Next Article in Journal
Single and Combined Effects of Aged Polyethylene Microplastics and Cadmium on Nitrogen Species in Stormwater Filtration Systems: Perspectives from Treatment Efficiency, Key Microbial Communities, and Nitrogen Cycling Functional Genes
Next Article in Special Issue
A Review on Multifunctional Polymer–MXene Hybrid Materials for Electronic Applications
Previous Article in Journal
Correction: Iqbal et al. Floating ZnO QDs-Modified TiO2/LLDPE Hybrid Polymer Film for the Effective Photodegradation of Tetracycline under Fluorescent Light Irradiation: Synthesis and Characterisation. Molecules 2021, 26, 2509
Previous Article in Special Issue
Sustainable MXene Synthesis via Molten Salt Method and Nano-Silicon Coating for Enhanced Lithium-Ion Battery Performance
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Ti3C2Tx MXene-Based Hybrid Photocatalysts in Organic Dye Degradation: A Review

1
Department of Chemistry and Biochemistry, North Carolina Central University, Durham, NC 27707, USA
2
Department of Physical and Applied Sciences, University of Houston-Clear Lake, Houston, TX 77058, USA
3
Department of Chemistry and Physics, McNeese State University, Lake Charles, LA 70605, USA
*
Authors to whom correspondence should be addressed.
Molecules 2025, 30(7), 1463; https://doi.org/10.3390/molecules30071463
Submission received: 14 February 2025 / Revised: 16 March 2025 / Accepted: 22 March 2025 / Published: 26 March 2025
(This article belongs to the Special Issue The Way Forward in MXenes Materials)

Abstract

:
This review provides an overview of the fabrication methods for Ti3C2Tx MXene-based hybrid photocatalysts and evaluates their role in degrading organic dye pollutants. Ti3C2Tx MXene has emerged as a promising material for hybrid photocatalysts due to its high metallic conductivity, excellent hydrophilicity, strong molecular adsorption, and efficient charge transfer. These properties facilitate faster charge separation and minimize electron–hole recombination, leading to exceptional photodegradation performance, long-term stability, and significant attention in dye degradation applications. Ti3C2Tx MXene-based hybrid photocatalysts significantly improve dye degradation efficiency, as evidenced by higher percentage degradation and reduced degradation time compared to conventional semiconducting materials. This review also highlights computational techniques employed to assess and enhance the performance of Ti3C2Tx MXene-based hybrid photocatalysts for dye degradation. It identifies the challenges associated with Ti3C2Tx MXene-based hybrid photocatalyst research and proposes potential solutions, outlining future research directions to address these obstacles effectively.

1. Introduction

Rapid industrialization has escalated the production of toxic and carcinogenic wastewater [1]. The discharge of untreated wastewater from textile, pharmaceutical, dyeing, and paper printing industries significantly increases dye concentrations in water sources, posing a severe environmental threat [2]. These wastewaters contain chromophores and auxochromes, which color water, block sunlight, and disrupt photosynthesis, thereby harming aquatic ecosystems [3]. Discharged dyes are classified as cationic (e.g., methylene blue (MB), crystal violet (CV), rhodamine B (RhB)), which carry positive charges, are prevalent in textiles, and present toxicity risks, or anionic (e.g., acid orange 7 (AO7), acid blue 92 (AB92), Congo red (CR)), which carry negative charges, are water soluble, and are difficult to remove [4]. Various degradation methods, including membrane filtration, oxidation, adsorption, and photocatalysis, have been successfully developed and applied to address this issue [5,6].
Photocatalysis stands out as a highly effective degradation method due to its eco-friendliness and cost-effectiveness in treating organic pollutants [7]. Consequently, extensive research has focused on developing photocatalytic materials with wide energy bandgaps, abundant surface-active sites, and efficient charge separation capabilities to optimize dye photodegradation [8]. While single-component semiconducting photocatalysts like TiO2, ZnO, ZnS, SnO2, graphitic carbon nitride (g-C3N4), and CdS have been widely employed [9,10], their photocatalytic activity is often limited by poor visible light absorption, rapid charge recombination, and low electron–hole mobility [11,12]. To address these limitations and enhance photocatalytic performance, various strategies have been explored, including heterojunction construction [13], element doping [14,15], and co-catalyst introduction [14].
Modifying semiconducting materials with a co-catalyst is an effective strategy to enhance their photocatalytic activity [16]. A co-catalyst facilitates charge separation by preventing electron–hole recombination, a common limitation in photocatalysis. Materials such as graphene [17], carbon nanotubes [18], and carbon quantum dots [19] have been utilized as co-catalysts to improve dye photodegradation. While significant progress has been achieved using these co-catalysts, they each have certain limitations. Graphene and carbon nanotubes face challenges due to their scarcity of functional groups [20], while carbon quantum dots are hindered by their relatively low conductivity and scalability issues [21]. Challenges such as insufficient surface functional groups, which restrict chemical bonding between the photocatalyst and co-catalyst, and lower electrical conductivity, which impedes charge migration, can hinder photocatalytic efficiency. Consequently, there remains a need for efficient and cost-effective co-catalysts to optimize the photocatalytic degradation of organic dyes.
MXenes, a class of two-dimensional (2D) layered materials composed of transition metal carbides, nitrides, or carbonitrides with the general formula Mn+1Xn (where M is a transition metal like Sc, Ti, Zr, Hf, V, Nb, Ta, or Mo, and X is carbon, nitrogen, or carbonitride), are promising for hybrid photocatalyst fabrication [22,23,24]. Their lower Fermi level compared to most studied semiconductors, abundant terminal functional groups, excellent metallic conductivity, and exposed terminal metal sites contribute to this potential. Specifically, the presence of abundant functional groups such as -OH, -O, and -F on Ti3C2Tx MXene facilitates strong interfacial chemical bonding with semiconductors, enabling the formation of a Schottky junction. This junction acts as an electron trap, effectively suppressing photoexcited electron–hole pair recombinations [25]. Furthermore, the excellent metallic conductivity of Ti3C2Tx MXene ensures rapid charge carrier migration, promoting efficient separation of photogenerated electrons and holes [26]. Lastly, the exposed terminal metal sites enhance reactivity compared to carbon-based materials, making Ti3C2Tx MXene a highly effective co-catalyst.
While Ti3C2Tx MXene exhibits limited standalone photocatalytic activity, requiring doping and UV radiation for optimal function [27], it plays a crucial role in composite photocatalysts. Beyond acting as a photogenerated charge acceptor co-catalyst, its large surface area and surface functionalities provide an excellent platform for the uniform growth, size control, and fine dispersion of photocatalysts, thereby exposing more surface-active sites [22,28,29,30]. Moreover, Ti3C2Tx MXene possesses distinctive physical and chemical properties, including high electrical conductivity, excellent hydrophilicity, mechanical stability, ion intercalation ability, and tunable surface functionalization [31,32,33]. These attributes render it an ideal component for composite photocatalyst materials. The integration of Ti3C2Tx MXene with semiconducting materials has yielded hybrid photocatalysts with micro/nano architectures and multi-junction nanocomposites. These hybrids leverage synergistic interactions between Ti3C2Tx MXene and conventional semiconductors or metal nanostructures, leading to enhanced charge separation and reduced recombination rates [23,34]. Consequently, Ti3C2Tx MXene-based hybrid photocatalysts demonstrate high effectiveness for the photodegradation of organic pollutants [35].
This review comprehensively explores Ti3C2Tx MXene-based hybrid photocatalysts for the degradation of cationic and anionic organic dyes. It details the synthesis of Ti3C2Tx MXene, its integration into hybrid photocatalysts, its role and effectiveness in dye degradation, and the underlying mechanisms. Furthermore, it discusses current challenges and emerging opportunities, providing insights into potential future research pathways.

2. Ti3C2Tx MXene and Synthesis Methods

2.1. Introduction to Ti3C2Tx MXenes

MXenes, first synthesized by Naguib et al. in 2011, are 2D crystals of transition metal carbides, nitrides, or carbonitrides [36]. These 2D materials are produced via top-down approaches, starting with their three-dimensional parent materials, MAX phases. MAX phases, represented by the general formula M(n+1)AXn (where n = 1, 2, or 3), consist of an early transition metal M (e.g., titanium, niobium, molybdenum, tantalum, vanadium, chromium), an A-group element A (primarily groups 13 and 14), and carbon, nitrogen, or carbonitride X layers [37]. Chemical etching removes the A layers from the MAX phase, resulting in the formation of 2D MXene crystals. Due to the predominantly aqueous synthesis methods, the M elements on the resulting MXene surfaces are terminated with functional groups, represented by Tx. These terminated MXenes are denoted by the general formula M(n+1)XnTx (where n = 1, 2, or 3), with Tx typically representing -OH, -O, or -F surface terminations [38,39,40,41,42].
MXenes have garnered significant attention due to their exceptional properties, including hydrophilicity, high electrical conductivity, and tunable bandgaps through surface termination modifications [42,43]. Ti3C2Tx, the inaugural member of the MXene family, remains the most extensively researched due to its durability, the relatively simple process of etching aluminum layers from its MAX phase precursor, and its remarkable physical and chemical characteristics [36,44,45]. Consequently, leveraging their unique properties, numerous other MXene species have been explored for a wide range of applications.

2.2. Synthesis of Ti3C2Tx MXenes

Hydrofluoric acid (HF) etching was the pioneering technique for transforming MAX phases into MXenes [36]. Despite its simplicity and ability to yield high-quality MXenes [33], making it a widely used method, HF’s extreme toxicity and corrosiveness pose significant drawbacks [46]. To mitigate these risks, alternative methods, such as in situ HF generation via a reaction between LiF and HCl, have been developed. This approach also facilitates Li+ cation intercalation between MXene layers [47]. A characteristic outcome of HF etching is the formation of MXene layers with surface terminations, primarily -O, -OH, and -F, which enable surface engineering and bandgap tuning through modification of the termination type and quantity [48]. Another modified etching technique is the molten salt method, which involves reacting MAX phases with a Lewis acidic molten salt at elevated temperatures [39]. Similar to HF etching, an intercalant is utilized; however, the molten salt method allows for greater control over surface terminations during synthesis. For instance, -F terminations produced by HF etching can degrade the electrochemical performance of Ti3C2Tx electrodes in supercapacitors. To address this, Guo et al. [49] employed a one-step LiCl-KCl-K2CO3 molten salt etch and delamination process to replace -F terminations with -O, reducing the -F content from 11.23 to 3.43 atomic percent. This resulted in improved specific capacity, capacitance retention, conductivity, and electrochemical activity-specific surface area.
The hydrothermal etching method was developed as a safer and more environmentally friendly alternative to the highly corrosive and hazardous HF etching process. This technique enables efficient exfoliation and the production of high-quality MXene flakes, offering the advantages of larger interlayer spacing and improved delamination properties. Hydrothermal etching utilizes high-pressure and high-temperature conditions in an aqueous MXene solution to produce high-purity multilayer MXenes [46]. For example, Peng et al. [33] synthesized Ti3C2Tx MXenes using this technique by reacting MAX phases with HCl and HCl + NaBF4 solutions, followed by heating in an autoclave. The resulting MXenes were delaminated using dimethyl sulfoxide (DMSO) and sonication. XRD analysis revealed that hydrothermal etching was significantly more effective at removing aluminum compared to traditional HF etching. Furthermore, dye adsorption studies with cationic MB and anionic methyl orange (MO) demonstrated that hydrothermally etched Ti3C2Tx exhibited superior adsorption performance, with lower residual dye concentrations compared to non-hydrothermal counterparts. Some hydrothermal methods incorporate microwaves to excite reagents and reduce reaction temperature and time [50].
The resulting flakes were cleaned, delaminated using tetramethylammonium hydroxide, cleaned again, and then incorporated into a composite with reduced graphene oxide, which effectively degraded dyes via photolysis [51]. Cao et al. [50] utilized microwave hydrothermal etching to rapidly create and oxidize Ti3C2Tx MXene nanosheets. After cleaning, this Ti3C2Tx was combined with TiO2 and CdZnS, again using microwave hydrothermal methods, to synthesize a synergistic photocatalytic semiconductive heterojunction. This heterojunction degraded RhB dye molecules by 29.33% within 90 min, a 31.17-fold improvement compared to single Ti3C2 [52]. To drastically reduce etching time (from approximately 2–3 days to 45 min) and eliminate -F terminations, Latif et al. [53] applied 5 to 30 M concentrations of NaOH etchant to Ti3AlC2 in a microwave hydrothermal reaction. Higher NaOH concentrations resulted in synthesized Ti3C2Tx MXene, with a 0.46% aluminum content, as determined by XRD, and a semiconductive bandgap energy of 1.30 to 1.60 eV.

3. Design and Fabrication of Ti3C2Tx MXene-Based Hybrid Photocatalysts

A wide range of methods have been employed to fabricate Ti3C2Tx MXene-based hybrid photocatalysts, involving the integration of Ti3C2Tx MXene nanosheets with diverse nanomaterials. As a result of their chemical synthesis, Ti3C2Tx MXene materials naturally acquire surface terminations, and their tunable surface chemistry provides potential pathways for hybridization with various material classes through covalent and non-covalent interactions [54]. The redox behavior of titanium (Ti) has been leveraged to develop TiO2-based semiconductors on the MXene surface, where MXene serves as a growth platform. The presence of hydrophilic surface functional groups such as -O-, -OH, and -F is theoretically expected to enhance chemical bonding with other semiconductor photocatalysts. However, the development of hybrid photocatalysts utilizing the MXene functional terminations through covalent bonding requires state-of-the-art approaches, and hence, most studies have primarily utilized non-covalent approaches, where charge interactions or hydrophilic groups act as adsorption or anchoring sites for metal salts, semiconductors, and other materials [55,56].
Tran et al. [57] fabricated TiO2/Ti3C2Tx composites through an in situ partial oxidation of Ti3C2Tx, resulting in microscale safflower-like structures composed of TiO2/Ti3C2Tx heterostructure nanorods. This transformation involved sequential hydrothermal oxidation, alkalization, ion exchange, and heat treatment, during which layered MXene flakes were fragmented into nanoparticles, from which TiO2/Ti3C2Tx nanorods grew radially. A schematic of this synthesis process is shown in Figure 1.
The resulting TiO2/Ti3C2Tx heterostructures demonstrated exceptional photocatalytic properties. Their photocurrent was ten times greater than that of pristine MXene. Furthermore, the photocatalytic degradation efficiency of rhodamine B (RhB) reached 95%, a significant improvement over the 19% achieved by MXene alone. Even after four cycles, the degradation efficiency remained above 95%, indicating excellent stability. This superior performance was attributed to the rapid generation of TiO2 carriers, suppressed charge carrier recombination, and enhanced light absorption due to the porous safflower-like structure. Similarly, Quyen et al. [58] employed a novel synthesis approach to create TiO2@Ti3C2Tx nanoflowers with a porous 3D framework derived from 2D Ti3C2Tx MXenes via hydrothermal oxidation combined with calcination. This in situ transformation converted the initially conductive Ti3C2Tx MXene into a semiconductor, forming a TiO2@Ti3C2Tx heterojunction. Compared to the 36% degradation rate of pure TiO2 for RhB, the TiO2@Ti3C2Tx composite achieved an impressive 97% degradation within 40 min of light irradiation. This enhancement was attributed to the electronic structure of Ti3C2Tx, whose Fermi level is lower than that of TiO2. Upon light excitation, photoinduced electrons transferred from the valance (CB) of TiO2 to metallic Ti3C2Tx, leaving holes in the valence band (VB) of TiO2. The resulting Schottky barrier at the interface suppressed electron diffusion back into TiO2, thereby reducing electron–hole recombination and improving photocatalytic efficiency. In a similar vein, an in situ solvothermal process was utilized to prepare facet-exposed TiO2/Ti3C2Tx [59]. Since the exposed crystal planes of TiO2 can effectively capture photogenerated holes and facilitate rapid migration to the Ti3C2Tx surface, the photocatalytic performance towards methyl orange was significantly superior compared to TiO2 and Ti3C2Tx alone.
Calcination, a simple method for preparing Ti3C2Tx MXene composites, involves heating powders of two materials at elevated temperatures. As illustrated in Figure 2, varying amounts of Ti3C2Tx MXene powders were thoroughly dispersed in a Zn2+-containing solution [60]. The resulting mixture was heated in an oven to evaporate the solvent. Subsequently, the powder was calcined at 550 °C for 4 h, with a heating rate of 5 °C/min in ambient air to synthesize ZnO/Ti3C2Tx hybrid structures. Compared to pristine ZnO, the hybrid structure exhibited reduced photoluminescence intensity, an enhanced Brunauer–Emmett–Teller (BET) surface area, and improved photocatalytic degradation efficiency for MO and RhB.
The wet impregnation method, where one material is deposited onto a solid support, can be used to prepare Ti3C2Tx MXene composites. Nasri et al. [6] mixed g-C3N4 and Ti3C2Tx MXene powders in varying weight proportions and sonicated the mixture until a slurry formed. The resulting slurry was dried overnight at 60 °C to obtain the Ti3C2Tx MXene/g-C3N4 composite photocatalyst. Figure 3 illustrates the fabrication process of this composite. They found that the 1 wt% Ti3C2Tx MXene/g-C3N4 heterostructure exhibited higher photocatalytic activity for methylene blue degradation compared to pure g-C3N4. This enhanced activity was attributed to intimate interfacial contact (observed via field emission scanning electron microscopy (FESEM) analysis), efficient photo-charge carrier transfer, and a larger BET surface area.
Electrostatically driven self-assembly is a facile approach for synthesizing MXene composites, particularly in solutions. Cai et al. [61] synthesized an Ag3PO4/Ti3C2Tx MXene Schottky catalyst that exhibited prominent photodegradation performance towards various organic dyes, including methyl orange and 2,4-dinitrophenol. In their process, Ti3C2Tx sheets were first dispersed in deionized (DI) water using sonication, followed by the addition of an AgNO3 aqueous solution to the Ti3C2Tx suspension under vigorous stirring. Subsequently, a Na2HPO4 aqueous solution was added dropwise to the mixture and stirred for 2 h. The resulting precipitate was washed multiple times with DI water and dried in a vacuum oven at 80 °C overnight. The negatively charged surface of Ti3C2Tx, due to abundant surface termination groups, facilitated its interaction with Ag3PO4. The apparent rate constant for 2,4-dinitrophenol degradation with Ag3PO4/Ti3C2Tx was 2.5 times higher than that of Ag3PO4/RGO and 10 times higher than that of Ag3PO4 alone. This enhanced photocatalytic activity of Ag3PO4/Ti3C2Tx was attributed to the sufficient and close interfacial contact between Ag3PO4 and Ti3C2Tx, unidirectional electron flow trapped by Ti3C2Tx across the Schottky barrier, and the stronger redox reactivity of surface metal Ti sites on Ti3C2Tx.
Ultrasonic forces have been employed in the fabrication of Ti3C2Tx MXene-based composites to disrupt electrostatic attractions and van der Waals interactions. For example, Lee et al. [62] prepared 2D/2D WO3/Ti3C2Tx heterojunction composites (Figure 4). WO3 nanosheets were dispersed in deionized (DI) water, followed by the addition of varying amounts of Ti3C2Tx nanosheets. The resulting suspension was sonicated, where cavitation bubbles generated localized high-temperature and high-pressure spots, facilitating physical and chemical interactions between WO3 and Ti3C2Tx. This process enabled the effective integration of WO3 nanosheets with Ti3C2Tx structures. The final suspension was dried at 100 °C for 12 h in an electric oven. The WO3/Ti3C2Tx heterojunction demonstrated significantly higher photoexcited carrier transfer and separation efficiency, resulting in exceptional photocatalytic performance for methylene blue (MB) degradation under visible light. Furthermore, photoelectrochemical analysis of WO3/Ti3C2Tx revealed improved charge carrier mobility, effectively reducing the carrier transport barrier between WO3 and Ti3C2Tx. A similar approach has been used to synthesize Ti3C2Tx/CuFe2O4 nanohybrids [63].
A similar process has been used to prepare ZnS nanoparticle/layered MXene sheets [64]. Likewise, manganese oxide-decorated 2D Ti3C2Tx MXene containing MnO2 nanopetals were also synthesized using the ultrasonic approach [65]. The synergistic effect of these two nanomaterials inhibited electron–hole pair recombination and improved surface activity. The nanocomposite exhibited high photocatalytic ability, degrading approximately 99% of MB within 30 min.
The sol-gel method employs metal alkoxide precursors to form gels via hydrolysis, typically at lower temperatures, followed by calcination. For instance, Iqbal et al. [66] utilized a double-solvent solvothermal method to synthesize BiFeO3 (BFO)/Ti3C2Tx nanohybrid. They separately ultrasonicated Ti3C2Tx MXene in deionized (DI) water and BFO nanoparticles in a mixture of acetic acid and ethylene glycol. The resulting solutions were combined and transferred to a Teflon-lined steel autoclave for solvothermal synthesis, where the mixture was heated at 160 °C for 2 h. The final product was washed and dried at 80 °C for 3 h. This nanohybrid exhibited a high BET surface area of 147 m2 g1, a low band gap of 1.96 eV, and a reduced recombination time. These attributes contributed to the material’s superior photocatalytic performance, demonstrated by the degradation of CR dye within a 42 min period under visible light irradiation. A similar process was used to prepare La- and Mn-co-doped BFO nanoparticles embedded in Ti3C2Tx sheets [67].
The solvothermal process, which involves reactions in a solvent under elevated temperatures and pressures within a sealed system, has been employed to synthesize MXene nanocomposites. For instance, Zheng et al. [68] synthesized SnO2/Ti3C2Tx composites, and Zhou et al. [69] prepared CeO2/Ti3C2Tx nanocomposites using CeO2 nanorods on Ti3C2Tx sheets. These nanocomposites demonstrated enhanced photocatalytic activity for RhB photodegradation under UV-light irradiation compared to pure CeO2 semiconductors and Ti3C2Tx. This enhancement was attributed to the composite’s narrower bandgap compared to CeO2, facilitating improved solar energy utilization and resulting in high pollutant degradation efficiency [69]. A similar process was utilized to prepare BiVO4/Ti3C2Tx nanocomposites [70]. This nanohybrid was synthesized by first preparing a 0.005 g/mL Ti3C2Tx MXene solution in D.I. water via 15 min ultrasonication. Separately, 0.40 g of BiVO4 was dispersed in a 1:1 ethanediol–ethanoic acid mixture and ultrasonicated at 80 °C for 30 min. The solutions were then mixed, stirred for 1 h, and subjected to hydrothermal treatment at 150 °C for 3 h in a Teflon-lined autoclave. The product was washed with D.I. water and ethanol, then dried at 80 °C for 6 h, yielding a nanocomposite powder. The fabrication of BiVO4/Ti3C2Tx is illustrated in Figure 5.
Fan et al. [71] synthesized monolayer Ti3C2Tx MXene nanosheets through LiF-HCl etching, followed by washing, centrifugation, and ultrasonic exfoliation. MoS2 nanosheets were produced via lithium-ion etching with n-butyllithium under argon for 48 h, then neutralized, sonicated, and centrifuged. To fabricate MoS2/MXene/NC composite microspheres, MoS2 and MXene dispersions were ultrasonicated, nebulized into droplets, and self-assembled in liquid wax at 150 °C. Subsequently, the microspheres were coated with polydopamine (PDA) using dopamine hydrochloride in Tris buffer. Vacuum carbonization at 700 °C yielded composite microspheres with varying PDA coating times (4, 8, 12 h) and MoS2-to-MXene ratios (1:3, 1:5, 1:7), as detailed in Figure 6.
It was found that the Schottky junction and heterojunction between Ti3C2Tx and MoS2 significantly prolonged the recombination time of electron–hole pairs and broadened the visible light absorption range. Consequently, the nanohybrid exhibited enhanced photocatalytic activity towards MO. It was also found that TiO2/Ti3C2Tx MXene was synthesized via a hydrothermal reaction involving nano-TiO2 and Ti3C2Tx MXene nanosheets [72]. The improved photocatalytic activity was attributed to the suppression of electron–hole recombination, which facilitated electron accumulation and enhanced electron transfer from TiO2 to MXene. Hydrothermal treatment of Ti3C2Tx nanosheets and AgNO3 at 160 °C for 12 h, with a 2 °C/min heating rate, resulted in the fabrication of AgNPs/TiO2/Ti3C2Tx composite [73]. During this process, silver nanoparticles (AgNPs) were deposited on the MXene surface via the self-reduction of silver salts, where MXene acted as a redox agent, facilitating the nucleation and growth of spherical Ag nanoparticles on the MXene nanosheets. The photocatalytic performance of the oxidized form was significantly improved compared to the pristine form. AgNPs/TiO2/Ti3C2Tx also showed superior degradation efficiency for MB and RhB compared to pristine MXene. Bi2WO6/Ti3C2Tx was synthesized using a similar process involving heating layered Ti3C2Tx MXene and Bi(NO3)3·5H2O at 160 °C for 16 h.

4. Photocatalytic Degradation of Dyes Using Ti3C2Tx MXene Hybrids

Photocatalysis presents a relatively safe and cost-effective strategy for degrading hazardous organic pollutants [41,74,75]. Ti3C2Tx MXene, with its unique lamellar structure, remarkably high metallic conductivity, and excellent hydrophilicity, has emerged as a promising member of the MXene family. Its distinctive properties have enabled its application as a photocatalyst for environmental remediation [76,77] and as a co-catalyst for enhancing the photocatalytic degradation potential of composite photocatalysts [34,78]. Ti3C2Tx MXene-based hybrid photocatalysts offer several advantages, including improved charge separation, efficient atomic utilization, tunable bandgap, optimized morphology, enhanced electron transfer efficiency, and increased photocatalytic activity [79,80,81]. This review assesses the effectiveness of Ti3C2Tx MXene-based hybrid photocatalysts by comparing their performance to non-hybrid photocatalysts in the degradation of both cationic and anionic dyes (Table 1).
Table 1 summarizes the results of studies on the photocatalytic degradation of anionic dyes (MO, CR) and cationic dyes (MB, RhB). Dey and Ratan Das [81] achieved 95% degradation of MO within 300 min using CdS as a non-hybrid photocatalyst. In another study, Mohammad Jafri et al. [85] utilized TiO2 nanofibers, achieving 95.2% degradation of MB within 240 min, and Mary et al. [88] reported 97.6% degradation of CR in 75 min using ZnO. Similarly, Fang et al. [90] prepared g-C3N4, achieving 75% degradation of RhB in 180 min. The ZnO standalone photocatalyst required 1260 min (21 h) to degrade 40.88% of MB [84], while MoSe2 achieved only 8.44% degradation in 120 min (2 h) [89]. These prolonged durations and low efficiencies render such photocatalytic processes economically and temporally inefficient for the practical degradation of organic dye pollutants. Furthermore, pristine MXene photocatalysts exhibit lower dye degradation efficiencies. For example, Qu et al. [92] reported that the alkalized Ti3C2Tx demonstrated reduced photocatalytic performance compared to Ti3C2Tx MXene-based hybrid photocatalysts. Specifically, the alkalized Ti3C2Tx achieved degradation rates of only 17.3% for RhB and 2.8% for MO within 120 min.
Conversely, Ti3C2Tx MXene-based hybrid photocatalysts have demonstrated remarkable potential for degrading both cationic and anionic dyes under light irradiation. For example, a Ti3C2Tx MXene-based hybrid photocatalyst achieved a degradation efficiency of 99.7% for MO and 100% for RhB, as detailed in Table 2. These photocatalysts effectively decompose dye molecules into less harmful degradation products, showcasing their efficiency in addressing organic dye pollutants. This high efficiency can be attributed to the rich surface chemistry, tunable bandgap structures, high electrical conductivity, hydrophilicity, thermal stability, and large specific surface area with abundant active sites. These properties facilitate efficient dye adsorption and subsequent photocatalytic degradation. For instance, Nasri et al. [6] reported 100% degradation efficiency for MB using a Ti3C2Tx/g-C3N4 hybrid within 180 min. This exceptional performance was attributed to the effective charge separation and transfer capabilities of Ti3C2Tx MXene, which significantly enhanced the generation of reactive species essential for dye degradation.
Ta et al. [60] demonstrated that ZnO/Ti3C2Tx achieved a 99.7% degradation efficiency for MO, while Bi2WO6/Ti3C2Tx effectively degraded RhB, with an impressive 99.9% efficiency within 20 min., Zhao and Cai [93] reported comparable results (see Table 2). These studies highlight the pivotal role of MXene-incorporated heterostructures, characterized by their high conductivity and ability to facilitate rapid electron transfer, which is essential for efficient photocatalysis [94]. Yao and Wang [95] demonstrated that MB achieved a 93.3% degradation effi-ciency within 160 min using a standalone TiO2 catalyst at a catalyst-to-dye ratio of 1.5:1. However, the MXene-based hybrid AgNPs/TiO2/Ti3C2Tx achieved 99% degrada-tion in just 30 min at a 2.5:1 catalyst-to-dye ratio, demonstrating superior efficiency in both time and degradation rate compared to non-hybrid photocatalysts [73].
Ti3C2Tx MXene-based hybrids represent a versatile and highly effective class of photocatalysts for the degradation of both anionic and cationic dyes, making them suitable for various environmental remediation applications. For example, Ti3C2Tx MXene-based nanocomposites prepared with Mn2O3 were evaluated for photocatalytic dye degradation under light, demonstrating efficient photocatalysis. The one-dimensional (1D) Mn2O3-Ti3C2Tx (20 wt%) nanocomposite achieved 100% degradation of MB within 25 min, effectively removing the dye [96]. Similarly, NiM-nO3/NiMn2O4-Ti3C2Tx MXene nanocomposites achieved 100% degradation of MB in 50 min, showcasing excellent dye removal efficiency [97]. Wang et al. [98] synthesized a Ti3C2Tx/Bi4Ti3O12 heterojunction via a facile in-situ solvothermal method, demonstrating exceptional visible-light-driven photocatalytic performance by achieving 100% degradation of MO and RhB within 60 and 50 min, respectively (see Table 2). These results highlight the potential of Ti3C2Tx MXene-based hierarchical composites for water remediation, offering a sustainable approach for the degradation of anionic and cationic organic pollutants. In an independent study, Iqbal et al. [66] reported that the BiFeO3 (BFO)/Ti3C2Tx MXene hybrid achieved 100% degradation of CR in 42 min, highlighting its potential for photocatalysis applications.
It was observed that factors such as the synthesis method, pH, catalyst-to-dye ratio, concentration, and structure significantly influence the functionality and effectiveness of photocatalysts. Based on this evaluation, it is concluded that Ti3C2Tx MXene-based hybrid photocatalysts exhibit superior photocatalytic performance and efficiency in degrading organic pollutant dyes compared to non-hybrid photocatalysts.
Table 2. Ti3C2Tx MXene-based hybrid photocatalysts with photocatalytic degradation performance towards cationic and anionic dyes.
Table 2. Ti3C2Tx MXene-based hybrid photocatalysts with photocatalytic degradation performance towards cationic and anionic dyes.
MXene-Based Hybrid PhotocatalystsDyesType of Dyes
(Based on Charge)
Degradation
Percentage/(Time)
References
TiO2/Ti3C2TxMOAnionic 92 (50 min)[59]
MoS2@Ti3C2MOAnionic98 (60 min)[99]
Ti3C2/TiO2/CuOMOAnionic99 (80 min)[100]
ZnO/Ti3C2TxMOAnionic99.7 (50 min) [60]
Ti3C2Tx/Bi4Ti3O12MOAnionic100 (60 min)[98]
TiO2/Ti3C2 MxeneMBCationic 96.44 (60 min)[101]
AgNPs/TiO2/Ti3C2Tx MBCationic 99 (30 min)[73]
Ti3C2/g-C3N4MBCationic 100 (180 min)[6]
NiMnO3/NiMn2O4-Ti3C2Tx MXeneMBCationic 100 (50 min)[97]
1D Mn2O3-Ti3C2Tx MBCationic 100 (25 min)[96]
Mn-codoped bismuth
ferrite/Ti3C2
CRAnionic93 (30 min)[67]
CoFe2O4@MXeneCRAnionic98.9 (30 min)[102]
BiVO4/Ti3C2CRAnionic99.5 (60 min)[70]
BGFO-20Sn/MXeneCRAnionic100 (120 min)[103]
BiFeO3 (BFO)/Ti3C2CRAnionic100 (42min)[66]
TiO2@Ti3C2RhBCationic97 (40 min)[58]
BiOBr/TiO2/
Ti3C2Tx
RhBCationic99.8 (30 min)[104]
Bi2WO6/Ti3C2RhBCationic99.9 (20 min)[93]
ZnS/MXeneRhBCationic100 (100 min)[64]
Ti3C2Tx/Bi4Ti3O12RhBCationic100 (50 min)[98]

5. Computational Studies and Simulations

The investigation of Ti3C2Tx MXene-based hybrids for photocatalytic applications is significantly enhanced by computational tools and techniques, which complement experimental approaches. These methods offer insights into atomic-level phenomena, guide material design, and predict performance under varying conditions.
Chen et al. [101] investigated the electronic and optical properties of -F terminated, -O terminated, and termination-free Ti3C2 in an MXene nanosheet/TiO2 composite using density functional theory (DFT). Their study revealed that surface terminations reduced the density of electronic states, lowered conductivity, and enhanced stability compared to the termination-free MXene. DFT analysis also demonstrated the feasibility of electron transfer from TiO2 to Ti3C2 and identified the Schottky barrier at the interface between the two materials. Furthermore, computational modeling highlighted the synergy between the composite components, showing an extended range of light absorption, suppressed electron–hole recombination, and improved hole oxidation efficiency in the VB of TiO2. These factors significantly enhanced the photocatalytic performance of the Ti3C2/TiO2 composite, making it a promising candidate for photocatalytic applications, such as treating organic pollutants like dye molecules [101]. Lemos et al. utilized a computational model to evaluate the performance of a Ti3C2Tx/TiO2 nanocomposite hybrid for photocatalyzed dye-sensitized solar cells. Through DFT calculations, they discovered that the anatase potential is reduced at the nanocomposite interface and that the nanocomposite exhibits improved photocarrier separation at the interface between the nanocomposite and the dye [105]. Furthermore, Yang et al. [106] conducted a computational analysis to confirm that a transition metal dichalcogenide/MXene photocatalyst hybrid, MoS2/Ti3C2, functions as a Schottky barrier. A Bader charge analysis revealed that the difference in work functions between MoS2 and Ti3C2, combined with a built-in electric field, facilitates the transfer of photogenerated electrons from MoS2 to the Ti3C2 electron sink. This efficient electron transfer enhances photocarrier separation, resulting in longer lasting photogenerated holes and exceptional photodegradation performance against rhodamine B dye in wastewater [106].
Several computational techniques have been employed to assess and optimize photocatalytic mechanisms in MXene-based composites for dye degradation applications. Liu et al. [107] developed a g-C3N4/Ti3C2 (CNTC) heterojunction by hybridizing 2D Ti3C2 MXene with three-dimensional (3D) g-C3N4 for enhanced photodegradation of RhB dye. This photocatalyst exhibited a high specific surface area (85.08 m2/g) and remarkable charge migration capabilities. To evaluate its improved photocatalytic performance, the researchers utilized DFT analysis to examine the differential charge density, electron distribution, and charge transfer dynamics between g-C3N4 and the Ti3C2 sink. The study revealed that the excellent conductivity of Ti3C2 stemmed from the overlap between the Fermi level and CB in the heterojunction. This led to the understanding that these 2D/3D heterojunctions significantly promote charge transfer and separation, which are essential for efficient photocatalysis. Additionally, the combination of a high specific surface area and abundant active sites makes the CNTC particularly effective for dye photodegradation [107]. Cheng et al. [108] synthesized a self-cleaning BiOCl-polypyrrole Ti3C2Tx MXene composite membrane with excellent photocatalytic activity and high flux, designed for filtering and degrading pollutants. The membrane’s performance was tested against various dyes. To gain deeper insight into the photocatalytic mechanisms, particularly charge separation and degradation, the researchers conducted DFT calculations. Density of state (DOS) simulations revealed that the chemisorption of oxygen into vacancies on the BiOCl surface generated superoxide radicals. These radicals enhanced the composite’s photocatalytic efficiency through redox interactions with dye molecules and other pollutants [108]. Wang et al. [109] synthesized an S-scheme Pt-MnO2/TiO2@Ti3C2Tx composite using an electrostatically self-assembled Ti-O-Mn bond and evaluated its oxidative photodegradation performance against MB, MO, and RhB dyes. To investigate the photocatalytic mechanisms, DFT calculations were performed, revealing that the Ti-O-Mn bond induced the formation of metastable Ti atoms and electrostatically adsorbed Mn2+ ions. This bond facilitated the separation of photoinduced carriers and optimized their transport pathways. Additionally, the DFT analysis identified the formation of S-scheme heterojunctions between MnO2 and TiO2 through the Ti-O-Mn bond, driving the flow of photogenerated carriers. These factors collectively enhanced the composite’s photocatalytic efficiency [109].
DFT has been employed in conjunction with the finite element method to explore ways to enhance the photocatalytic activity and membrane permeability of a novel MXene-based composite membrane, N-doped Bi2O2CO3@Ti3C2Tx/Polyethersulfone, designed for oil/water separation and dye degradation [110]. Through DFT analysis of the electron distribution and band structure of doped versus undoped Bi2O2CO3, it was found that N doping improved conductivity, enhanced electron transition activity, and facilitated photogenerated carrier transport (attributed to VB dispersion), thereby making Bi2O2CO3@Ti3C2Tx/Polyethersulfone more effective for photocatalytic applications [110].

6. Other Applications of Ti3C2Tx MXene-Based Hybrid Photocatalysts

Beyond organic dye degradation, Ti3C2Tx MXene-based hybrid photocatalysts have demonstrated potential for various other applications. They have shown considerable promise in wastewater treatment, particularly in degrading organic pollutants such as dyes, pharmaceuticals, and pesticides [111]. Composite materials of Ti3C2Tx MXene with semiconductors such as TiO2 or ZnO have exhibited even greater photocatalytic efficiencies; the synergy between MXenes and these semiconductors results in improved light absorption and charge separation [112,113]. For instance, Ti3C2Tx/TiO2 composites demonstrate enhanced photocatalytic activity under visible light due to the synergistic effects of both materials. These composites can efficiently degrade various dyes under sunlight, including RhB and MO, making them suitable for sustainable and cost-effective wastewater treatment [114]. Furthermore, doping MXenes with other elements or combining them with carbon-based materials like graphene can further enhance their photocatalytic properties. Nitrogen-doped Ti3C2Tx MXene shows improved photocatalytic degradation of antibiotics such as tetracycline under visible light, highlighting their potential in treating pharmaceutical contaminants in wastewater [115]. Ti3C2Tx MXene-based hybrid photocatalysts are also being explored for air purification applications, particularly in removing volatile organic compounds (VOCs) and other airborne pollutants. Ti3C2Tx MXene, when combined with photocatalysts like TiO2, can effectively degrade formaldehyde and toluene, common indoor air pollutants, under UV and visible light [116]. Incorporating noble metals like Au or Ag onto Ti3C2Tx MXene can further enhance their photocatalytic performance by creating localized surface plasmon resonance, increasing light absorption, and improving the degradation rates of VOCs [117]. Additionally, Z-scheme heterostructures involving Ti3C2Tx MXene have been developed to mimic natural photosynthesis, achieving efficient separation and transfer of photogenerated charge carriers and enhancing photocatalytic degradation of air pollutants [118]. The most applied technology is the Ti3C2Tx MXene-based TiO2 photocatalyst, which utilizes photocatalysis for glass cleaning, where UV or visible light activates TiO2, generating reactive oxygen species that decompose organic pollutants on the glass surface. This self-cleaning mechanism maintains transparency, reduces manual cleaning, and prevents pollutant buildup, enhancing the efficiency and durability of glass surfaces.
In addition to the degradation of dyes, the degradation of other organic pollutants, such as pharmaceutical waste and VOCs, has emerged as a challenging task requiring immediate attention [119]. In recent years, extensive studies [120] have highlighted photocatalysis as an attractive method for the efficient degradation of organic pollutants. The photocatalytic performance of a catalyst is defined by properties such as its surface-to-volume ratio, light interaction, and mechanical stability [121]. Among the variety of materials reported thus far, Ti3C2Tx MXenes have gained significant attention as photocatalytic materials [122]. Furthermore, the integration of Ti3C2Tx MXenes with other nanomaterials can significantly improve photocatalytic performance [123].
Kumar et al. [124] reported a novel photocatalyst composed of g-C3N4, Ti3C2Tx, and Au nanoparticles for the degradation of cefixime. Figure 7a illustrates the degradation mechanism of cefixime decomposition under light. The composition with 3 wt% Ti3C2Tx demonstrated the highest degradation, achieving 64.69% cefixime removal in 105 min under visible light irradiation, as shown in Figure 7b,c. Similarly, Diao et al. [125] reported efficient photocatalytic degradation of tetracycline hydrochloride using MXene-based photocatalysts comprising g-C3N4/Ti3C2Tx/TiO2. The reported ternary catalysts exhibited superior performance, chemical and photostability, and recyclability. Another MXene-based ternary photocatalyst, reported by Zhou et al. [126], facilitated photocatalytic degradation of enoxacin under visible light, where Ti3C2Tx improved charge separation at the interface, resulting in efficient degradation. Rdewi et al. [127] reported a ZnO-TiO2-MXene photocatalyst for the decomposition of carbamazepine molecules in wastewater under solar irradiation. The presented results showed 99.6% removal efficiency at pH 7, attributed to improved charge carrier transport and a reduced recombination rate due to the incorporation of TiO2-MXene with ZnO photocatalysts. The excellent photocatalytic performance also demonstrated reusability across multiple decomposition cycles with exceptional efficiency. Similarly, Abbas et al. [128] reported a ZnO-TiO2-MXene photocatalyst for the decomposition of ceftriaxone sodium molecules in water using simulated solar light.
Sukidpaneenid et al. [129] reported Ti3C2Tx/TiO2 photocatalysts for the degradation of enrofloxacin antibiotics in water. The precise control over TiO2 loading on MXene, achieved through variations in hydrothermal processes, played a crucial role in tuning photocatalytic properties. Additionally, the intercalation of sodium ions significantly improved adsorption, and the synergistic effect of TiO2 loading and NaCl treatment led to the efficient removal of enrofloxacin. Shahzad et al. [130] demonstrated the degradation of carbamazepine (CBZ) under direct sunlight and UV light using Ti3C2Tx-based heterostructure photocatalysts with {001} TiO2. The results showed that the Kapp value under UV irradiation was higher than under sunlight for CBZ degradation. Also, the effects of pH on degradation performance were considered. Mohanty et al. [131] reported a series of SrTiO3/Ti3C2Tx-based photocatalysts decorated with Au nanoparticles for the degradation of colorless organic pollutants such as ciprofloxacin under sunlight. The significant enhancement in photocatalytic degradation for the plasmon-mediated heterostructure catalysts was attributed to the absorption of a broad solar spectrum, charge separation, and charge transport. Du et al. [132] demonstrated the photocatalytic degradation of tetracycline using CeO2-Ti3C2Tx-TiO2 (CeMXT). The composite exhibited an excellent degradation efficiency of 94.70% for 22.19 mg/L in 104.13 min with 0.65 g/L of catalyst at pH 4.72. This excellent degradation efficiency was attributed to increased radical generation and charge separation.
Sergi et al. [133] reported a series of TiO2-Ti3C2Tx MXene photocatalysts with controlled composition for improved photocatalytic removal of benzene. The incorporation of TiO2 with Ti3C2Tx MXene enhanced optical absorption, leading to improved photocatalytic performance. The report demonstrated the potential of Ti3C2Tx MXenes for VOC removal and the ability of heterostructures to enhance photocatalysis. Furthermore, Huang et al. [116] reported the degradation of formaldehyde (HCHO) and acetone (CH3COCH3) using Bi2WO6/Ti3C2 (BT4) under light irradiation. Bi2WO6 photocatalysts suffer from a high recombination rate, and combination with Ti3C2 can significantly improve charge separation. Ti3C2-mediated charge separation caused charge transfer at the interface, significantly improving the photocatalytic performance of BT4. Additionally, DFT simulations indicated higher adsorption of formaldehyde and acetone on the Ti3C2 surface compared to the Bi2WO6 surface, synergistically tuning photocatalytic performance. Notably, the degradation of formaldehyde and acetone for BT4 was 2 and 6.6 times higher, respectively, than for Bi2WO6. Mo et al. [134] demonstrated photocatalytic and photothermal removal of VOCs (phenol) from water using a TiO2/Ti3C2Tx/C3N4/PVA (TTCP) hydrogel under sunlight. Figure 7d shows the SEM image of TTCP. The VOC removal efficiency varied from 69.4% to 100% at different phenol concentrations (1–50 mg/L), as shown in Figure 7e,f. The membrane significantly lowered the total dissolved solids (TDSs) and TOC. The TDS level was reduced by more than two orders of magnitude, and the TOC removal efficiency was observed to be 80%.
Figure 7. (a) Schematic of the cefixime degradation mechanism under light. (b) Kinetics of cefixime degradation. (c) Histogram of degradation rate (%). Reproduced with permission from [124], (d) SEM image of the tetracalcium phosphate (TTCP) hydrogel. (e) Total organic carbon (TOC) of source water under different phenol concentrations (orange column), distilled water without photocatalyst (green column), and distilled water with TTCP hydrogel (violet column). (f) TOC of distilled water with different catalysts. Reproduced with permission from [134]. (g) Schematic illustration of the synthesis process of the Ti3C2Tx MXene/CeO2 photocatalysts (inset SEM image). Comparative representation of the production of (h) C2H5OH and (i) CH4 under solar light illumination with different catalysts. Reproduced with permission from [135].
Figure 7. (a) Schematic of the cefixime degradation mechanism under light. (b) Kinetics of cefixime degradation. (c) Histogram of degradation rate (%). Reproduced with permission from [124], (d) SEM image of the tetracalcium phosphate (TTCP) hydrogel. (e) Total organic carbon (TOC) of source water under different phenol concentrations (orange column), distilled water without photocatalyst (green column), and distilled water with TTCP hydrogel (violet column). (f) TOC of distilled water with different catalysts. Reproduced with permission from [134]. (g) Schematic illustration of the synthesis process of the Ti3C2Tx MXene/CeO2 photocatalysts (inset SEM image). Comparative representation of the production of (h) C2H5OH and (i) CH4 under solar light illumination with different catalysts. Reproduced with permission from [135].
Molecules 30 01463 g007
The photocatalytic reduction of CO2 is another aspect of environmental remediation applications. Reducing CO2 into useful byproducts using light-activated catalysts represents a futuristic pathway towards sustainability [90,136]. The excellent charge separation and slow recombination rate of Ti3C2Tx MXene demonstrate its potential as a catalytic material for CO2 reduction [137]. In this regard, Cao et al. [52] demonstrated 2D/2D ultrathin Ti3C2Tx/Bi2WO6 nanosheets prepared by in situ growth of Bi2WO6 nanosheets over Ti3C2Tx nanosheets. The proposed hybrid catalyst exhibited improved efficiency towards CO2 reduction under light due to reduced charge transfer distance, improved surface-to-volume ratio, and increased adsorption sites. The hybrid catalyst showed improved production of CH4 and CH3OH compared to Bi2WO6. Quantitatively, the production of CH4 increased from 0.41 μmol g1 h1 to 1.78 μmol g1 h1, while the production of CH3OH increased from 0.07 μmol g1 h1 to 0.58 μmol g1 h1. Similarly, Mishra et al. [135] investigated a Ti3C2Tx-CeO2 hybrid catalyst with varying Ti3C2Tx ratios for the photocatalytic reduction of CO2. The schematic synthesis process for the hybrid catalysts is shown in Figure 7g. Surprisingly, the hybrid catalyst with 5 wt% Ti3C2Tx/CeO2 produced 6127.04 μmol g1 of ethanol and 129.5 μmol g1 of methane within 5 h, significantly higher than CeO2, as shown in Figure 7h,i. Another report by Li et al. [137] demonstrated the potential of a Ti3C2-based hybrid catalyst comprising g-C3N4/ZnO/Ti3C2Tx for CO2 reduction into methane (CH4) and carbon monoxide (CO). The hybrid framework exhibited a notably higher production rate of 1.41 μmol g1 h1 towards CO, increased by factors of 2.7 and 1.7 compared to ZnO and g-C3N4, respectively.
Ti3C2Tx MXene has been extensively studied in recent years for its broad range of photocatalytic applications in environmental remediation, as evidenced by the literature. Reported results highlight the potential of Ti3C2Tx MXene-based catalysts for degrading dyes, pharmaceutical wastes, and VOCs, as well as for CO2 reduction. These photocatalysts benefit from a high surface area, efficient charge separation at the interface, and excellent light absorption, which contribute significantly to their performance.

7. Working Mechanism of Ti3C2Tx MXene-Based Hybrid Photocatalyst

The electronic structure of Ti3C2Tx MXene-based photocatalysts plays a crucial role in their ability to degrade targeted pollutants through photocatalysis. Generally, an ideal photocatalyst should possess a suitable bandgap, strong absorption in the visible range, prolonged charge separation lifetimes, and adequate redox potential [64]. To enhance photocatalysis efficiency, co-catalysts are employed to separate photogenerated charge carriers. When materials such as ZnO, CdS, TiO2, and Ti3C2 are used as co-catalysts, Schottky junctions formed with Ti3C2Tx facilitate the rapid dissociation of these charge carriers [138]. As discussed in the previous section, Ti3C2Tx MXene-based photocatalyst composites demonstrate superior photocatalytic performance compared to non-hybrid photocatalysts. This is attributed to the abundant active sites available on Ti3C2Tx MXene-based hybrids, their enhanced bandgap, improved light-harvesting capability, reduced charge carrier recombination, and extended photoelectron lifetime [139].
During photocatalytic degradation, Ti3C2Tx MXene-based hybrid photocatalysts absorb visible light, exciting photogenerated electrons into the CB and leaving holes in the VB [22]. These excited charge carriers are subsequently transferred to the Ti3C2Tx MXene at the interface, primarily due to the higher potential of Ti3C2Tx. While the Schottky junction formed between n-type TiO2 and Ti3C2 facilitates hole transfer from TiO2 to Ti3C2, the inherent band bending in n-type TiO2 creates an energy barrier that inhibits direct electron transfer from Ti3C2 to TiO2 [140]. This limitation means that photogenerated electrons on Ti3C2 primarily participate in reactions with adsorbed oxygen, leading to the formation of superoxide radicals (O2). Consequently, the primary role of the Ti3C2 in this hybrid system is to act as a hole reservoir, preventing hole recombination at the Ti3C2–TiO2 interface, rather than directly contributing to electron transfer towards reduction reactions on the TiO2 surface. In this process, photogenerated electrons from the Ti3C2Tx MXenes migrate to the surface and react with O2 to produce O2, and holes from the TiO2 react with hydroxyl ions (OH) to produce hydroxyl radicals (OH). These radicals cause the degradation of cationic and anionic dyes of organic pollutants [22,141,142].
A photocatalytic mechanism for the Ti3C2Tx–TiO2 hybrid, illustrating anionic dye (e.g., MO) degradation, is presented in Figure 8a. Initially, the light source provides high-energy photons, which activate the TiO2 component, generating photoinduced electrons in its CB and holes in its VB. These photoelectrons rapidly migrate from the CB of TiO2 to the Ti3C2Tx MXene, facilitated by its high electrical conductivity [73,114]. Consequently, Ti3C2Tx MXene accumulates a negative charge, while TiO2 becomes positively charged, leading to the formation of a Schottky barrier at the Ti3C2Tx–TiO2 interface, which functions as a space-charge layer. Subsequently, the photogenerated electrons on Ti3C2Tx MXene migrate to its surface, where they react with O2 molecules to generate superoxide radicals (O2) [114,143]. Simultaneously, the photogenerated holes react with adsorbed OH to form hydroxyl radicals (OH) [58]. These radicals play a crucial role in MO degradation. The photocatalytic reaction mechanism facilitated by the MXene-based hybrid photocatalyst (TiO2/Ti3C2Tx) is illustrated in Figure 8a [114]. The mechanism can be roughly described through the following reactions:
Ti3C2–TiO2 + hϑ → TiO2 (holes) + Ti3C2 (electrons)
holes + OHOH
electrons + O2O2−
Dye (Anionic or cationic) + O2− → CO2 + H2O
Dye (Anionic or cationic) + OH→ CO2 + H2O
For the degradation of cationic dye like MB, O2− is produced through a reduction reaction between electrons and O2 molecules, leading to the direct degradation of MB. Additionally, light-induced holes partially oxidize MB and partially react with water to generate OH, ultimately breaking down MB [144]. As evidenced by Figure 8b, OH and holes are the primary reactive species in the TiO2/Ti3C2Tx MXene composite photocatalysis reaction system [72]. Ultimately, these radicals (OH, O2−), possessing strong oxidizing abilities, degrade both anionic and cationic dyes, such as MO and MB molecules, directly into their oxidation products [145,146]. Therefore, a similar photocatalytic reaction mechanism occurs for both cationic and anionic dyes under light irradiation.

8. Challenges and Future Directions

While Ti3C2Tx MXene-based hybrid photocatalysts have demonstrated significant promise for organic dye pollutant degradation and environmental remediation, several challenges must be addressed to further enhance their photocatalytic performance and enable practical application.
One of the most significant challenges in MXene research is the development of green synthesis methods. Current synthesis processes often involve harsh and hazardous chemicals for etching and exfoliation, such as hydrofluoric acid, tetramethylammonium hydroxide, and tetrabutylammonium hydroxide [36]. Eliminating these chemicals is crucial for the wider applicability of MXenes.
Furthermore, the limited yield of high-quality MXene products during synthesis is another major obstacle. Current large-scale synthesis processes are inefficient, hindering commercial use. Therefore, future research efforts should focus on developing mass-production methods for high-quality, uniformly delaminated single- to few-layer MXene nanosheets. This will require a deeper understanding of the reaction kinetics and thermodynamics of the synthesis process to achieve uniform and scalable production.
The long-term stability of MXenes in harsh environmental conditions, such as oxygen, high humidity, and acidic or alkaline media, remains a significant concern [38]. Corrosion and degradation over time can compromise the catalytic efficiency of composite materials and hinder their reuse. Thermal stability is also a considerable challenge for MXene-based photocatalysts, as elevated temperatures can accelerate MXene oxidation [147].
Significant efforts have been made to address the challenges associated with MXene-based hybrid photocatalysts. The development of efficient MXene-based hybrid photocatalysts for practical applications requires a scalable synthesis of high-quality MXenes. However, achieving this remains challenging without a comprehensive understanding of the reaction kinetics and thermodynamics. Currently, the kinetics and thermodynamics of A-layer etching in MAX phases to obtain corresponding MXene layers are still underexplored [148,149,150]. Additionally, the photothermal effect of MXenes is a crucial thermodynamic factor that significantly impacts photocatalytic reactions. Upon irradiation, MXenes generate heat, raising the surrounding temperature and influencing the thermodynamics of charge carriers [151]. Accelerated thermodynamics up to an optimal temperature enhances photocatalytic efficiency; however, excessive heat can negatively impact reactant adsorption [152]. Hence, future research should focus on in-depth studies of reaction kinetics and thermodynamics to achieve uniform and large-scale MXene synthesis and MXene-based hybrid photocatalyst fabrication for practical applications.
Environmentally friendly, large-scale synthesis routes and effective antioxidation strategies for MXenes should be explored to ensure their long-term stability and sustainability. Future Ti3C2Tx-based hybrid photocatalyst materials should also utilize their functional terminations, which serve as active sites for reaction centers, promoting covalent interactions with photocatalyst materials. This approach may enhance hybrid materials’ robustness, environmental stability, and efficiency.
To transition Ti3C2Tx MXene-based hybrid photocatalysts from laboratory to real-world applications, strategies for integrating them into practical systems, such as water treatment plants or portable water purification devices, should be explored. The development of user-friendly photocatalytic reactors or devices utilizing MXene-based hybrid materials for organic dye and pollutant degradation will advance these materials to the forefront of environmental remediation technologies. Conventional methods for discovering materials with desired properties are often time-consuming and complex. To address these challenges, a machine learning (ML)-driven approach can be employed to predict and understand the properties of functional materials more efficiently [153,154]. When combined with experimental data, data-driven ML models enable researchers to uncover natural correlations between catalyst properties and the degradation rates of contaminants in water resources. Significant progress has already been made in applying ML to study the catalytic activity of both standalone photocatalysts, such as ZnO [155] and TiO2 [156], as well as nanocomposites, including Bi2WO6/MIL-53 [155] and CuWO4@TiO2 [157], for degrading various organic dyes. However, the application of such modern artificial intelligence (AI)-based approaches in the field of MXene-based hybrid photocatalysts remains limited. To enhance the efficiency of Ti3C2Tx MXene-based hybrid photocatalysts, a predictive design strategy integrating AI-driven modeling with existing databases can be utilized. This methodology enables the identification and screening of optimal material combinations to enhance photocatalytic performance.

9. Conclusions

In summary, this review has examined the fabrication and evaluated the performance of Ti3C2Tx MXene-based hybrid photocatalysts for the degradation of organic dye pollutants. Ti3C2Tx MXene exhibits exceptional physical and chemical properties, including high electrical conductivity, excellent hydrophilicity, adsorption capability, and efficient charge transfer, while also suppressing electron–hole recombination during photocatalysis. This makes Ti3C2Tx MXene a promising, low-cost, and scalable alternative to noble metal catalysts, offering exceptional catalytic performance in hybrid photocatalysts. The evaluation demonstrated that Ti3C2Tx MXene-based hybrid photocatalysts significantly enhance dye degradation efficiency, as evidenced by increased percentage degradation and reduced degradation time, compared to non-hybrid or pure semiconducting materials. This review also provided a comprehensive understanding of the dye degradation mechanisms involving Ti3C2Tx MXene-based hybrid photocatalysts and highlighted various computational studies and simulations that have advanced research in this field. Furthermore, the application of ML techniques holds significant promise for optimizing the design and performance of these photocatalysts. Integrating ML into future research can accelerate the discovery of novel MXene-based photocatalysts with tailored properties. Finally, the challenges associated with Ti3C2Tx MXene-based hybrid photocatalysts were thoroughly identified, and future research directions were suggested to effectively address these challenges.

Funding

T.L. sincerely acknowledges support for this project provided by the Faculty Research Support Fund (FRSF), UHCL (Award # $2408), and NASA Bridge Program Seed Funding under grant # 80NSSC24K0417. F.Y. acknowledges support from the U.S. National Science Foundation (NSF) under grant # DMR-2122044 and the U.S. Army Research Office (ARO) under grant # W911NF2210109. J.P. acknowledges support from the Endowed Professorship in Science # 45 for 2024–2025 from McNeese State University.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AbbreviationDefinition
1DOne-dimensional
2DTwo-dimensional
3DThree-dimensional
AB92Acid blue 92
AIArtificial intelligence
AO7Acid Orange 7
BETBrunauer–Emmett–Teller
CBZCarbamazepine
CBConduction band
CNTCg-C3N4/Ti3C2
CRCongo red
CVCrystal violet
DIDeionized
DMSODimethyl sulfoxide
DOSDensity of state
DFTDensity functional theory
FESEMField emission scanning electron microscopy
g-C3N4Graphitic carbon nitride
HFHydrogen fluoride
MBMethylene blue
MLMachine learning
PDAPoly-dopamine
RhBRhodamine B
SEMScanning electron microscopy
TOCTotal organic carbon
TTCPTetra calcium phosphate
TDSsTotal dissolved solids
UVUltraviolet
VBValence band
VOCsVolatile organic compounds
XRDX-ray diffraction

References

  1. Singh, B.J.; Chakraborty, A.; Sehgal, R. A Systematic Review of Industrial Wastewater Management: Evaluating Challenges and Enablers. J. Environ. Manag. 2023, 348, 119230. [Google Scholar] [CrossRef]
  2. Ahmed, J.; Thakur, A.; Goyal, A. Industrial Wastewater and Its Toxic Effects. In Biological Treatment of Industrial Wastewater; Shah, M.P., Ed.; The Royal Society of Chemistry: London, UK, 2021; pp. 1–14. ISBN 978-1-83916-279-4. [Google Scholar]
  3. López-Ahumada, E.; Salazar-Hernández, M.; Talavera-López, A.; Solis-Marcial, O.J.; Hernández-Soto, R.; Ruelas-Leyva, J.P.; Hernández, J.A. Removal of Anionic and Cationic Dyes Present in Solution Using Biomass of Eichhornia Crassipes as Bioadsorbent. Molecules 2022, 27, 6442. [Google Scholar] [CrossRef]
  4. Salleh, M.A.M.; Mahmoud, D.K.; Karim, W.A.W.A.; Idris, A. Cationic and Anionic Dye Adsorption by Agricultural Solid Wastes: A Comprehensive Review. Desalination 2011, 280, 1–13. [Google Scholar] [CrossRef]
  5. Zhong, Q.; Li, Y.; Zhang, G. Two-Dimensional MXene-Based and MXene-Derived Photocatalysts: Recent Developments and Perspectives. Chem. Eng. J. 2021, 409, 128099. [Google Scholar] [CrossRef]
  6. Nasri, M.S.I.; Samsudin, M.F.R.; Tahir, A.A.; Sufian, S. Effect of MXene Loaded on G-C3N4 Photocatalyst for the Photocatalytic Degradation of Methylene Blue. Energies 2022, 15, 955. [Google Scholar] [CrossRef]
  7. Soni, V.; Singh, P.; Phan Quang, H.H.; Parwaz Khan, A.A.; Bajpai, A.; Van Le, Q.; Thakur, V.K.; Thakur, S.; Nguyen, V.-H.; Raizada, P. Emerging Architecture Titanium Carbide (Ti3C2Tx) MXene Based Photocatalyst toward Degradation of Hazardous Pollutants: Recent Progress and Perspectives. Chemosphere 2022, 293, 133541. [Google Scholar] [CrossRef]
  8. Khan, S.; Noor, T.; Iqbal, N.; Yaqoob, L. Photocatalytic Dye Degradation from Textile Wastewater: A Review. ACS Omega 2024, 9, 21751–21767. [Google Scholar] [CrossRef]
  9. Gupta, T.; Chauhan, R.P. Photocatalytic Degradation of Water Pollutants Using II-VI Semiconducting Catalysts: A Comprehensive Review. J. Environ. Chem. Eng. 2021, 9, 106734. [Google Scholar] [CrossRef]
  10. Rafiq, A.; Ikram, M.; Ali, S.; Niaz, F.; Khan, M.; Khan, Q.; Maqbool, M. Photocatalytic Degradation of Dyes Using Semiconductor Photocatalysts to Clean Industrial Water Pollution. J. Ind. Eng. Chem. 2021, 97, 111–128. [Google Scholar] [CrossRef]
  11. Guo, R.; Wang, J.; Bi, Z.; Chen, X.; Hu, X.; Pan, W. Recent Advances and Perspectives of g–C3N4–Based Materials for Photocatalytic Dyes Degradation. Chemosphere 2022, 295, 133834. [Google Scholar] [CrossRef]
  12. Yadav, S.; Shakya, K.; Gupta, A.; Singh, D.; Chandran, A.R.; Varayil Aanappalli, A.; Goyal, K.; Rani, N.; Saini, K. A Review on Degradation of Organic Dyes by Using Metal Oxide Semiconductors. Env. Sci. Pollut. Res. 2022, 30, 71912–71932. [Google Scholar] [CrossRef]
  13. Wu, L.; Li, Q.; Yang, C.; Ma, X.; Zhang, Z.; Cui, X. Constructing a Novel TiO2/γ-Graphyne Heterojunction for Enhanced Photocatalytic Hydrogen Evolution. J. Mater. Chem. A 2018, 6, 20947–20955. [Google Scholar] [CrossRef]
  14. Li, X.; Yu, J.; Wageh, S.; Al-Ghamdi, A.A.; Xie, J. Graphene in Photocatalysis: A Review. Small 2016, 12, 6640–6696. [Google Scholar] [CrossRef] [PubMed]
  15. Chitara, B.; Dimitrov, E.; Liu, M.; Seling, T.R.; Kolli, B.S.C.; Zhou, D.; Yu, Z.; Shringi, A.K.; Terrones, M.; Yan, F. Charge Transfer Modulation in Vanadium-Doped WS2/Bi2O2Se Heterostructures. Small 2023, 19, 2302289. [Google Scholar] [CrossRef]
  16. Yang, J.; Wang, D.; Han, H.; Li, C. Roles of Cocatalysts in Photocatalysis and Photoelectrocatalysis. Acc. Chem. Res. 2013, 46, 1900–1909. [Google Scholar] [CrossRef]
  17. Ramalingam, G.; Perumal, N.; Priya, A.K.; Rajendran, S. A Review of Graphene-Based Semiconductors for Photocatalytic Degradation of Pollutants in Wastewater. Chemosphere 2022, 300, 134391. [Google Scholar] [CrossRef]
  18. Zhou, W.; Pan, K.; Qu, Y.; Sun, F.; Tian, C.; Ren, Z.; Tian, G.; Fu, H. Photodegradation of Organic Contamination in Wastewaters by Bonding TiO2/Single-Walled Carbon Nanotube Composites with Enhanced Photocatalytic Activity. Chemosphere 2010, 81, 555–561. [Google Scholar] [CrossRef]
  19. Sharma, S.; Dutta, V.; Singh, P.; Raizada, P.; Rahmani-Sani, A.; Hosseini-Bandegharaei, A.; Thakur, V.K. Carbon Quantum Dot Supported Semiconductor Photocatalysts for Efficient Degradation of Organic Pollutants in Water: A Review. J. Clean. Prod. 2019, 228, 755–769. [Google Scholar] [CrossRef]
  20. Tong, T.; Zhang, M.; Chen, W.; Huo, X.; Xu, F.; Yan, H.; Lai, C.; Wang, W.; Hu, S.; Qin, L.; et al. Recent Advances in Carbon-Based Material/Semiconductor Composite Photoelectrocatalysts: Synthesis, Improvement Strategy, and Organic Pollutant Removal. Coord. Chem. Rev. 2024, 500, 215498. [Google Scholar] [CrossRef]
  21. Nenashev, G.V.; Istomina, M.S.; Kryukov, R.S.; Kondratev, V.M.; Shcherbakov, I.P.; Petrov, V.N.; Moshnikov, V.A.; Aleshin, A.N. Effect of Carbon Dots Concentration on Electrical and Optical Properties of Their Composites with a Conducting Polymer. Molecules 2022, 27, 8000. [Google Scholar] [CrossRef]
  22. Li, X.; Bai, Y.; Shi, X.; Su, N.; Nie, G.; Zhang, R.; Nie, H.; Ye, L. Applications of MXene (Ti3C2Tx) in Photocatalysis: A Review. Mater. Adv. 2021, 2, 1570–1594. [Google Scholar] [CrossRef]
  23. Im, J.K.; Sohn, E.J.; Kim, S.; Jang, M.; Son, A.; Zoh, K.-D.; Yoon, Y. Review of MXene-Based Nanocomposites for Photocatalysis. Chemosphere 2021, 270, 129478. [Google Scholar] [CrossRef] [PubMed]
  24. Murali, G.; Reddy Modigunta, J.K.; Park, Y.H.; Lee, J.-H.; Rawal, J.; Lee, S.-Y.; In, I.; Park, S.-J. A Review on MXene Synthesis, Stability, and Photocatalytic Applications. ACS Nano 2022, 16, 13370–13429. [Google Scholar] [CrossRef] [PubMed]
  25. Ning, X.; Hao, A.; Cao, Y.; Chen, R.; Xie, J.; Lu, Z.; Hu, J.; Jia, D. Construction of MXene/Bi2WO6 Schottky Junction for Highly Efficient Piezocatalytic Hydrogen Evolution and Unraveling Mechanism. Nano Lett. 2024, 24, 3361–3368. [Google Scholar] [CrossRef]
  26. Sinopoli, A.; Othman, Z.; Rasool, K.; Mahmoud, K.A. Electrocatalytic/Photocatalytic Properties and Aqueous Media Applications of 2D Transition Metal Carbides (MXenes). Curr. Opin. Solid. State Mater. Sci. 2019, 23, 100760. [Google Scholar] [CrossRef]
  27. Cui, L.; Wen, J.; Deng, Q.; Du, X.; Tang, T.; Li, M.; Xiao, J.; Jiang, L.; Hu, G.; Cao, X.; et al. Improving the Photocatalytic Activity of Ti3C2 MXene by Surface Modification of N Doped. Materials 2023, 16, 2836. [Google Scholar] [CrossRef]
  28. Huang, K.; Li, C.; Li, H.; Ren, G.; Wang, L.; Wang, W.; Meng, X. Photocatalytic Applications of Two-Dimensional Ti3C2 MXenes: A Review. ACS Appl. Nano Mater. 2020, 3, 9581–9603. [Google Scholar] [CrossRef]
  29. Tang, R.; Xiong, S.; Gong, D.; Deng, Y.; Wang, Y.; Su, L.; Ding, C.; Yang, L.; Liao, C. Ti3C2 2D MXene: Recent Progress and Perspectives in Photocatalysis. ACS Appl. Mater. Interfaces 2020, 12, 56663–56680. [Google Scholar] [CrossRef]
  30. Illahi, C.; Hutabarat, W.E.F.; Nurdini, N.; Failamani, F.; Kadja, G.T.M. Photocatalytic Degradation of Azo Dyes over MXene-Based Catalyst: Recent Developments and Future Prospects. Next Nanotechnol. 2024, 6, 100055. [Google Scholar] [CrossRef]
  31. Goel, N.; Kushwaha, A.; Kumar, M. Two-Dimensional MXenes: Recent Emerging Applications. RSC Adv. 2022, 12, 25172–25193. [Google Scholar] [CrossRef]
  32. Kitchamsetti, N.; De Barros, A.L.F. Recent Advances in MXenes Based Composites as Photocatalysts: Synthesis, Properties and Photocatalytic Removal of Organic Contaminants from Wastewater. ChemCatChem 2023, 15, e202300690. [Google Scholar] [CrossRef]
  33. Peng, C.; Zhou, T.; Wei, P.; Xu, W.; Pan, H.; Peng, F.; Jia, J.; Zhang, K.; Yu, H. Photocatalysis over MXene-Based Hybrids: Synthesis, Surface Chemistry, and Interfacial Charge Kinetics. APL Mater. 2021, 9, 070703. [Google Scholar] [CrossRef]
  34. Javaid, A.; Latif, S.; Imran, M.; Hussain, N.; Bilal, M.; Iqbal, H.M.N. MXene-Based Hybrid Composites as Pho-tocatalyst for the Mitigation of Pharmaceuticals. Chemosphere 2022, 291, 133062. [Google Scholar] [CrossRef] [PubMed]
  35. Iravani, S.; Varma, R.S. MXene-Based Photocatalysts in Degradation of Organic and Pharmaceutical Pollutants. Molecules 2022, 27, 6939. [Google Scholar] [CrossRef]
  36. Naguib, M.; Kurtoglu, M.; Presser, V.; Lu, J.; Niu, J.; Heon, M.; Hultman, L.; Gogotsi, Y.; Barsoum, M.W. Two-Dimensional Nanocrystals Produced by Exfoliation of Ti3AlC2. Advanced Materials 2011, 23, 4248–4253. [Google Scholar] [CrossRef]
  37. Ayad, M.M.; El-Nasr, A.A. Anionic Dye (Acid Green 25) Adsorption from Water by Using Polyaniline Nano-tubes Salt/Silica Composite. J. Nanostruct Chem. 2012, 3, 3. [Google Scholar] [CrossRef]
  38. Limbu, T.B.; Chitara, B.; Orlando, J.D.; Garcia Cervantes, M.Y.; Kumari, S.; Li, Q.; Tang, Y.; Yan, F. Green Synthesis of Reduced Ti3C2Tx MXene Nanosheets with Enhanced Conductivity, Oxidation Stability, and SERS Activity. J. Mater. Chem. C 2020, 8, 4722–4731. [Google Scholar] [CrossRef]
  39. Wang, D.; Zhou, C.; Filatov, A.S.; Cho, W.; Lagunas, F.; Wang, M.; Vaikuntanathan, S.; Liu, C.; Klie, R.F.; Ta-lapin, D.V. Direct Synthesis and Chemical Vapor Deposition of 2D Carbide and Nitride MXenes. Science 2023, 379, 1242–1247. [Google Scholar] [CrossRef]
  40. Ayodhya, D. A Review of Recent Progress in 2D MXenes: Synthesis, Properties, and Applications. Diam. Relat. Mater. 2023, 132, 109634. [Google Scholar] [CrossRef]
  41. Lei, J.-C.; Zhang, X.; Zhou, Z. Recent Advances in MXene: Preparation, Properties, and Applications. Front. Phys. 2015, 10, 276–286. [Google Scholar] [CrossRef]
  42. Xu, X.; Zhang, C.; Yin, J.; Smajic, J.; Bahabri, M.; Lei, Y.; Hedhili, M.N.; Hota, M.K.; Shi, L.; Guo, T.; et al. Aniso-tropic Superconducting Nb2CTx MXene Processed by Atomic Exchange at the Wafer Scale. Adv. Mater. 2024, 36, 2305326. [Google Scholar] [CrossRef]
  43. Kamysbayev, V.; Filatov, A.S.; Hu, H.; Rui, X.; Lagunas, F.; Wang, D.; Klie, R.F.; Talapin, D.V. Covalent Surface Modifications and Superconductivity of Two-Dimensional Metal Carbide MXenes. Science 2020, 369, 979–983. [Google Scholar] [CrossRef] [PubMed]
  44. Gogotsi, Y.; Anasori, B. The Rise of MXenes. ACS Nano 2019, 13, 8491–8494. [Google Scholar] [CrossRef] [PubMed]
  45. Verger, L.; Xu, C.; Natu, V.; Cheng, H.-M.; Ren, W.; Barsoum, M.W. Overview of the Synthesis of MXenes and Other Ultrathin 2D Transition Metal Carbides and Nitrides. Curr. Opin. Solid. State Mater. Sci. 2019, 23, 149–163. [Google Scholar] [CrossRef]
  46. Liu, H.; Guo, H.; Gao, Z.; Pan, H.; Zhen, J.; Linghu, J.; Li, Z. Applications of Artificial Intelligence in Materials Research for Fuel Cells. AI Mater. 2025, 1, 1–50. [Google Scholar] [CrossRef]
  47. Alhabeb, M.; Maleski, K.; Anasori, B.; Lelyukh, P.; Clark, L.; Sin, S.; Gogotsi, Y. Guidelines for Synthesis and Processing of Two-Dimensional Titanium Carbide (Ti3C2Tx MXene). Chem. Mater. 2017, 29, 7633–7644. [Google Scholar] [CrossRef]
  48. Koh, S.W.; Rekhi, L.; Arramel; Birowosuto, M.D.; Trinh, Q.T.; Ge, J.; Yu, W.; Wee, A.T.S.; Choksi, T.S.; Li, H. Tuning the Work Function of MXene via Surface Functionalization. ACS Appl. Mater. Interfaces 2024, 16, 66826–66836. [Google Scholar] [CrossRef]
  49. Guo, L.; Jiang, W.-Y.; Shen, M.; Xu, C.; Ding, C.-X.; Zhao, S.-F.; Yuan, T.-T.; Wang, C.-Y.; Zhang, X.-Q.; Wang, J.-Q. High Capacitance of MXene (Ti3C2Tx) Through Intercalation and Surface Modification in Molten Salt. Electrochim. Acta 2022, 401, 139476. [Google Scholar] [CrossRef]
  50. Cao, Z.; Yin, Q.; Zhang, Y.; Li, Y.; Yu, C.; Zhang, M.; Fan, B.; Shao, G.; Wang, H.; Xu, H.; et al. Heterostructure Composites of TiO2 and CdZnS Nanoparticles Decorated on Ti3C2Tx Nanosheets and Their Enhanced Photocatalytic Performance by Microwave Hydrothermal Method. J. Alloys Compd. 2022, 918, 165681. [Google Scholar] [CrossRef]
  51. Purbayanto, M.A.K.; Bury, D.; Chandel, M.; Shahrak, Z.D.; Mochalin, V.N.; Wójcik, A.; Moszczyńska, D.; Wojciechowska, A.; Tabassum, A.; Naguib, M.; et al. Ambient Processed rGO/Ti3CNTx MXene Thin Film with High Oxidation Stability, Photosensitivity, and Self-Cleaning Potential. ACS Appl. Mater. Interfaces 2023, 15, 44075–44086. [Google Scholar] [CrossRef]
  52. Cao, S.; Shen, B.; Tong, T.; Fu, J.; Yu, J. 2D/2D Heterojunction of Ultrathin MXene/Bi2 WO6 Nanosheets for Improved Photocatalytic CO2 Reduction. Adv. Funct. Mater. 2018, 28, 1800136. [Google Scholar] [CrossRef]
  53. Latif, F.E.A.; Khalid, M.; Numan, A.; Manaf, N.A.; Mubarak, N.M.; Zaharin, H.A.; Abdullah, E.C. Micro-Wave-Assisted Hydrothermal Synthesis of Ti3C2Tx MXene: A Sustainable and Scalable Approach Using Alkaline Etchant. J. Mol. Struct. 2025, 1329, 141407. [Google Scholar] [CrossRef]
  54. Saini, H.; Srinivasan, N.; Šedajová, V.; Majumder, M.; Dubal, D.P.; Otyepka, M.; Zbořil, R.; Kurra, N.; Fischer, R.A.; Jayaramulu, K. Emerging MXene@Metal–Organic Framework Hybrids: Design Strategies toward Versatile Applications. ACS Nano 2021, 15, 18742–18776. [Google Scholar] [CrossRef] [PubMed]
  55. Qamar, M.A.; Ali, S.K. Functionalized MXenes for Enhanced Visible-Light Photocatalysis: A Focus on Surface Termination Engineering and Composite Design. Inorganics 2025, 13, 45. [Google Scholar] [CrossRef]
  56. Hong, L.; Guo, R.; Yuan, Y.; Ji, X.; Li, Z.; Lin, Z.; Pan, W. Recent Progress of Two-Dimensional MXenes in Photocatalytic Applications: A Review. Mater. Today Energy 2020, 18, 100521. [Google Scholar] [CrossRef]
  57. My Tran, N.; Thanh Hoai Ta, Q.; Noh, J.-S. Unusual Synthesis of Safflower-Shaped TiO2/Ti3C2 Heterostructures Initiated from Two-Dimensional Ti3C2 MXene. Appl. Surf. Sci. 2021, 538, 148023. [Google Scholar] [CrossRef]
  58. Quyen, V.T.; Ha, L.T.T.; Thanh, D.M.; Le, Q.V.; Viet, N.M.; Nham, N.T.; Thang, P.Q. Advanced Synthesis of MXene-Derived Nanoflower-Shaped TiO2@Ti3C2 Heterojunction to Enhance Photocatalytic Degradation of Rhodamine B. Environ. Technol. Innov. 2021, 21, 101286. [Google Scholar] [CrossRef]
  59. Chen, J.; Zheng, H.; Zhao, Y.; Que, M.; Lei, X.; Zhang, K.; Luo, Y. Preparation of Facet Exposed TiO2/Ti3C2T Composites with Enhanced Photocatalytic Activity. J. Phys. Chem. Solids 2020, 145, 109565. [Google Scholar] [CrossRef]
  60. Ta, Q.T.H.; Tran, N.M.; Noh, J.-S. Rice Crust-Like ZnO/Ti3C2Tx MXene Hybrid Structures for Improved Photocatalytic Activity. Catalysts 2020, 10, 1140. [Google Scholar] [CrossRef]
  61. Cai, T.; Wang, L.; Liu, Y.; Zhang, S.; Dong, W.; Chen, H.; Yi, X.; Yuan, J.; Xia, X.; Liu, C.; et al. Ag3PO4/Ti3C2 MXene Interface Materials as a Schottky Catalyst with Enhanced Photocatalytic Activities and Anti-Photocorrosion Performance. Appl. Catal. B Environ. 2018, 239, 545–554. [Google Scholar] [CrossRef]
  62. Jin Lee, D.; Mohan Kumar, G.; Sekar, S.; Chang Jeon, H.; Young Kim, D.; Ilanchezhiyan, P. Ultrasonic Pro-cessing of WO3 Nanosheets Integrated Ti3C2 MXene 2D-2D Based Heterojunctions with Synergistic Effects for Enhanced Water Splitting and Environmental Remediation. Ultrason. Sonochemistry 2023, 101, 106681. [Google Scholar] [CrossRef]
  63. Alsafari, I.A.; Munir, S.; Zulfiqar, S.; Saif, M.S.; Warsi, M.F.; Shahid, M. Synthesis, Characterization, Photocatalytic and Antibacterial Properties of Copper Ferrite/MXene (CuFe2O4/Ti3C2) Nanohybrids. Ceram. Int. 2021, 47, 28874–28883. [Google Scholar] [CrossRef]
  64. Liu, X.; Liu, Q.; Chen, C. Ultrasonic Oscillation Synthesized ZnS Nanoparticles/Layered MXene Sheet with Outstanding Photocatalytic Activity Under Visible Light. Vacuum 2021, 183, 109834. [Google Scholar] [CrossRef]
  65. Ishfaq, M.; Rasheed, A.; Ajmal, S.; Dastgeer, G.; Naz, T.; Baig, M.M.; Lee, S.G. Synthesis and Characterization of a Novel MnO2@MXene-Based 2D/3D Hierarchical Z-Scheme for Sustainable Environmental Applications. Ceram. Int. 2024, 50, 9801–9810. [Google Scholar] [CrossRef]
  66. Iqbal, M.A.; Tariq, A.; Zaheer, A.; Gul, S.; Ali, S.I.; Iqbal, M.Z.; Akinwande, D.; Rizwan, S. Ti3C2-MXene/Bismuth Ferrite Nanohybrids for Efficient Degradation of Organic Dyes and Colorless Pollutants. ACS Omega 2019, 4, 20530–20539. [Google Scholar] [CrossRef]
  67. Iqbal, M.A.; Ali, S.I.; Amin, F.; Tariq, A.; Iqbal, M.Z.; Rizwan, S. La- and Mn-Codoped Bismuth Ferrite/Ti3C2 MXene Composites for Efficient Photocatalytic Degradation of Congo Red Dye. ACS Omega 2019, 4, 8661–8668. [Google Scholar] [CrossRef]
  68. Zheng, W.; Zhang, P.; Tian, W.; Wang, Y.; Zhang, Y.; Chen, J.; Sun, Z. Microwave-Assisted Synthesis of SnO2-Ti3C2 Nanocomposite for Enhanced Supercapacitive Performance. Mater. Lett. 2017, 209, 122–125. [Google Scholar] [CrossRef]
  69. Zhou, W.; Zhu, J.; Wang, F.; Cao, M.; Zhao, T. One-Step Synthesis of Ceria/Ti3C2 Nanocomposites with En-hanced Photocatalytic Activity. Mater. Lett. 2017, 206, 237–240. [Google Scholar] [CrossRef]
  70. Sajid, M.M.; Khan, S.B.; Javed, Y.; Amin, N.; Zhang, Z.; Shad, N.A.; Zhai, H. Bismuth Vanadate/MXene (Bi-VO4/Ti3C2) Heterojunction Composite: Enhanced Interfacial Control Charge Transfer for Highly Efficient Visible Light Photocatalytic Activity. Env. Sci. Pollut. Res. 2021, 28, 35911–35923. [Google Scholar] [CrossRef]
  71. Fan, Y.; Liu, Z.; Li, Q.; Zhao, K.; Ahmad, M.; Liu, P.; Zhang, Q.; Zhang, B. Preparation of MoS2/MXene/NC Porous Composite Microspheres with Wrinkled Surface and Their Microwave Absorption Performances. ACS Appl. Mater. Interfaces 2023, 15, 41720–41731. [Google Scholar] [CrossRef]
  72. Zhang, S.; Cai, M.; Wu, J.; Wang, Z.; Lu, X.; Li, K.; Lee, J.-M.; Min, Y. Photocatalytic Degradation of TiO2 via Incorporating Ti3C2 MXene for Methylene Blue Removal from Water. Catal. Commun. 2023, 174, 106594. [Google Scholar] [CrossRef]
  73. Othman, Z.; Sinopoli, A.; Mackey, H.R.; Mahmoud, K.A. Efficient Photocatalytic Degradation of Organic Dyes by AgNPs/TiO2/Ti3C2Tx MXene Composites under UV and Solar Light. ACS Omega 2021, 6, 33325–33338. [Google Scholar] [CrossRef] [PubMed]
  74. Ighalo, J.O.; Smith, M.L.; Mayyahi, A.A.; Amama, P.B. MXenes in Photocatalytic NOx Abatement: Current Innovations, Opportunities, and Challenges. Appl. Catal. B Environ. Energy 2024, 358, 124352. [Google Scholar] [CrossRef]
  75. Ahmaruzzaman, M. MXenes and MXene-Supported Nanocomposites: A Novel Materials for Aqueous Environmental Remediation. RSC Adv. 2022, 12, 34766–34789. [Google Scholar] [CrossRef]
  76. Ahmaruzzaman, M. MXene-Based Novel Nanomaterials for Remediation of Aqueous Environmental Pollutants. Inorg. Chem. Commun. 2022, 143, 109705. [Google Scholar] [CrossRef]
  77. Atri, S.; Loni, E.; Zazimal, F.; Hensel, K.; Caplovicova, M.; Plesch, G.; Lu, X.; Nagarajan, R.; Naguib, M.; Monfort, O. MXene-Derived Oxide Nanoheterostructures for Photocatalytic Sulfamethoxazole Degradation. ACS Appl. Nano Mater. 2024, 7, 16506–16515. [Google Scholar] [CrossRef]
  78. Kumar, A.; Dixit, U.; Singh, K.; Gupta, S.P.; Beg, M.S. Structure and Properties of Dyes and Pigments. In Dyes and Pigments-Novel Applications and Waste Treatment; Papadakis, R., Ed.; IntechOpen: London, UK, 2021; ISBN 978-1-83968-614-6. [Google Scholar]
  79. Wang, C.; Ye, J.; Liang, L.; Cui, X.; Kong, L.; Li, N.; Cheng, Z.; Peng, W.; Yan, B.; Chen, G. Application of MXene-Based Materials in Fenton-like Systems for Organic Wastewater Treatment: A Review. Sci. Total Environ. 2023, 862, 160539. [Google Scholar] [CrossRef]
  80. Amrillah, T.; Supandi, A.R.; Puspasari, V.; Hermawan, A.; Seh, Z.W. MXene-Based Photocatalysts and Electrocatalysts for CO2 Conversion to Chemicals. Trans. Tianjin Univ. 2022, 28, 307–322. [Google Scholar] [CrossRef]
  81. Dey, P.C.; Das, R. Enhanced Photocatalytic Degradation of Methyl Orange Dye on Interaction with Synthesized Ligand Free CdS Nanocrystals Under Visible Light Illumination. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2020, 231, 118122. [Google Scholar] [CrossRef]
  82. Medina, J.C.; Bizarro, M.; Silva-Bermudez, P.; Giorcelli, M.; Tagliaferro, A.; Rodil, S.E. Photocatalytic Discoloration of Methyl Orange Dye by δ-Bi2O3 Thin Films. Thin Solid. Film. 2016, 612, 72–81. [Google Scholar] [CrossRef]
  83. 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 Approaches for Environmental Safety. Ecotoxicol. Environ. Saf. 2022, 231, 113160. [Google Scholar] [CrossRef] [PubMed]
  84. Srisai, J.; Muangnapoh, T.; Vas-Umnuay, P. Comparative Study on Photocatalytic Degradation of Methylene Blue Using Pristine ZnO and Ni/ZnO Composite Films. Mater. Today Proc. 2022, 66, 3168–3173. [Google Scholar] [CrossRef]
  85. Mohammad Jafri, N.; Jaafar, J.; Alias, N.; Samitsu, S.; Aziz, F.; Wan Salleh, W.; Mohd Yusop, M.; Othman, M.; Rahman, M.; Ismail, A.; et al. Synthesis and Characterization of Titanium Dioxide Hollow Nanofiber for Photocata-lytic Degradation of Methylene Blue Dye. Membranes 2021, 11, 581. [Google Scholar] [CrossRef] [PubMed]
  86. Seling, T.R.; Katzbaer, R.R.; Thompson, K.L.; Aksoy, S.E.; Chitara, B.; Shringi, A.K.; Schaak, R.E.; Riaz, U.; Yan, F. Transition Metal-Doped CuO Nanosheets for Enhanced Visible-Light Photocatalysis. J. Photochem. Photobiol. A Chem. 2024, 448, 115356. [Google Scholar] [CrossRef]
  87. Seling, T.R.; Kumar Shringi, A.; Wang, K.; Riaz, U.; Yan, F. Bi2O2S Nanosheets for Effective Visible Light Photocatalysis of Anionic Dye Degradation. Mater. Lett. 2024, 361, 136136. [Google Scholar] [CrossRef]
  88. Mary, A.S.; Norbert, A.; Shaji, S.; Philip, R.R. Electrochemically Anodized Solid and Stable ZnO Nanorods as an Adsorbent/Nanophotocatalyst: ROS Mediated Degradation of Azo Dyes Congo Red and Methyl Orange. J. Clean. Prod. 2023, 428, 139466. [Google Scholar] [CrossRef]
  89. Mittal, H.; Khanuja, M. Optimization of MoSe2 Nanostructure by Surface Modification Using Conducting Polymer for Degradation of Cationic and Anionic Dye: Photocatalysis Mechanism, Reaction Kinetics and Intermediate Product Study. Dye. Pigment. 2020, 175, 108109. [Google Scholar] [CrossRef]
  90. Fang, S.; Lv, K.; Li, Q.; Ye, H.; Du, D.; Li, M. Effect of Acid on the Photocatalytic Degradation of Rhodamine B over G-C3N4. Appl. Surf. Sci. 2015, 358, 336–342. [Google Scholar] [CrossRef]
  91. Revathi, B.; Balakrishnan, L.; Pichaimuthu, S.; Nirmala Grace, A.; Krishna Chandar, N. Photocatalytic Degradation of Rhodamine B Using BiMnO3 Nanoparticles under UV and Visible Light Irradiation. J. Mater. Sci. Mater. Electron. 2020, 31, 22487–22497. [Google Scholar] [CrossRef]
  92. Qu, J.; Teng, D.; Zhang, X.; Yang, Q.; Li, P.; Cao, Y. Preparation and Regulation of Two-Dimensional Ti3C2Tx MXene for Enhanced Adsorption–Photocatalytic Degradation of Organic Dyes in Wastewater. Ceram. Int. 2022, 48, 14451–14459. [Google Scholar] [CrossRef]
  93. Zhao, D.; Cai, C. Layered Ti3C2 MXene Modified Two-Dimensional Bi2WO6 Composites with Enhanced Visible Light Photocatalytic Performance. Mater. Chem. Front. 2019, 3, 2521–2528. [Google Scholar] [CrossRef]
  94. Akbari, M.; Rasouli, J.; Rasouli, K.; Ghaedi, S.; Mohammadi, M.; Rajabi, H.; Sabbaghi, S. MXene-Based Compo-site Photocatalysts for Efficient Degradation of Antibiotics in Wastewater. Sci. Rep. 2024, 14, 31498. [Google Scholar] [CrossRef]
  95. Yao, J.; Wang, C. Decolorization of Methylene Blue with TiO2 Sol via UV Irradiation Photocatalytic Degradation. Int. J. Photoenergy 2010, 2010, 1–6. [Google Scholar] [CrossRef]
  96. Kalaiselvi, C.; Krishna Chandar, N. Accordion-like Multilayer Ti3C2Tx MXene Sheets Decorated 1D Mn2O3 Nanorods-Based Nanocomposites: An Efficient Catalyst for Swift Removal of Single and Mixed Dyes. J. Phys. Chem. Solids 2023, 182, 111591. [Google Scholar] [CrossRef]
  97. Chandiran, K.; Pandian, M.S.; Balakrishnan, S.; Pitchaimuthu, S.; Chen, Y.-S.; Nagamuthu Raja, K.C. Ti3C2Tx MXene Decorated with NiMnO3/NiMn2O4 Nanoparticles for Simultaneous Photocatalytic Degradation of Mixed Cationic and Anionic Dyes. Colloids Surf. A Physicochem. Eng. Asp. 2024, 692, 133888. [Google Scholar] [CrossRef]
  98. Wang, Y.; Chen, J.; Que, M.; Wu, Q.; Wang, X.; Zhou, Y.; Ma, Y.; Li, Y.; Yang, X. MXene-Derived Ti3C2Tx/Bi4Ti3O12 Heterojunction Photocatalyst for Enhanced Degradation of Tetracycline Hydrochloride, Rho-damine B, and Methyl Orange under Visible-Light Irradiation. Appl. Surf. Sci. 2023, 639, 158270. [Google Scholar] [CrossRef]
  99. Jiao, S.; Liu, L. Friction-Induced Enhancements for Photocatalytic Degradation of MoS2@Ti3C2 Nanohybrid. Ind. Eng. Chem. Res. 2019, 58, 18141–18148. [Google Scholar] [CrossRef]
  100. Lu, Y.; Yao, M.; Zhou, A.; Hu, Q.; Wang, L. Preparation and Photocatalytic Performance of Ti3C2/TiO2/CuO Nanocomposites. J. Nanomater. 2017, 2017, 1–5. [Google Scholar] [CrossRef]
  101. Chen, L.; Ye, X.; Chen, S.; Ma, L.; Wang, Z.; Wang, Q.; Hua, N.; Xiao, X.; Cai, S.; Liu, X. Ti3C2 MXene Nanosheet/TiO2 Composites for Efficient Visible Light Photocatalytic Activity. Ceram. Int. 2020, 46, 25895–25904. [Google Scholar] [CrossRef]
  102. Ayub, A.; Kim, B.; Lim, Y.; Devarayapalli, K.C.; Kim, G.; Lee, D.S. Hydrothermal Synthesis of Cobalt Ferrite-Functionalized Ti3C2Tx MXene for the Degradation of Congo Red via Peroxymonosulfate Activation System. J. Alloys Compd. 2023, 963, 171294. [Google Scholar] [CrossRef]
  103. Tariq, A.; Ali, S.I.; Akinwande, D.; Rizwan, S. Efficient Visible-Light Photocatalysis of 2D-MXene Nanohybrids with Gd3+- and Sn4+ -Codoped Bismuth Ferrite. ACS Omega 2018, 3, 13828–13836. [Google Scholar] [CrossRef] [PubMed]
  104. Xu, T.; Wang, J.; Cong, Y.; Jiang, S.; Zhang, Q.; Zhu, H.; Li, Y.; Li, X. Ternary BiOBr/TiO2/Ti3C2T MXene Nano-composites with Heterojunction Structure and Improved Photocatalysis Performance. Chin. Chem. Lett. 2020, 31, 1022–1025. [Google Scholar] [CrossRef]
  105. Lemos, H.G.; Ronchi, R.M.; Portugal, G.R.; Rossato, J.H.H.; Selopal, G.S.; Barba, D.; Venancio, E.C.; Rosei, F.; Arantes, J.T.; Santos, S.F. Efficient Ti3C2Tx MXene/TiO2 Hybrid Photoanodes for Dye-Sensitized Solar Cells. ACS Appl. Energy Mater. 2022, 5, 15928–15938. [Google Scholar] [CrossRef]
  106. Yang, J.; Zhu, Q.; Xie, Z.; Wang, Y.; Wang, J.; Peng, Y.; Fang, Y.; Deng, L.; Xie, T.; Xu, L. Enhancement Mechanism of Photocatalytic Activity for MoS2/Ti3C2 Schottky Junction: Experiment and DFT Calculation. J. Alloys Compd. 2021, 887, 161411. [Google Scholar] [CrossRef]
  107. Liu, D.; Li, C.; Ge, J.; Zhao, C.; Zhao, Q.; Zhang, F.; Ni, T.; Wu, W. 3D Interconnected G-C3N4 Hybridized with 2D Ti3C2 MXene Nanosheets for Enhancing Visible Light Photocatalytic Hydrogen Evolution and Dye Contaminant Elimination. Appl. Surf. Sci. 2022, 579, 152180. [Google Scholar] [CrossRef]
  108. Cheng, X.; Liao, J.; Xue, Y.; Lin, Q.; Yang, Z.; Yan, G.; Zeng, G.; Sengupta, A. Ultrahigh-Flux and Self-Cleaning Composite Membrane Based on BiOCl-PPy Modified MXene Nanosheets for Contaminants Removal from Wastewater. J. Membr. Sci. 2022, 644, 120188. [Google Scholar] [CrossRef]
  109. Wang, B.; Lin, L.; Chen, Y.; Yang, Q.; Xiong, Y.; Zhang, L.; Dai, X.; Jiang, Y.; Zhong, C.; Liao, J.; et al. Rational Construction of S-Scheme Pt-MnO2/TiO2@Ti3C2Tx via Ti-O-Mn Bond for Distinguished Charge Transfer in Photocatalytic Wastewater Environmental Governance and Hydrogen Production. Compos. Sci. Technol. 2023, 241, 110137. [Google Scholar] [CrossRef]
  110. Lin, Q.; Zeng, G.; Yan, G.; Luo, J.; Cheng, X.; Zhao, Z.; Li, H. Self-Cleaning Photocatalytic MXene Composite Membrane for Synergistically Enhanced Water Treatment: Oil/Water Separation and Dyes Removal. Chem. Eng. J. 2022, 427, 131668. [Google Scholar] [CrossRef]
  111. Chen, C.; Wang, B.; Xu, J.; Fei, L.; Raza, S.; Li, B.; Zeng, Q.; Shen, L.; Lin, H. Recent Advancement in Emerging MXene-Based Photocatalytic Membrane for Revolutionizing Wastewater Treatment. Small 2024, 20, 2311427. [Google Scholar] [CrossRef]
  112. Khadidja, M.F.; Fan, J.; Li, S.; Li, S.; Cui, K.; Wu, J.; Zeng, W.; Wei, H.; Jin, H.-G.; Naik, N.; et al. Hierarchical ZnO/MXene Composites and Their Photocatalytic Performances. Colloids Surf. A Physicochem. Eng. Asp. 2021, 628, 127230. [Google Scholar] [CrossRef]
  113. Xue, H.; Yan, Q.; Chen, L.; Wang, Y.; Xie, X.; Sun, J. Ti3C2 MXene Assembled with TiO2 for Efficient Photocatalytic Mineralization of Gaseous O-Xylene. Appl. Surf. Sci. 2023, 608, 155136. [Google Scholar] [CrossRef]
  114. Hieu, V.Q.; Phung, T.K.; Nguyen, T.-Q.; Khan, A.; Doan, V.D.; Tran, V.A.; Le, V.T. Photocatalytic Degradation of Methyl Orange Dye by Ti3C2–TiO2 Heterojunction Under Solar Light. Chemosphere 2021, 276, 130154. [Google Scholar] [CrossRef] [PubMed]
  115. Mousavi, S.M.; Mohtaram, M.S.; Rasouli, K.; Mohtaram, S.; Rajabi, H.; Sabbaghi, S. Efficient Visi-ble-Light-Driven Photocatalytic Degradation of Antibiotics in Water by MXene-Derived TiO2-Supported SiO2/Ti3C2 Composites: Optimisation, Mechanism and Toxicity Evaluation. Environ. Pollut. 2025, 367, 125624. [Google Scholar] [CrossRef] [PubMed]
  116. Huang, G.; Li, S.; Liu, L.; Zhu, L.; Wang, Q. Ti3C2 MXene-Modified Bi2WO6 Nanoplates for Efficient Photodegradation of Volatile Organic Compounds. Appl. Surf. Sci. 2020, 503, 144183. [Google Scholar] [CrossRef]
  117. Zhou, X.; Liu, G.; Yu, J.; Fan, W. Surface Plasmon Resonance-Mediated Photocatalysis by Noble Metal-Based Composites Under Visible Light. J. Mater. Chem. 2012, 22, 21337. [Google Scholar] [CrossRef]
  118. Qin, J.; Hu, X.; Li, X.; Yin, Z.; Liu, B.; Lam, K. 0D/2D AgInS2/MXene Z-Scheme Heterojunction Nanosheets for Improved Ammonia Photosynthesis of N2. Nano Energy 2019, 61, 27–35. [Google Scholar] [CrossRef]
  119. Yusuf, B.O.; Umar, M.; Aliyu, M.; Alhassan, A.M.; Awad, M.M.; Taialla, O.A.; Abdullahi, A.S.; Musa, J.N.; Al-hooshani, K.R.; Ganiyu, S.A. Recent Advances and Future Prospects of MXene-Based Photocatalysts in Environmental Remediations. J. Environ. Chem. Eng. 2024, 12, 114812. [Google Scholar] [CrossRef]
  120. Hassaan, M.A.; El-Nemr, M.A.; Elkatory, M.R.; Ragab, S.; Niculescu, V.-C.; El Nemr, A. Principles of Photocatalysts and Their Different Applications: A Review. Top. Curr. Chem. (Z) 2023, 381, 31. [Google Scholar] [CrossRef]
  121. Kuang, P.; Low, J.; Cheng, B.; Yu, J.; Fan, J. MXene-Based Photocatalysts. J. Mater. Sci. Technol. 2020, 56, 18–44. [Google Scholar] [CrossRef]
  122. Parwaiz, S.; Sohn, Y.; Khan, M.M. Insights into MXenes and MXene-Based Heterostructures for Various Photocatalytic Applications. Mater. Sci. Semicond. Process. 2025, 186, 109099. [Google Scholar] [CrossRef]
  123. You, Z.; Liao, Y.; Li, X.; Fan, J.; Xiang, Q. State-of-the-Art Recent Progress in MXene-Based Photocatalysts: A Comprehensive Review. Nanoscale 2021, 13, 9463–9504. [Google Scholar] [CrossRef] [PubMed]
  124. Kumar, A.; Majithia, P.; Choudhary, P.; Mabbett, I.; Kuehnel, M.F.; Pitchaimuthu, S.; Krishnan, V. MXene Coupled Graphitic Carbon Nitride Nanosheets Based Plasmonic Photocatalysts for Removal of Pharmaceutical Pollutant. Chemosphere 2022, 308, 136297. [Google Scholar] [CrossRef] [PubMed]
  125. Diao, Y.; Yan, M.; Li, X.; Zhou, C.; Peng, B.; Chen, H.; Zhang, H. In-Situ Grown of g-C3N4/Ti3C2/TiO2 Nanotube Arrays on Ti Meshes for Efficient Degradation of Organic Pollutants Under Visible Light Irradiation. Colloids Surf. A Physicochem. Eng. Asp. 2020, 594, 124511. [Google Scholar] [CrossRef]
  126. Zhou, Y.; Yu, M.; Zhan, R.; Wang, X.; Peng, G.; Niu, J. Ti3C2 MXene-Induced Interface Electron Separation in g-C3N4/Ti3C2 MXene/MoSe2 Z-Scheme Heterojunction for Enhancing Visible Light-Irradiated Enoxacin Degrada-tion. Sep. Purif. Technol. 2021, 275, 119194. [Google Scholar] [CrossRef]
  127. Rdewi, E.H.; Abbas, K.K.; AbdulkadhimAl-Ghaban, A.M.H. Removal Pharmaceutical Carbamazepine from Wastewater Using ZnO-TiO2-MXene Heterostructural Nanophotocatalyst Under Solar Light Irradiation. Mater. Today Proc. 2022, 60, 1702–1711. [Google Scholar] [CrossRef]
  128. Abbas, K.K.; AbdulkadhimAl-Ghaban, A.M.H.; Rdewi, E.H. Synthesis of a Novel ZnO/TiO2-Nanorod MXene Heterostructured Nanophotocatalyst for the Removal Pharmaceutical Ceftriaxone Sodium from Aqueous Solution under Simulated Sunlight. J. Environ. Chem. Eng. 2022, 10, 108111. [Google Scholar] [CrossRef]
  129. Sukidpaneenid, S.; Chawengkijwanich, C.; Pokhum, C.; Isobe, T.; Opaprakasit, P.; Sreearunothai, P. Multi-Function Adsorbent-Photocatalyst MXene-TiO2 Composites for Removal of Enrofloxacin Antibiotic from Water. J. Environ. Sci. 2023, 124, 414–428. [Google Scholar] [CrossRef]
  130. Shahzad, A.; Rasool, K.; Nawaz, M.; Miran, W.; Jang, J.; Moztahida, M.; Mahmoud, K.A.; Lee, D.S. Hetero-structural TiO2/Ti3C2Tx (MXene) for Photocatalytic Degradation of Antiepileptic Drug Carbamazepine. Chem. Eng. J. 2018, 349, 748–755. [Google Scholar] [CrossRef]
  131. Mohanty, S.; Sharma, M.; Kumar, A.; Krishnan, V. Hot Electron-Mediated Photocatalytic Degradation of Ciprofloxacin Using Au-Decorated SrTiO3-and Ti3C2 MXene-Based Interfacial Heterostructure Nanoarchitectonics. J. Phys. Chem. C 2023, 127, 17711–17722. [Google Scholar] [CrossRef]
  132. Du, X.; Ye, L.; Zhu, J.; Ye, Y.; Wang, A.; Zhang, H.; Xu, Z.; Dai, L.; Wang, Y. Novel Cerium (IV) Oxide-Ti3C2-Titanium Dioxide Heterostructure Photocatalyst for Pharmaceutical Pollutants Removal: Photocatalyst Characterization, Process Optimization and Transformation Pathways. Surf. Interfaces 2024, 46, 103892. [Google Scholar] [CrossRef]
  133. Sergiienko, S.A.; Tobaldi, D.M.; Lajaunie, L.; Lopes, D.V.; Constantinescu, G.; Shaula, A.L.; Shcherban, N.D.; Shkepu, V.I.; Calvino, J.J.; Frade, J.R.; et al. Correction: Photocatalytic Removal of Benzene over Ti3C2Tx MXene and TiO2–MXene Composite Materials Under Solar and NIR Irradiation. J. Mater. Chem. C 2023, 11, 5225. [Google Scholar] [CrossRef]
  134. Mo, H.; Wang, Y. A Bionic Solar-Driven Interfacial Evaporation System with a Photothermal-Photocatalytic Hydrogel for VOC Removal during Solar Distillation. Water Res. 2022, 226, 119276. [Google Scholar] [CrossRef] [PubMed]
  135. Mishra, R.P.; Mrinalini, M.; Kumar, N.; Bastia, S.; Chaudhary, Y.S. Efficient Photocatalytic CO2 Reduction with High Selectivity for Ethanol by Synergistically Coupled MXene-Ceria and the Charge Carrier Dynamics. Langmuir 2023, 39, 14189–14203. [Google Scholar] [CrossRef] [PubMed]
  136. Sun, Y.; Meng, X.; Dall’Agnese, Y.; Dall’Agnese, C.; Duan, S.; Gao, Y.; Chen, G.; Wang, X.-F. 2D MXenes as Co-Catalysts in Photocatalysis: Synthetic Methods. Nano-Micro Lett. 2019, 11, 79. [Google Scholar] [CrossRef]
  137. Li, J.; Wang, Y.; Wang, Y.; Guo, Y.; Zhang, S.; Song, H.; Li, X.; Gao, Q.; Shang, W.; Hu, S.; et al. MXene Ti3C2 Decorated G-C3N4/ZnO Photocatalysts with Improved Photocatalytic Performance for CO2 Reduction. Nano Mater. Sci. 2023, 5, 237–245. [Google Scholar] [CrossRef]
  138. Irfan, M.; Ahmad, I.; Shukrullah, S.; Hussain, H.; Atif, M.; Legutko, S.; Petru, J.; Hatala, M.; Naz, M.Y.; Rahman, S. Construction of 0D/2D Schottky Heterojunctions of ZnO and Ti3C2 Nanosheets with the Enriched Transfer of Interfacial Charges for Photocatalytic Hydrogen Evolution. Materials 2022, 15, 4557. [Google Scholar] [CrossRef]
  139. Colmenares, J.; Kuna, E. Photoactive Hybrid Catalysts Based on Natural and Synthetic Polymers: A Comparative Overview. Molecules 2017, 22, 790. [Google Scholar] [CrossRef]
  140. Peng, C.; Yang, X.; Li, Y.; Yu, H.; Wang, H.; Peng, F. Hybrids of Two-Dimensional Ti3C2 and TiO2 Exposing {001} Facets toward Enhanced Photocatalytic Activity. ACS Appl. Mater. Interfaces 2016, 8, 6051–6060. [Google Scholar] [CrossRef]
  141. Peng, Y.; Cai, P.; Yang, L.; Liu, Y.; Zhu, L.; Zhang, Q.; Liu, J.; Huang, Z.; Yang, Y. Theoretical and Experimental Studies of Ti3C2 MXene for Surface-Enhanced Raman Spectroscopy-Based Sensing. ACS Omega 2020, 5, 26486–26496. [Google Scholar] [CrossRef]
  142. Hosseini, S.F.; Seyed Dorraji, M.S.; Rasoulifard, M.H. Boosting Photo-Charge Transfer in 3D/2D TiO2@Ti3C2 MXene/Bi2S3 Schottky/Z-Scheme Heterojunction for Photocatalytic Antibiotic Degradation and H2 Evolution. Compos. Part B Eng. 2023, 262, 110820. [Google Scholar] [CrossRef]
  143. Wang, D.; Zhao, L.; Song, D.; Qiu, J.; Kong, F.; Guo, L.-H. A Formation Model of Superoxide Radicals Photogenerated in Nano-TiO2 Suspensions. RSC Adv. 2019, 9, 29429–29432. [Google Scholar] [CrossRef] [PubMed]
  144. Farghaly, A.; Maher, E.; Gad, A.; El-Bery, H. Synergistic Photocatalytic Degradation of Methylene Blue Using TiO2 Composites with Activated Carbon and Reduced Graphene Oxide: A Kinetic and Mechanistic Study. Appl. Water Sci. 2024, 14, 228. [Google Scholar] [CrossRef]
  145. Islam Molla, M.A.; Tateishi, I.; Furukawa, M.; Katsumata, H.; Suzuki, T.; Kaneco, S. Evaluation of Reaction Mechanism for Photocatalytic Degradation of Dye with Self-Sensitized TiO2 Under Visible Light Irradiation. OJINM 2017, 7, 1–7. [Google Scholar] [CrossRef]
  146. Kumar Mandal, R.; Ghosh, S.; Pal Majumder, T. Comparative Study between Degradation of Dyes (MB, MO) in Monotonous and Binary Solution Employing Synthesized Bimetallic (Fe-CdO) NPs Having Antioxidant Property. Results Chem. 2023, 5, 100788. [Google Scholar] [CrossRef]
  147. Cao, F.; Zhang, Y.; Wang, H.; Khan, K.; Tareen, A.K.; Qian, W.; Zhang, H.; Ågren, H. Recent Advances in Oxidation Stable Chemistry of 2D MXenes. Adv. Mater. 2022, 34, 2107554. [Google Scholar] [CrossRef]
  148. Lim, K.R.G.; Shekhirev, M.; Wyatt, B.C.; Anasori, B.; Gogotsi, Y.; Seh, Z.W. Fundamentals of MXene Synthesis. Nat. Synth. 2022, 1, 601–614. [Google Scholar] [CrossRef]
  149. Mashtalir, O.; Naguib, M.; Dyatkin, B.; Gogotsi, Y.; Barsoum, M.W. Kinetics of Aluminum Extraction from Ti3AlC2 in Hydrofluoric Acid. Mater. Chem. Phys. 2013, 139, 147–152. [Google Scholar] [CrossRef]
  150. Wei, Y.; Zhang, P.; Soomro, R.A.; Zhu, Q.; Xu, B. Advances in the Synthesis of 2D MXenes. Adv. Mater. 2021, 33, 2103148. [Google Scholar] [CrossRef]
  151. Xu, D.; Li, Z.; Li, L.; Wang, J. Insights into the Photothermal Conversion of 2D MXene Nanomaterials: Synthesis, Mechanism, and Applications. Adv. Funct. Mater. 2020, 30, 2000712. [Google Scholar] [CrossRef]
  152. Jin, C.; Sun, D.; Sun, Z.; Rao, S.; Wu, Z.; Cheng, C.; Liu, L.; Liu, Q.; Yang, J. Interfacial Engineering of Ni-Phytate and Ti3C2Tx MXene-Sensitized TiO2 toward Enhanced Sterilization Efficacy under 808 Nm NIR Light Irradiation. Appl. Catal. B Environ. 2023, 330, 122613. [Google Scholar] [CrossRef]
  153. Lu, B.; Xia, Y.; Ren, Y.; Xie, M.; Zhou, L.; Vinai, G.; Morton, S.A.; Wee, A.T.S.; Van Der Wiel, W.G.; Zhang, W.; et al. When Machine Learning Meets 2D Materials: A Review. Adv. Sci. 2024, 11, 2305277. [Google Scholar] [CrossRef]
  154. Malakar, P.; Thakur, M.S.H.; Nahid, S.M.; Islam, M.M. Data-Driven Machine Learning to Predict Mechanical Properties of Monolayer Transition-Metal Dichalcogenides for Applications in Flexible Electronics. ACS Appl. Nano Mater. 2022, 5, 16489–16499. [Google Scholar] [CrossRef]
  155. Anandhi, G.; Iyapparaja, M. Photocatalytic Degradation of Drugs and Dyes Using a Maching Learning Approach. RSC Adv. 2024, 14, 9003–9019. [Google Scholar] [CrossRef] [PubMed]
  156. Javed, M.F.; Shahab, M.Z.; Asif, U.; Najeh, T.; Aslam, F.; Ali, M.; Khan, I. Evaluation of Machine Learning Models for Predicting TiO2 Photocatalytic Degradation of Air Contaminants. Sci. Rep. 2024, 14, 13688. [Google Scholar] [CrossRef] [PubMed]
  157. Ahmed, Y.; Dutta, K.R.; Nepu, S.N.C.; Prima, M.; AlMohamadi, H.; Akhtar, P. Optimizing Photocatalytic Dye Degradation: A Machine Learning and Metaheuristic Approach for Predicting Methylene Blue in Contaminated Water. Results Eng. 2025, 25, 103538. [Google Scholar] [CrossRef]
Figure 1. Schematic illustration of the TiO2/Ti3C2 hybrid synthesis process through partial oxidation of Ti3C2. Reproduced with permission from [57].
Figure 1. Schematic illustration of the TiO2/Ti3C2 hybrid synthesis process through partial oxidation of Ti3C2. Reproduced with permission from [57].
Molecules 30 01463 g001
Figure 2. Schematic illustration of the ZnO/Ti3C2Tx hybrid photocatalyst synthesis process. Reproduced with permission from [60].
Figure 2. Schematic illustration of the ZnO/Ti3C2Tx hybrid photocatalyst synthesis process. Reproduced with permission from [60].
Molecules 30 01463 g002
Figure 3. Schematic illustration of the Ti3C2Tx MXene/g-C3N4 photocatalyst synthesis process using the wet impregnation method. Reproduced with permission from [6].
Figure 3. Schematic illustration of the Ti3C2Tx MXene/g-C3N4 photocatalyst synthesis process using the wet impregnation method. Reproduced with permission from [6].
Molecules 30 01463 g003
Figure 4. Schematic illustration of the 2D/2D WO3/Ti3C2Tx heterojunction formation process. Reproduced with permission from [62].
Figure 4. Schematic illustration of the 2D/2D WO3/Ti3C2Tx heterojunction formation process. Reproduced with permission from [62].
Molecules 30 01463 g004
Figure 5. Flow chart of the synthesis process for BiVO4/Ti3C2Tx nanocomposite [70].
Figure 5. Flow chart of the synthesis process for BiVO4/Ti3C2Tx nanocomposite [70].
Molecules 30 01463 g005
Figure 6. Schematic of the synthesis and fabrication of MoS2/Ti3C2Tx MXene/N-doped carbon composite microspheres. Reproduced with permission from [71].
Figure 6. Schematic of the synthesis and fabrication of MoS2/Ti3C2Tx MXene/N-doped carbon composite microspheres. Reproduced with permission from [71].
Molecules 30 01463 g006
Figure 8. Proposed mechanisms of photocatalytic degradation: (a) anionic dye (MO), reproduced with permission from [114], and (b) cationic dye (MB) over TiO2/Ti3C2Tx composite, reproduced with permission from [72].
Figure 8. Proposed mechanisms of photocatalytic degradation: (a) anionic dye (MO), reproduced with permission from [114], and (b) cationic dye (MB) over TiO2/Ti3C2Tx composite, reproduced with permission from [72].
Molecules 30 01463 g008
Table 1. Selected nonhybrid photocatalysts with photocatalytic degradation performance towards cationic and anionic dyes.
Table 1. Selected nonhybrid photocatalysts with photocatalytic degradation performance towards cationic and anionic dyes.
Nonhybrid
Photocatalysts
DyesType of DyesDegradation
Percentage (%)
References
CdSMOAnionic95 (300 min)[81]
δ-Bi2O3MOAnionic98 (180 min)[82]
TiO2 NPsMOAnionic~95 (~120 min)[83]
ZnOMBCationic40.88 (21 h)[84]
TiO2 Hollow NanofiberMBCationic95.2 (4 h)[85]
CuOMBCationic62 (270 min)[86]
Bi2S2O3CRAnionic82 (75 min)[87]
ZnOCRAnionic97.6 (75 min)[88]
MoSe2CRAnionic8.44 (120 min)[89]
g-C3N4RhBCationic75 (180 min)[90]
BiMnO3RhBCationic68 (75 min)[91]
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

Seling, T.R.; Songsart-Power, M.; Shringi, A.K.; Paudyal, J.; Yan, F.; Limbu, T.B. Ti3C2Tx MXene-Based Hybrid Photocatalysts in Organic Dye Degradation: A Review. Molecules 2025, 30, 1463. https://doi.org/10.3390/molecules30071463

AMA Style

Seling TR, Songsart-Power M, Shringi AK, Paudyal J, Yan F, Limbu TB. Ti3C2Tx MXene-Based Hybrid Photocatalysts in Organic Dye Degradation: A Review. Molecules. 2025; 30(7):1463. https://doi.org/10.3390/molecules30071463

Chicago/Turabian Style

Seling, Tank R., Mackenzie Songsart-Power, Amit Kumar Shringi, Janak Paudyal, Fei Yan, and Tej B. Limbu. 2025. "Ti3C2Tx MXene-Based Hybrid Photocatalysts in Organic Dye Degradation: A Review" Molecules 30, no. 7: 1463. https://doi.org/10.3390/molecules30071463

APA Style

Seling, T. R., Songsart-Power, M., Shringi, A. K., Paudyal, J., Yan, F., & Limbu, T. B. (2025). Ti3C2Tx MXene-Based Hybrid Photocatalysts in Organic Dye Degradation: A Review. Molecules, 30(7), 1463. https://doi.org/10.3390/molecules30071463

Article Metrics

Back to TopTop