation of the photocatalytic degradation mechanism of MB. Reproduced with permission from [77] (copyright Elsevier).
Figure 4. (a) Synthesis process for the YTCN photocatalyst [76]; and (b) schematic representation of the photocatalytic degradation mechanism of MB. Reproduced with permission from [77] (copyright Elsevier).
Catalysts 15 00893 g004
Figure 5. (A) Composite material adsorption capacity; (B) MB removal ratio in 30 min; (C) Langmuir–Hinshelwood model fitting curve; and (D) photocatalytic MB degradation cycling studies. Reproduced with permission from [78] (copyright Elsevier).
Figure 5. (A) Composite material adsorption capacity; (B) MB removal ratio in 30 min; (C) Langmuir–Hinshelwood model fitting curve; and (D) photocatalytic MB degradation cycling studies. Reproduced with permission from [78] (copyright Elsevier).
Catalysts 15 00893 g005
Figure 6. The degradation mechanism of MB dye exposed to UV and visible light using (a) TiO2−x and (b) RGO-TiO2−x. Reproduced with permission from [79] (copyright Elsevier).
Figure 6. The degradation mechanism of MB dye exposed to UV and visible light using (a) TiO2−x and (b) RGO-TiO2−x. Reproduced with permission from [79] (copyright Elsevier).
Catalysts 15 00893 g006
Figure 7. (a) Synthesis pathway of the ZnO@CBC composite. (b) Mechanism of MB degradation via photocatalysis using ZnO@CBC. Reproduced with permission from [80] (copyright Elsevier).
Figure 7. (a) Synthesis pathway of the ZnO@CBC composite. (b) Mechanism of MB degradation via photocatalysis using ZnO@CBC. Reproduced with permission from [80] (copyright Elsevier).
Catalysts 15 00893 g007
Figure 8. Illustrates the preparation of AC and the AC@Fe3O4 composite. Reproduced with permission from [81] (copyright Wiley-VCH).
Figure 8. Illustrates the preparation of AC and the AC@Fe3O4 composite. Reproduced with permission from [81] (copyright Wiley-VCH).
Catalysts 15 00893 g008
Figure 9. Preparation of porous activated carbon from jute sticks and the corresponding carboxylation process. Reproduced with permission from [83] (copyright Elsevier).
Figure 9. Preparation of porous activated carbon from jute sticks and the corresponding carboxylation process. Reproduced with permission from [83] (copyright Elsevier).
Catalysts 15 00893 g009
Figure 10. Proposed mechanism of MB degradation using MnO2-g-C3N4-ZnO. Reproduced with permission from [84] (copyright Wiley-VCH).
Figure 10. Proposed mechanism of MB degradation using MnO2-g-C3N4-ZnO. Reproduced with permission from [84] (copyright Wiley-VCH).
Catalysts 15 00893 g010
Figure 11. Three-dimensional response surface (right) and contour plots (left) showing the combined effects of (A) pH and oxidant dose, (B) MB concentration and oxidant dose, and (C) pH and dye concentration on degradation efficiency using the g-C3N4/GO/CuFe2O4 composite. Reproduced with permission from [86] (copyright SPRINGER Nature).
Figure 11. Three-dimensional response surface (right) and contour plots (left) showing the combined effects of (A) pH and oxidant dose, (B) MB concentration and oxidant dose, and (C) pH and dye concentration on degradation efficiency using the g-C3N4/GO/CuFe2O4 composite. Reproduced with permission from [86] (copyright SPRINGER Nature).
Catalysts 15 00893 g011
Figure 12. Photodegradation mechanism of MB by CA/ZnS-Ag composites irradiated by visible light. Reproduced with permission from [87] (copyright SPRINGER Nature).
Figure 12. Photodegradation mechanism of MB by CA/ZnS-Ag composites irradiated by visible light. Reproduced with permission from [87] (copyright SPRINGER Nature).
Catalysts 15 00893 g012
Figure 13. Proposed mechanism of charge separation and transfer during photocatalytic degradation and hydrogen evolution using S-gTAHP-15h. Reproduced with permission from [88] (copyright Elsevier).
Figure 13. Proposed mechanism of charge separation and transfer during photocatalytic degradation and hydrogen evolution using S-gTAHP-15h. Reproduced with permission from [88] (copyright Elsevier).
Catalysts 15 00893 g013
Figure 14. (a) Schematic of the synthesis, structure, and catalytic use of Cu-g-C3N4/BC composites. (b) Suggested photo-Fenton mechanism for dye degradation under visible light using Cu-g-C3N4/BC. Reproduced with permission from [90] (copyright Elsevier).
Figure 14. (a) Schematic of the synthesis, structure, and catalytic use of Cu-g-C3N4/BC composites. (b) Suggested photo-Fenton mechanism for dye degradation under visible light using Cu-g-C3N4/BC. Reproduced with permission from [90] (copyright Elsevier).
Catalysts 15 00893 g014
Figure 15. Cyclic stability and reproducibility of the WS2/GO/Au electrocatalyst for complete degradation of MB dye; superimposed UV-Vis spectra illustrating consistent electrocatalytic performance. Reproduced with permission from [91] (copyright Elsevier).
Figure 15. Cyclic stability and reproducibility of the WS2/GO/Au electrocatalyst for complete degradation of MB dye; superimposed UV-Vis spectra illustrating consistent electrocatalytic performance. Reproduced with permission from [91] (copyright Elsevier).
Catalysts 15 00893 g015
Figure 16. NiO/g-C3N4 Z-scheme heterojunction fabrication for improved MB dye photocatalytic degradation. Reproduced with permission from [92] (copyright Elsevier).
Figure 16. NiO/g-C3N4 Z-scheme heterojunction fabrication for improved MB dye photocatalytic degradation. Reproduced with permission from [92] (copyright Elsevier).
Catalysts 15 00893 g016
Figure 17. Proposed sono-photocatalytic mechanism for binary GO/Fe3O4 NCs. Reproduced with permission from [93] (copyright Elsevier).
Figure 17. Proposed sono-photocatalytic mechanism for binary GO/Fe3O4 NCs. Reproduced with permission from [93] (copyright Elsevier).
Catalysts 15 00893 g017
Figure 18. Nix/MoSx/MOF-2@g-C3N4 carbon nanostructures for efficient photocatalytic removal of MB. Reproduced with permission from [94] (copyright Elsevier).
Figure 18. Nix/MoSx/MOF-2@g-C3N4 carbon nanostructures for efficient photocatalytic removal of MB. Reproduced with permission from [94] (copyright Elsevier).
Catalysts 15 00893 g018
Figure 19. (a) Surface schematic of carbon-based nanostructures: MWCNT, GNR, and CNx. (b) MB dye degradation via photocatalysis involving OH, O2, and HO2 radicals. Reproduced with permission from [95] (copyright Elsevier).
Figure 19. (a) Surface schematic of carbon-based nanostructures: MWCNT, GNR, and CNx. (b) MB dye degradation via photocatalysis involving OH, O2, and HO2 radicals. Reproduced with permission from [95] (copyright Elsevier).
Catalysts 15 00893 g019
Figure 20. A schematic diagram of organic pollutant photocatalytic degradation by g-C3N4@WDC. Reproduced with permission from [98] (copyright Elsevier).
Figure 20. A schematic diagram of organic pollutant photocatalytic degradation by g-C3N4@WDC. Reproduced with permission from [98] (copyright Elsevier).
Catalysts 15 00893 g020
Figure 21. Plausible mechanism of MB degradation by N-CDs@ZnO under UV light via ROS generation. Reproduced with permission from [99] (copyright Royal Society of Chemistry).
Figure 21. Plausible mechanism of MB degradation by N-CDs@ZnO under UV light via ROS generation. Reproduced with permission from [99] (copyright Royal Society of Chemistry).
Catalysts 15 00893 g021
Figure 22. Schematic illustration of the enhanced photocatalytic mechanism of TiO2@CFs under light irradiation. Reproduced with permission from [100] (copyright SPRINGER Nature).
Figure 22. Schematic illustration of the enhanced photocatalytic mechanism of TiO2@CFs under light irradiation. Reproduced with permission from [100] (copyright SPRINGER Nature).
Catalysts 15 00893 g022
Figure 23. Schematic of material synthesis. Reproduced with permission from [109] (copyright Elsevier).
Figure 23. Schematic of material synthesis. Reproduced with permission from [109] (copyright Elsevier).
Catalysts 15 00893 g023
Figure 24. Schematic of the ZnCO3-Ag2CO3 band structure and possible mechanism of MB degradation. Reproduced with permission from [112] (copyright SPRINGER Nature Link).
Figure 24. Schematic of the ZnCO3-Ag2CO3 band structure and possible mechanism of MB degradation. Reproduced with permission from [112] (copyright SPRINGER Nature Link).
Catalysts 15 00893 g024
Figure 25. (a) Schematic illustration of the photocatalytic degradation mechanism of RhB and MB using PEDOT/Ag2SeO3 (3:1:1–4 mL) nanohybrids. Reproduced with permission from [114]. (b) Photocatalytic setup under UV and visible light irradiation. Reproduced with permission from [115] (copyright Wiley). (c) Proposed mechanism for MB photodegradation using MRGO 20 NPs nanocomposite. Reproduced with permission from [115] (copyright SPRINGER Nature Link).
Figure 25. (a) Schematic illustration of the photocatalytic degradation mechanism of RhB and MB using PEDOT/Ag2SeO3 (3:1:1–4 mL) nanohybrids. Reproduced with permission from [114]. (b) Photocatalytic setup under UV and visible light irradiation. Reproduced with permission from [115] (copyright Wiley). (c) Proposed mechanism for MB photodegradation using MRGO 20 NPs nanocomposite. Reproduced with permission from [115] (copyright SPRINGER Nature Link).
Catalysts 15 00893 g025
Figure 26. (a) Proposed visible-light-driven photocatalytic degradation mechanism of MB using FeSe2/ZnO composites. Reproduced with permission from [116]. (b) Schematic representation of the photocatalytic mechanism. Reproduced with permission from [117] (copyright Elsevier).
Figure 26. (a) Proposed visible-light-driven photocatalytic degradation mechanism of MB using FeSe2/ZnO composites. Reproduced with permission from [116]. (b) Schematic representation of the photocatalytic mechanism. Reproduced with permission from [117] (copyright Elsevier).
Catalysts 15 00893 g026
Figure 27. Preparation and photodegradation of MB. Reproduced with permission from [118] (copyright Elsevier).
Figure 27. Preparation and photodegradation of MB. Reproduced with permission from [118] (copyright Elsevier).
Catalysts 15 00893 g027
Figure 28. (a,b) Photodegradation efficiency of MB degradation by TiO2-C@N at varying initial concentrations (pH 6). (c) Reusability of TiO2-C@N in the dark. (d) Reusability under UV light. Reproduced with permission from [122] (copyright Royal Society of Chemistry).
Figure 28. (a,b) Photodegradation efficiency of MB degradation by TiO2-C@N at varying initial concentrations (pH 6). (c) Reusability of TiO2-C@N in the dark. (d) Reusability under UV light. Reproduced with permission from [122] (copyright Royal Society of Chemistry).
Catalysts 15 00893 g028
Figure 29. (a) Degradation rate [125], (b) mechanism of interphase charge transfer [125], and (c) nanospherical inter-layered of g-C3N4/TiO2 photocatalyst. Reproduced with permission from [125] (copyright SPRINGER Nature Link).
Figure 29. (a) Degradation rate [125], (b) mechanism of interphase charge transfer [125], and (c) nanospherical inter-layered of g-C3N4/TiO2 photocatalyst. Reproduced with permission from [125] (copyright SPRINGER Nature Link).
Catalysts 15 00893 g029
Figure 30. (a) Photodegradation of methylene blue (MB) showing the decrease in C/C0 with irradiation time [126], (b) kinetic analysis of MB degradation using the linear plot of ln(C0/C) [126], and (c) illustration of the synthesis route for spinach-derived boron-doped g-C3N4/TiO2 composites for enhanced MB degradation. Reproduced with permission from [126] (copyright Elsevier).
Figure 30. (a) Photodegradation of methylene blue (MB) showing the decrease in C/C0 with irradiation time [126], (b) kinetic analysis of MB degradation using the linear plot of ln(C0/C) [126], and (c) illustration of the synthesis route for spinach-derived boron-doped g-C3N4/TiO2 composites for enhanced MB degradation. Reproduced with permission from [126] (copyright Elsevier).
Catalysts 15 00893 g030
Figure 31. (a) The proposed photocatalytic mechanism of Cr-TiO2/carbon nanocomposite. Reproduced with permission from [128] (copyright SPRINGER Nature Link). (b) Schematic illustration of the photodegradation process of MB. Reproduced with permission from [129] (copyright Elsevier).
Figure 31. (a) The proposed photocatalytic mechanism of Cr-TiO2/carbon nanocomposite. Reproduced with permission from [128] (copyright SPRINGER Nature Link). (b) Schematic illustration of the photodegradation process of MB. Reproduced with permission from [129] (copyright Elsevier).
Catalysts 15 00893 g031
Figure 32. Band alignment of (a) P25 TiO2 [131], (b) Cu-TiO2 [131], and (c) Cu-TiO2/g-C3N4. Reproduced with permission from [131] (copyright Elsevier).
Figure 32. Band alignment of (a) P25 TiO2 [131], (b) Cu-TiO2 [131], and (c) Cu-TiO2/g-C3N4. Reproduced with permission from [131] (copyright Elsevier).
Catalysts 15 00893 g032
Figure 33. (a) The possible photocatalytic mechanism for the degradation of pollutants by nanoparticles. Reproduced with permission from [130]. (b) Z-scheme showing the charge separation and band alignment in CoO/ZnO NPs with low H2O2 for MB degradation. Reproduced with permission from [132] (copyright Elsevier).
Figure 33. (a) The possible photocatalytic mechanism for the degradation of pollutants by nanoparticles. Reproduced with permission from [130]. (b) Z-scheme showing the charge separation and band alignment in CoO/ZnO NPs with low H2O2 for MB degradation. Reproduced with permission from [132] (copyright Elsevier).
Catalysts 15 00893 g033
Figure 34. Proposed Z-scheme for enhanced MB degradation by g-CN/ZnO under visible light. Reproduced with permission from [136] (copyright Multidisciplinary Digital Publishing Institute).
Figure 34. Proposed Z-scheme for enhanced MB degradation by g-CN/ZnO under visible light. Reproduced with permission from [136] (copyright Multidisciplinary Digital Publishing Institute).
Catalysts 15 00893 g034
Figure 35. Schematic illustration of the photocatalytic mechanism of Ag–ZnO nanocomposites under solar light irradiation. (a) Charge transfer and reactive oxygen species (ROS) generation via electron–hole separation and trapping at the Ag surface. (b) Surface plasmon resonance (SPR) effect of Ag nanoparticles enhancing electron excitation, charge separation, and subsequent degradation of methylene blue (MB) into CO2, H2O, and other products. Reproduced with permission from [138] (copyright SPRINGER Nature Link).
Figure 35. Schematic illustration of the photocatalytic mechanism of Ag–ZnO nanocomposites under solar light irradiation. (a) Charge transfer and reactive oxygen species (ROS) generation via electron–hole separation and trapping at the Ag surface. (b) Surface plasmon resonance (SPR) effect of Ag nanoparticles enhancing electron excitation, charge separation, and subsequent degradation of methylene blue (MB) into CO2, H2O, and other products. Reproduced with permission from [138] (copyright SPRINGER Nature Link).
Catalysts 15 00893 g035
Figure 36. Charge transfer pathways and MB degradation mechanism of BiOI/C composite. Reproduced with permission from [141] (copyright SPRINGER Nature Link).
Figure 36. Charge transfer pathways and MB degradation mechanism of BiOI/C composite. Reproduced with permission from [141] (copyright SPRINGER Nature Link).
Catalysts 15 00893 g036
Figure 38. Photocatalytic degradation curves of MB under visible light are presented as follows: (a) for TC, (b) for T C@NH2-MIL-125(Ti), (c) a comparison of the photocatalytic activity of the samples, (d) the kinetic curve representing ln(C0/C) as a function of time, and (e) the synthesis of flower-shaped Ce-MOF/CdIn2S4/CdS, its electron transfer, and outstanding photocatalytic performance in the degradation of MB. Reproduced with permission from [152] (copyright 2025, Elsevier).
Figure 38. Photocatalytic degradation curves of MB under visible light are presented as follows: (a) for TC, (b) for T C@NH2-MIL-125(Ti), (c) a comparison of the photocatalytic activity of the samples, (d) the kinetic curve representing ln(C0/C) as a function of time, and (e) the synthesis of flower-shaped Ce-MOF/CdIn2S4/CdS, its electron transfer, and outstanding photocatalytic performance in the degradation of MB. Reproduced with permission from [152] (copyright 2025, Elsevier).
Catalysts 15 00893 g038
Figure 39. (a) Possible mechanism for adsorption behavior of Ni(20)-ZIF-8 [158], and (b) illustrates ZIF-8/POTS coatings preparation through simple chemical modification and spraying technology, including the self-cleaning and photocatalytic properties. Reproduced with permission from [158] (copyright Elsevier).
Figure 39. (a) Possible mechanism for adsorption behavior of Ni(20)-ZIF-8 [158], and (b) illustrates ZIF-8/POTS coatings preparation through simple chemical modification and spraying technology, including the self-cleaning and photocatalytic properties. Reproduced with permission from [158] (copyright Elsevier).
Catalysts 15 00893 g039
Figure 40. (a) Mechanism for photocatalytic performance enhancement of MnMg-MOF [159], (b) UNiMOF/Ti3C2 [159], (c) ZIF-8/Ti3C2Tx [163], and (d) adsorption curves of MB on Ti3C2Tx, ZIF-8, and ZIF-8/Ti3C2Tx. Reproduced with permission from [163] (copyright Elsevier).
Figure 40. (a) Mechanism for photocatalytic performance enhancement of MnMg-MOF [159], (b) UNiMOF/Ti3C2 [159], (c) ZIF-8/Ti3C2Tx [163], and (d) adsorption curves of MB on Ti3C2Tx, ZIF-8, and ZIF-8/Ti3C2Tx. Reproduced with permission from [163] (copyright Elsevier).
Catalysts 15 00893 g040
Figure 42. (a) Proposed mechanism for MB degradation using ZIF-8 decomposed with NDCQDs. (b) Pseudo-second-order kinetics model for MB degradation. (c) Intra-particle diffusion mechanism of CoFe2O4/SiO2/Cu-MOF [172]. (d) Reusability of CoFe2O4/SiO2/Cu-MOF in MB degradation. Reproduced with permission from [173] (copyright Elsevier).
Figure 42. (a) Proposed mechanism for MB degradation using ZIF-8 decomposed with NDCQDs. (b) Pseudo-second-order kinetics model for MB degradation. (c) Intra-particle diffusion mechanism of CoFe2O4/SiO2/Cu-MOF [172]. (d) Reusability of CoFe2O4/SiO2/Cu-MOF in MB degradation. Reproduced with permission from [173] (copyright Elsevier).
Catalysts 15 00893 g042
Figure 43. Schematic illustration of the preparation of MOF-5, GO, and MOF-5/GO. Reproduced with permission from [175] (copyright Elsevier).
Figure 43. Schematic illustration of the preparation of MOF-5, GO, and MOF-5/GO. Reproduced with permission from [175] (copyright Elsevier).
Catalysts 15 00893 g043
Table 1. Comparative analysis of photocatalysis and other treatment techniques.
Table 1. Comparative analysis of photocatalysis and other treatment techniques.
Sr. No.Treatment MethodsDescriptionAdvantagesLimitationsRefs.
1PhotocatalysisUtilizes light-activated catalysts.Environmentally friendly, cost-effective, high energy saving, minimal secondary pollutants, and the easiest catalyst loading.Difficulty in regeneration and recovery.[29]
2BiologicalUses microorganisms to degrade the material.Cost-effective and excellent reduction of color.Not applicable to highly concentrated organic waste, difficult to control, and has low efficiency in dye degradation.[30]
3Chemical precipitationUse of various chemicals, such as lime or aluminum sulfate, to convert dissolved material into a solid substance.Energy consumption is low and effective in removing organic halogens.Excessive use of chemicals (such as lime, oxidants, and H2S) and the development of sludge.[31]
4Membrane filtrationUse of a semi-permeable membrane to eliminate the contaminants.Rapid and eco-friendly method.High operational cost and membrane fouling reduce efficiency.[32]
5AdsorptionPollutants adhere to porous materials.Simple and cost-effective.The generation of secondary waste and adsorbent saturation requires regeneration.[33]
6Coagulation/flocculation Use of coagulants such as alum, potash alum, and polyaluminum chloride for the removal of materials.Simple and good for removing pollutants.Requires a high dose of chemicals, the formation of massive sludge, and larger particles.[34,35]
Table 2. Summary of carbon-based composite photocatalysts for the degradation of MB.
Table 2. Summary of carbon-based composite photocatalysts for the degradation of MB.
PhotocatalystsSynthesis MethodBand GapEfficiencyRef.
GO/PAN/CQDHydrothermal process1.79 eV100%[45]
CdS/CQDs/g-C3N4Calcination process2.68 eV86%[69]
MWCNTs/TiO2Liquid phase deposition method2.8–2.95 eV90%[70]
Ag/TiO2/CNTSonochemical method3.2 eV98%[72]
Lignin-based carbon/cadmium sulfide compositeIn situ precipitation method2.38 eV91.7%[75]
Yb-TiO2/g-C3N5Hydrothermal process2.77 eV96.57%[76]
PcDNPIMCAdler–Longo method-95.5%[77]
TiO2/C-550Sol-gel method2.7 eV100%[78]
RGO-TiO2−xHydrothermal process1.8 eV-[79]
ZnO@CBCImpregnation–Pyrolysis method~3.20 eV99.6%[80]
AC@Fe3O4Calcination and coprecipitation94.6%~1.7 to 2.0 eV[81]
ZnO/g-C3N4In situ synthesis97%3.02 to 2.94 eV[82]
ZnO/JSAC-COOHydrothermal method97.56%-[83]
MnO2-ZnO-g-C3N4Sol-gel method94%2.0 to 2.5 eV[84]
Ag2S/BSCN(SILAR) method96.48%~2.1–2.3 eV[85]
g-C3N4/GO/CuFe2O4In situ hydrothermal method99%2.31 eV[86]
CA/ZnS-AgPrecipitation method68.39%2.68 eV[87]
S-scheme/gC3N4/TiO2/SiO2/PANElectrospinning, calcination, hydrothermal, and freeze-drying techniques99.43%2.71 eV[88]
AC-ZrO2/CeO2 NCsCo-precipitation97.91%2.2–3.0 eV[89]
Cu-g-C3N4/BCIn situ pyrolysis32.7%2.06–2.24 eV[90]
WS2/GO/AuHydrothermal and laser ablation99.00%-[91]
NiO-doped C3N4Ultrasonic method92%2.95 eV[92]
g-C3N4/WS2One-pot hydrothermal method95.5%∼2.7 eV[101]
NiO/Cd/g-C3N4Microwave-assisted81.8%-[102]
NiO/ZnO/g-C3N4Hydrothermal method∼95%~2.68 eV[103]
GO/Fe3O4Chemical precipitation method98.68%1.96 eV[93]
g-C3N4/ZnOHydrothermal method97.7%2.91 eV[104]
Ni/Mo.S2/MOF-2@g-C3N4Solvothermal method91%1.75 eV[94]
Eggshell-based activated carbonChemical activation method83%-[96]
TiO2/COne-pot liquid phase reaction25.1%-[97]
g-C3N4@wood-derived carbonCarbonization98%-[98]
Bio-CDs co-doped with S/ClHydrothermal process94.2-[105]
CDs co-doped with N/SHydrothermal process2.34 > 100%-[106]
N-CDs@ZnOWet-impregnation method99%2.97 eV[99]
TiO2@CFsThermo-treatment after electrospinning94.54%-[100]
ZCF@MB-MIPChemical precipitation95.8%3.37 eV[107]
10%CCO/CeO2-nanocompositeCo-precipitation method followed by calcination85%2.75 eV[108]
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

Majeed, A.; Iqbal, M.A.; Do, T.-O. Advances in Composite Photocatalysts for Efficient Degradation of Organic Pollutants: Strategies, Challenges, and Future Perspectives. Catalysts 2025, 15, 893. https://doi.org/10.3390/catal15090893

AMA Style

Majeed A, Iqbal MA, Do T-O. Advances in Composite Photocatalysts for Efficient Degradation of Organic Pollutants: Strategies, Challenges, and Future Perspectives. Catalysts. 2025; 15(9):893. https://doi.org/10.3390/catal15090893

Chicago/Turabian Style

Majeed, Adnan, Muhammad Adnan Iqbal, and Trong-On Do. 2025. "Advances in Composite Photocatalysts for Efficient Degradation of Organic Pollutants: Strategies, Challenges, and Future Perspectives" Catalysts 15, no. 9: 893. https://doi.org/10.3390/catal15090893

APA Style

Majeed, A., Iqbal, M. A., & Do, T.-O. (2025). Advances in Composite Photocatalysts for Efficient Degradation of Organic Pollutants: Strategies, Challenges, and Future Perspectives. Catalysts, 15(9), 893. https://doi.org/10.3390/catal15090893

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