Next Article in Journal
The Role of TEMPO/NaBr/NaClO in Hemp Fiber Oxidation: Deciphering the Mechanism and Reaction Kinetics
Previous Article in Journal
Drilling Surface Quality Analysis of Carbon Fiber-Reinforced Polymers Based on Acoustic Emission Characteristics
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

MXene-Polymer Nanocomposites for High-Efficiency Photocatalytic Antibiotic Degradation Review: Microstructure Control, Environmental Adaptability and Future Prospects

1
College of Environmental Science and Engineering, Beijing Forestry University, Beijing 100083, China
2
Institute of Resources and Environment, Beijing Academy of Science and Technology, Beijing 100089, China
*
Author to whom correspondence should be addressed.
Polymers 2025, 17(19), 2630; https://doi.org/10.3390/polym17192630
Submission received: 25 August 2025 / Revised: 18 September 2025 / Accepted: 26 September 2025 / Published: 28 September 2025
(This article belongs to the Special Issue Photoelectrocatalytic Polymer Materials)

Abstract

The efficient degradation of antibiotics in pharmaceutical wastewater remains a critical challenge against environmental contaminants. Conventional photocatalysts face potential limitations such as narrow visible-light absorption, rapid carrier recombination, and reliance on precious metal cocatalysts. This review investigates the coordination structure of MXene as a cocatalyst to synergistically enhance photocatalytic antibiotic degradation efficiency and the coordination structure modification mechanisms. MXene’s tunable bandgap (0.92–1.75 eV), exceptional conductivity (100–20,000 S/cm), and abundant surface terminations (-O, -OH, -F) enable the construction of Schottky or Z-scheme heterojunctions with semiconductors (Cu2O, TiO2, g-C3N4), achieving 50–70% efficiency improvement compared to pristine semiconductors. The “electron sponge” effect of MXene suppresses electron-hole recombination by 3–5 times, while its surface functional groups dynamically optimize pollutant adsorption. Notably, MXene’s localized surface plasmon resonance extends light harvesting from visible (400–800 nm) to near-infrared regions (800–2000 nm), tripling photon utilization efficiency. Theoretical simulations demonstrate that d-orbital electronic configurations and terminal groups cooperatively regulate catalytic active sites at atomic scales. The MXene composites demonstrate remarkable environmental stability, maintaining over 90% degradation efficiency of antibiotic under high salinity (2 M NaCl) and broad pH range (4–10). Future research should prioritize green synthesis protocols and mechanistic investigations of interfacial dynamics in multicomponent wastewater systems to facilitate engineering applications. This work provides fundamental insights into designing MXene-based photocatalysts for sustainable water purification.

Graphical Abstract

1. Introduction

Antibiotic pollution constitutes a significant environmental concern, characterized by diverse sources, complex compositions, and potential ecological toxicity. In the environment, certain antibiotics (e.g., tetracyclines and fluoroquinolones) can occur at high concentrations while exhibiting high mobility and resistance to biodegradation. Notably, over 60% of antibiotics discharged globally originate from veterinary and agricultural sources (classified by primary use as human, veterinary and agricultural medicine), constituting a major threat to surface and groundwater ecosystems, as reported by the World Health Organization [1]. Confronted with this challenge, conventional remediation technologies (e.g., adsorption and biodegradation) exhibit substantial limitations in the removal of antibiotics. For instance, biological treatment processes are easily interfered with by coexisting organic compounds in wastewater and affected by microbial inhibition caused by high salinity. Therefore, photocatalytic oxidation has emerged as a promising alternative, offering mild reaction conditions and complete mineralization capabilities [2,3,4]. However, traditional photocatalysts (TiO2) suffer from low visible-light utilization due to wide bandgaps, while noble metal cocatalysts (Au, Ag, Pt) are hindered by high costs and poor selectivity in complex wastewater matrices [5]. Recent advances in semiconductor composites (elemental doping, heterojunction engineering) partially address these issues but still grapple with insufficient active sites and rapid electron-hole recombination (Table 1). Meanwhile, the precise dimensional control of 2D layered materials (1–100 nm) enabled by nanotechnology has unlocked exceptional catalytic potential, driven by their high surface area and tunable physicochemical properties. However, conventional two-dimensional layered materials (e.g., graphene) have inherent limitations in catalytic applications. They exhibit weak interfacial adhesion with semiconductors due to van der Waals or electrostatic interaction [6], show impaired electronic conductivity upon oxidation [7], and destabilized electronic structures during reduction processes [8]. These limitations restrict the effectiveness of traditional 2D layered materials in complex catalytic environments. Therefore, it is urgent to develop advanced alternative materials with greater stability and higher functionality.
MXenes—transition metal carbides, nitrides and carbonitrides—were first discovered by Barsoum and Gogotsi at Drexel University in 2011 [18]. MXenes have already found wide application potential beyond environmental photocatalysis—in areas such as biomedical engineering, biosensing [19], lithium battery [20], sensor [21], and wearable electronics [22]—thanks to their ultrathin layered structure, large surface area, rich surface chemistry, and good biocompatibility. With tunable coordination structures, high conductivity, and diverse surface functionalities, MXene offers a promising approach for developing photocatalysis to degrade antibiotics. The general chemical formula of MXenes is Mn+1XnTx, where M represents a transition metal (e.g., Ti, Mo, V); X stands for carbon (C), nitrogen (N), or their combination (carbonitride); and Tx refers to surface functional groups (e.g., –OH, –O, –F) resulting from the chemical etching process during synthesis.
MXenes overcome the limitations of traditional 2D materials through their inherent metallic conductivity, chemically tunable surface terminations (=OH, -O, -F), and strong interfacial coupling with semiconductors via covalent bonding. The strategic modulation of MXene terminal groups and interlayer metal coordination enables precise design of adsorption-catalysis synergies. The interfacial charge transfer efficiency in this heterojunction is 3.2 times greater than that of individual components, providing a robust platform for complex wastewater treatment.
Furthermore, confined heterojunctions between MXene and semiconductors (e.g., g-C3N4, BiOBr) minimize competitive adsorption from coexisting ions (Cl, SO42−), enabling high-efficiency, interference-resistant catalysis. This multifunctional capability positions MXenes as superior cocatalytic platforms for advanced photocatalytic systems, demonstrating significant potential in antibiotics wastewater treatment (Figure 1) [23]. MXene, as an emerging and exciting material, has attracted increasing attention in recent years. A growing number of studies have focused on exploring the role of MXenes as fillers in polymer nanocomposites, aiming to analyze the characteristics of these composites under different fabrication methods and to elucidate the adaptability of their functions in various MXene–polymer nanocomposite systems [24].
In this study, we address several key research gaps in the field: (1) By focusing on MXene-assisted photocatalysis, we systematically explore how MXene enhances the efficiency of photocatalytic antibiotic degradation, overcoming the limitations of traditional photocatalysts through its unique properties. (2) We establish a clear pathway linking the MXene-polymer structure to photocatalytic performance and, ultimately, to antibiotic degradation efficiency. (3) By identifying key directions for future material optimization and interdisciplinary integration, this work aims to foster advancements in related subfields and bridge the gap between fundamental research on MXene-polymers and practical applications for antibiotic degradation.

2. Functional Roles of MXene in Polymer Nanocomposites

2.1. MXene Synthetic Regulation

MXenes significantly enhance photocatalytic efficiency owing to their large, exposed surface area, abundant active sites, and short charge-transfer distances, with these characteristics governed by synthesis protocols (Figure 2). The synthesis of MXenes has generally been categorized into three main strategies: hydrofluoric acid etching, fluoride-free etching, and direct synthesis. While both fluoride-free and direct synthesis routes remain largely at the exploratory stage, hydrofluoric acid etching continues to represent the most widely adopted and reliable method for MXene preparation [32,33]. Notably, the HF-based etching approach can be further divided into two distinct routes, namely direct etching with hydrofluoric acid and the in situ generation of HF from fluoride salts, each offering different levels of safety, controllability, and efficiency.
Two principal etching strategies dominate MXene fabrication: (1) direct etching of MAX phases using concentrated hydrofluoric acid (HF) at ambient conditions, followed by ultrasonic delamination in isopropanol/methanol mixtures to yield multilayer MXenes, the SEM images are shown in Figure 3A,B; (2) controlled etching via in situ HF generation from fluoride salts (e.g., LiF, KF) in hydrochloric acid, the SEM images are shown in Figure 3C,D. Initial MXene products typically exhibit accordion-like multilayered structures stabilized by van der Waals forces or hydrogen bonding [25,26,34,35]. The in situ etching strategy, which produces hydrofluoric acid from fluoride salts and hydrochloric acid, has been demonstrated to be particularly advantageous for obtaining monolayer MXenes.
Ti3C2Tx MXene prepared via an exfoliation method was combined with Cu2O through a precipitation approach to construct Ti3C2Tx-nanosheets/Cu2O composites. A Schottky heterojunction was formed at their interface, which significantly enhanced charge separation, with the Ti3C2Tx nanosheets acting as efficient electron acceptors. Reactive species such as superoxide radicals (·O2) and holes (h+) selectively degraded tetracycline, achieving a removal efficiency of 97.6% within 50 min, compared to only 62% for pristine Cu2O (Figure 4) [14]. Post-etching treatment of Ti3C2Tx via exfoliation is an effective strategy to enlarge its specific surface area.
However, semiconductor catalysts grown on such structures often nucleate at edges rather than on the basal planes, which weakens the interfacial interactions between MXenes and semiconductors and diminishes the dual role of MXenes as a growth platform and charge reservoir [35,37]. It is worth noting that the surface end groups and morphological features of MXenes can regulate this growth behavior. Gentile et al. reports that aggressive acid etching favors fluorine-terminated (-F) MXenes with crumpled morphologies, whereas mild conditions preserve hydroxyl (-OH) terminations [38]. Moreover, hydrophilic surface groups (-OH, -F, -O) can facilitate hydrogen bonding with intercalants such as tetrabutylammonium hydroxide or hydrazine hydrate. Introducing macromolecules (e.g., cetyltrimethylammonium bromide, CTAB) not only enlarge interlayer spacing but also suppresses restacking via steric hindrance [39], offering novel pathways for structural modulation. Therefore, the synthesis process of MXenes can be precisely controlled to obtain a larger specific surface area, more active sites, and shorter charge migration paths.

2.2. Electrical Conductivity and Band Structure Regulation of MXenes

MXenes derive their excellent conductivity from transition-metal frameworks (e.g., Ti, Mo) covalently bonded with C/N and a high electron density near the Fermi level, imparting quasi-metallic properties. Ti3C2Tx, the most studied MXene, reaches 6000–8000 S/cm, exceeding graphene and most 2D materials. Mo2CTx shows lower values (2000–4000 S/cm) due to stronger Mo–C scattering, while nitride MXenes (e.g., Ti2NTx: 1000–3000 S/cm) still surpass conventional oxides such as TiO2 (~10−9 S/cm). Overall, MXene conductivities range from 100 to 20,000 S/cm, enabling efficient charge storage and transport. However, resistivity increases with thickness (e.g., doubling as Ti3C2Tx grows from 3 to 9 nm) [40,41], making single-layer nanosheets superior for catalytic applications.
Importantly, electronic transport and band structures (0.92–1.75 eV) can be tuned via surface terminations (-OH, -O, -F, -S, -Cl) and interlayer spacing. These functional groups, formed during HF etching of MAX phases, can be tailored through etchant conditions and post-treatments (e.g., intercalation, annealing), allowing programmable electronic structures [42,43,44]. For example, theoretical calculations show that the band gap of Sc2C can be tuned from 0.44 eV (–OH) to 1.07 eV (–F) and 1.85 eV (–O), depending on the electronegativity of the termination and the resulting charge transfer between terminations and the metal layers [45]. Such tunability is critical for optimizing charge transfer in photocatalysis, highlighting the importance of controlling MXene morphology and surface terminations.
MXenes, composed of transition metals and tunable surface terminations, synergize with primary catalysts to achieve efficient photocatalytic pollutant degradation. With a lower Fermi level than semiconductors [46], MXenes function as universal co-catalysts by suppressing electron–hole recombination while enhancing catalyst dispersion and adsorption capacity. For instance, N-doped Ti3C2Tx MXene synthesized via in situ polymerization of dopamine hydrochloride exhibits highly dispersed active sites and elevated conduction band potential (−0.78 eV vs. NHE), which facilitates charge transfer. Nitrogen doping induces charge delocalization and leverages Ti d-orbital contributions, resulting in a 2.3-fold increase in electron transfer rate compared to undoped Ti3C2Tx [47]. Similarly, in co-doped ZnTiO3/Ti3C2Tx nanohybrids, the -OH, -O, and -F terminations impart a strongly negative surface charge (−32 mV) [48]. This electrostatic repulsion reduces interference from background ions (HCO3, Cl, SO42−, NO3), thereby minimizing competitive adsorption and enhancing tetracycline degradation efficiency by 45% under visible-light irradiation. Surface terminations can also be modulated post-synthesis: hydrothermal treatment replaces –F groups with –O/–OH, while calcination converts -OH to -O groups [49]. Such directional conversion optimizes electronic properties, promoting reactive oxygen species (e.g., O2) generation and enhancing light absorption. For instance, F-CoFe2O4@MXene composites achieve 91% of crystal violet and 87% of bisphenol A degradation under 140 min solar irradiation, attributed to fluorine-enhanced charge separation and broad-spectrum photon utilization [50]. Critically, MXene’s abundant and naturally tunable terminal groups (–OH, –O, –F, etc.) eliminate the need for external dopants, providing a sustainable route to in situ construction of high-performance co-catalytic interfaces. This intrinsic multifunctionality establishes MXene as a versatile platform for designing high-performance, cost-effective photocatalytic systems.

2.3. Theoretical and Experimental Coupling Analysis of MXene Photocatalytic Mechanism

To deeply analyze the action mechanism of MXene-based photocatalytic materials, it is necessary to combine theoretical calculations with advanced experimental characterization techniques to reveal their electronic structures, interfacial charge transfer, and reaction pathways at the atomic scale. In this section, through multi-scale simulations and dynamic in situ analysis, a complete logical chain of “structure–performance–mechanism” is constructed, providing theoretical support for the directional regulation of the coordination structure of MXene.

2.3.1. Theoretical Calculations Reveal the Electronic Properties and Catalytic Active Sites of MXene

The electronic properties of MXenes are intrinsically governed by their composition (transition metals) and surface chemistry (end group). The d-electron count of the transition metal strongly modulates their density of states near the Fermi level, influencing both conductivity and catalytic behavior. For instance, Ti-based MXenes (e.g., Ti3C2) exhibit metallic behavior with high DOS at the Fermi level, while O-terminated Ti2C MXene (Ti2CO2) displays a finite band gap (~0.2–1.2 eV depending on model and strain), indicative of semiconducting characteristics. Surface functionalization with electronegative terminations (F, OH, O) further tailors MXene’s electronic structure: termination with O tends to open a gap, while mixed terminations (e.g., OH/O) introduce localized states reducing mobility [51]. For example, many-body GW/DFT calculations for monolayer Ti2CO2 report an indirect band gap of ~1.15 eV and that under small tensile strain (~4%) Ti2CO2 undergoes an indirect-to-direct band-gap transition [52]. Also, studies of Ti3C2 with different terminations (F, OH, O) show significant shifts in electronic transport and optical absorption associated with surface chemistry [53]. These tunability channels are unique to MXenes and enable precise design of band structure for photocatalytic optimization.

2.3.2. Theoretical Calculations Reveal That the Electronic Properties of MXene Are Related to the Photocatalytic Mechanism

High electrical conductivity, synonymous with elevated charge mobility, is pivotal for enhancing photocatalytic reaction rates. Both properties, as manifestations of MXene’s electronic characteristics, ultimately govern its catalytic potential. Han et al. employed four-point probe measurements to quantify MXene conductivities, showing that Ti-based MXenes possess high conductivity, which significantly exceeds that of some Nb- or Mo-based analogs. Surface terminations further modulate conductivity: Hart’s group demonstrated that F-termination desorption during in situ annealing of Ti3C2Tx films enhances conductivity by a notable margin [54]. Transition metal d-electron configurations influence the density of states near the Fermi level. Surface functionalization with electronegative terminations (e.g., F, –OH, O) also tailors the band structure: NaOH treatment followed by vacuum annealing removes F terminations, increasing O termination prevalence and improving conductivity in Ti3C2Tx thin films [55].
This study reveals the unique advantages of MXene in photocatalysis through multi-scale theoretical calculations. ① A novel “d-orbital engineering–surface termination” strategy bypasses traditional doping-dependent bandgap tuning. ② The built-in electric field effect induced by the gradient functionalization of MXene was discovered, enabling the efficient separation of photogenerated carriers (with the quantum efficiency increased by 2.3 times). ③ The Ti d-band center (−2.1 eV) and O-termination p-orbital hybridization synergistically lower molecular adsorption energy. Therefore, MXene achieves efficient integration of light absorption, charge separation, and surface reactivity without external modification through its inherently tunable electronic structure and surface chemistry, surpassing the limitations of materials such as graphene.

2.4. How MXenes Improve Photocatalytic Activity

MXenes (e.g., Ti3C2Tx) enhance photocatalysis primarily by accelerating charge separation/transport, engineering built-in junctions at semiconductor interfaces, modulating surface chemistry (terminations/doping), and providing high-affinity adsorption sites that pre-concentrate pollutants [56]. These attributes translate to faster kinetics and higher antibiotic-degradation efficiencies under visible/simulated solar light.
MXenes possess quasi-metallic conductivity and a Fermi level favorable for accepting photogenerated electrons from semiconductors, suppressing e/h+ recombination and boosting reaction rates [57]. In Ti3C2Tx/Cu2O Schottky heterojunctions, Ti3C2Tx acts as an electron acceptor, delivering 97.6% tetracycline removal within 50 min under visible light—far above pristine Cu2O—demonstrating efficient interfacial charge extraction [14]. Integrating MXene with oxides (e.g., TiO2) or other semiconductors forms built-in fields that promote directional carrier migration and preserve strong redox potentials [58]. MXene–TiO2 hybrids with exposed (001) facets show markedly improved tetracycline degradation under sunlight/NIR due to optimized band alignment and interface contact [59]; more broadly, such heterojunctions are known to outperform single components in antibiotic removal.
MXene surface groups (–O/–OH/–F) and heteroatom doping tune band structure, work function, and adsorption energetics. N-doped Ti3C2Tx (via dopamine-assisted routes or post-treatments) exhibits enhanced electron delocalization and faster interfacial transfer, leading to stronger visible-light activity for antibiotic degradation compared with undoped MXene [47].
The 2D, high-surface-area MXene scaffold offers abundant adsorption sites (π–π/chemisorption, electrostatic interactions), enriching antibiotic molecules near reactive centers and lowering apparent activation barriers. MXene-based hybrids (e.g., CuFe2O4/MXene; ZnTiO3/Ti3C2Tx) show higher tetracycline/sulfonamide removal under visible light than their MXene-free counterparts [56], consistent with an adsorption-assisted photocatalytic pathway.
Implication for antibiotic wastewater: Collectively, these mechanisms—rapid electron extraction, favorable band alignment, tunable surface chemistry, and adsorption enrichment—explain the consistent gains observed for MXene-based systems in degrading tetracycline and related antibiotics under visible/solar irradiation. The functions of photocatalytic activity optimization are listed in Table 2 [41].
MXene (especially Ti3C2Tx) faces challenges in terms of stability in aqueous media and oxygen-rich environments, mainly because the exposed titanium atoms on its surface are easily attacked by dissolved oxygen to form Ti-O bonds, which in turn generate TiO2 nanoparticles. These TiO2 nanoparticles can serve as nucleation sites and partially dissolve under the corrosive effect of KOH. This dissolution-recrystallization process ultimately leads to the formation of nanowires, which instead effectively enhances the overall stability of the material [62].
Currently, with the rapid development of 2D MXene, the existing theoretical models have many limitations in accurately predicting the electronic band structures and plasmonic resonances of MXene materials. DFT (Density Functional Theory) calculations require a periodic and ordered crystal model. However, the atomic arrangements of MXene, especially the atomic arrangements of the surface functional groups Tx, are diverse and randomly distributed. The decomposition and oxidation of MXene under catalytically relevant conditions should be considered, and the active phases for specific reactions should be specified. In addition, DFT also has self-interaction error (SIE), especially for strongly correlated electron systems (such as metals containing d or f orbitals). The SIE leads to inaccurate descriptions of electron-electron interactions, and accurately describing the band structure remains a challenge. Temperature effects, van der Waals interactions, charge transfer kinetics, and multiscale modeling are also ignored in DFT calculations. To overcome these limitations, more advanced theoretical methods need to be adopted. Such as hybrid functionals, GW approximation, and the Bethe-Salpeter equation for exciton effects. Moreover, machine learning techniques can be applied to the discovery of MXene materials, and deep learning techniques can be used to eliminate the discrepancies between theoretical predictions and experimental results [63]. Such data-driven discoveries can lead to more accurate and efficient material development and applications.

3. MXene-Assisted Catalytic Optimization Mechanism

3.1. Interfacial Charge Transfer Mechanism and Theoretical Calculation Between MXene and Semiconductors

The first research report on MXene predicted that Ti3C2 MXene has semi-metallic properties through density functional theory (DFT) calculations. The analysis of the density of states (DOS) near its Fermi level shows that the Ti 3d orbitals dominate the electron transport behavior, which is consistent with the metallic conductivity of the MAX-phase parent materials (such as Tin+1AlCn), as illustrated in Figure 5 [64,65,66]. After the Al layer in the MAX phase is removed by chemical etching, the electronic structure of Tin+1Cn MXene changes significantly. The broken Ti-Al bonds lead to the redistribution of electrons in the Ti 3d orbitals, forming delocalized Ti-Ti metallic bonding states (local DOS peaks appear near the Fermi level) [66]. In addition, to study the enhancement effect of Ti3C2 on the photoactivity of TiO2, DFT was used to analyze the energy band structures of Ti3C2 and anatase-phase TiO2. The schematic diagram of their atomic structures is shown (Figure 5a) [67]. Figure 5b shows the total density of states (Total) and the projected density of states (PDOS) of Ti3C2 MXene. The horizontal axis represents energy (eV), and the vertical axis represents the value of the density of states. The total and projected DOS (Figure 5b) demonstrate that the Fermi level (EF, 0 eV) lies within the conduction band (energy range: −2 to +4 eV), confirming metallic conductivity with abundant free electrons. Ti 3d orbitals dominate the conduction band (>0 eV), contributing delocalized electrons, while C 2p orbitals hybridize with Ti in the valence band (<0 eV) to stabilize the structure. This metallic nature enables Ti3C2 to act as an electron highway, rapidly extracting photogenerated electrons when forming a Schottky junction with TiO2. Electron injection into MXene leaves holes in TiO2’s valence band, achieving 5–10 times enhanced charge separation efficiency. Electrostatic potential mapping (Figure 5c) reveals a potential barrier within Ti3C2, with lower potentials at the edges and elevated potentials centrally. Calculated work function (4.46 eV vs. vacuum) and Fermi level (+0.04 eV vs. NHE) indicate moderate electron emission capability [67]. According to Equation (1),
E F = E v a c φ
where Evac is the energy of a stationary electron in a vacuum near the surface and EF determines the calculation of the structure of the ground state electron. DFT simulations further show that anatase TiO2 exhibits a 3.09 eV bandgap, with its conduction band minimum (ECB) at +0.25 eV vs. NHE. The significant band offset (ΔE = 0.21 eV) between TiO2’s ECB and Ti3C2’s EF drives spontaneous electron transfer from TiO2 to MXene, forming a Schottky junction that suppresses electron-hole recombination. As shown in Figure 5d, a slab model was used to simulate Ti3C2 QD, and a side view of the heterojunction shows the tightly bound Ti3C2 QD/PGCN [68]. As illustrated in Figure 5e, Ti3C2 quantum dots (QDs) exhibit excellent electrical conductivity. For the Ti3C2 QD/PGCN heterojunction, the valence band maximum (VBM) is predominantly contributed by N atoms, while the conduction band minimum (CBM) mainly originates from Ti atoms (Figure 5g). This electronic configuration implies that photogenerated electrons are expected to transfer from PGCN to Ti3C2 QDs. Furthermore, the partial overlap between the conduction and valence bands of the heterojunction suggests enhanced electrical conductivity compared to pristine PGCN, which synergistically promotes visible-to-near-infrared light absorption [68]. These findings demonstrate that under near-infrared irradiation, the rapid extraction of photogenerated electrons from PGCN by Ti3C2 QDs effectively prolongs charge carrier lifetime, thereby significantly improving the photocatalytic performance of the heterojunction system. The calculated electrostatic potentials for PGCN, Ti3C2 QD are shown in Figure 5h,i. The work functions of PGCN and Ti3C2 QD are 4.67 and 6.18 eV, respectively.
The results show that the high electrical conductivity of Ti3C2 accelerates electron migration and reduces the interfacial resistance; the Schottky barrier drives the unidirectional injection of electrons, and holes accumulate on the surface of TiO2 to participate in the oxidation reaction [59]. Therefore, the high electrical conductivity and good energy structure of MXenes make them a good reservoir for capturing and shuttling the photoelectrons generated by semiconductors, thus promoting the separation of carriers and enhancing the photoactivity of MXenes-based composites.
Theoretical calculations further reveal the microscopic basis of MXene’s tunable electronic structure and support its universality as an electron acceptor. For instance, the energy band structure of a single pristine Ti3C2 layer is like that of a typical semi-metal with a finite density of states at the Fermi level. When its surface is terminated with F groups, a distinct separation between the conduction band and the valence band can be observed, and thus the energy band structure exhibits semiconductor characteristics. This implies that the energy band structure of MXenes evidently varies with the change in surface functional groups, which will affect the role of MXenes in accepting charge carriers in semiconductors.

3.2. MXene Acts as an “Electronic Sponge” in Z-Type Heterojunction

MXene’s foremost advantage in hybrid photocatalyst design lies in its ability to simplify heterojunction architectures. The Fermi level, a critical determinant of a material’s magnetic and photoelectrochemical properties, positions MXene as a universal cocatalyst for semiconductors due to its lower Fermi level relative to most semiconductors [46]. This energy alignment suppresses electron-hole recombination while enhancing photocatalyst dispersion and adsorption capacity. For instance, Ti3C2Tx MXene has been integrated with TiO2 [69], ZnIn2S4 [70], and Ag3PO4 [71] to form heterointerfaces, where heterojunctions (e.g., Schottky, Z-scheme) are strategically constructed to minimize charge carrier recombination and maximize photoactivity [72]. MXene’s inherent high conductivity enables rapid electron transfer from the semiconductor conduction band to MXene upon photoexcitation. The transferred electrons are confined within MXene due to the built-in electric field at the heterojunction interface, preventing back-recombination with holes and enhancing photocatalytic efficiency by 3–5× compared to standalone semiconductors [23,73]. As illustrated in Figure 6, MXene predominantly forms Schottky junctions, van der Waals heterostructures, or Z-scheme configurations with common photocatalysts [72,74]. The weak in-plane interactions of MXene’s 2D structure facilitate seamless stacking with other 2D materials (e.g., graphene, MoS2), forming van der Waals heterojunctions with optimized interfacial charge transfer pathways. Therefore, MXene offers core advantages for designing high-performance, structurally simplified hybrid photocatalytic systems through its intrinsic Fermi level position, high conductivity, and two-dimensional structural characteristics.
MXene, as a pioneering two-dimensional material, has high conductivity like that of precious metals, which is one of its most notable characteristics. This high conductivity stems from its delocalized electron system, enabling it to exhibit excellent performance in electron storage and transport [26,34,41,72]. The unique electronic structure of MXene, characterized by a Fermi level (Ef) positioned within the conduction band (CB), enables spontaneous electron transfer when interfaced with semiconductors. When MXene’s Ef lies below that of a coupled semiconductor, photoexcited electrons migrate from the semiconductor to MXene, establishing an intrinsic electric field at the interface. This field simultaneously enhances charge separation efficiency and enables directional electron transport across extended spatial domains [75]. The interfacial energy alignment creates a space-charge layer near the semiconductor surface, inducing upward band bending that forms a Schottky junction. This energy barrier effectively suppresses charge recombination by preventing electron backflow into the semiconductor [76]. This highly efficient charge management mechanism has been strongly validated in experiments. Ag3PO4/Ti3C2 Schottky photocatalysts prepared using electrostatic self-assembly technology exhibit excellent degradation performance against antibiotics, persistent organic pollutants, and dyes under visible light. Notably, these MXene-based hybrids address the inherent photocorrosion limitations of pristine Ag3PO4 while outperforming both pure Ag3PO4 and Ag3PO4/RGO composites in photocatalytic activity and operational stability [71]. Band engineering through heterojunction design has proven crucial for optimizing charge transfer pathways in photocatalytic systems. Among various configurations, Z-scheme heterojunctions stand out as particularly efficient architectures for directional charge migration while preserving strong redox potentials.
The formation of Z-scheme heterojunctions necessitates stringent criteria, including interleaved band alignment [77], substantial work function disparity [78], and robust interfacial coupling forces [79]. Notably, MXene inherently fulfills these prerequisites for constructing Z-scheme heterojunctions with diverse semiconductors, owing to its tunable electronic structure and surface reactivity (Figure 6c) [80]. Beyond conventional heterojunction engineering in single-component photocatalysts, MXene quantum dots (MQDs) have demonstrated exceptional versatility in modulating interfacial architectures for Z-scheme configurations within multicomposite photocatalytic systems. Mimicking the electron transport chain in natural photosynthesis, Z-scheme heterojunctions enable asymmetric coupling of two semiconductors’ band structures to establish a staggered charge transfer pathway. This architecture drives photogenerated electrons from the conduction band (ECB) of Semiconductor A to the valence band (EVB) of Semiconductor B, achieving spatial separation of electron-hole pairs. The resultant charge redistribution concentrates holes in semiconductors with high oxidation potentials (e.g., TiO2) while accumulating electrons in materials with strong reduction potentials (e.g., g-C3N4). This synergistic configuration simultaneously preserves the redox capabilities of both components and amplifies photocatalytic driving forces. Through calcination processes, a novel dual Z-scheme heterojunction composed of graphitic carbon nitride/Ti3C2 MXene/black phosphorus (CN/MX/BP, CXB) was synthesized. This system achieved exceptional photocatalytic degradation efficiency (>99%) for ciprofloxacin (CIP) within 60 min under visible light irradiation (λ > 420 nm) [81]. Similarly, a meticulously engineered In2S3/Ti3C2 MXene quantum dot/SmFeO3 Z-scheme heterojunction demonstrated 98% sulfamethoxazole degradation efficiency within 120 min under visible light exposure [82]. The enhancement of photocatalytic performance is due to the role of MXene as a metallic conductive electron mediator, which enables the directional transfer of charges. Meanwhile, the surface polarization effect effectively suppresses the recombination of electrons and holes, allowing the optimized interfacial structure to promote the occurrence of multistep redox reactions.
In Z-type heterojunctions, MXene’s unique electronic structure enables a dual key role as an electron acceptor and a hole transport mediator, thereby strengthening bidirectional charge management (Figure 6b). Specifically, the Fermi-level/work-function of Ti3C2Tx is positioned to accept photogenerated electrons from many n-type semiconductors, facilitating electron trapping at the MXene interface and suppressing recombination [83,84]. MXene’s high electrical conductivity promotes rapid interfacial electron extraction from Semiconductor A (e.g., TiO2), while its surface terminations (–OH/–O/–F) provide favorable interfacial interactions—including hydrogen bonding—with the valence band of Semiconductor B (e.g., g-C3N4) (Table 3), enabling directional hole transport within Z-scheme assemblies [85]. In additional, Ti3C2Tx exhibits broadband plasmonic/photothermal response extending from the visible to the near-infrared (≈400–2500 nm), which enhances local electromagnetic fields at MXene–semiconductor interfaces and accelerates carrier migration to reactive sites [86]. Collectively, these effects underpin the “electron-sink” function of MXene in Z-scheme systems and account for the consistently improved photocatalytic performance reported for MXene-coupled heterojunctions in pollutant degradation.
The core strategy for achieving the above performance optimization lies in the construction of photocatalyst composite materials. This strategy aims to improve overall performance through structural optimization, enhanced material/energy transfer, and precise band structure modulation. The core principle involves establishing point-to-surface or surface-to-surface interfacial contacts between materials, where the interfacial characteristics crucially determine the photocatalytic performance beyond individual component properties. These interfaces facilitate synergistic interactions through ① tailored electron transport mechanisms, ② reduced electron-hole recombination rates, and ③ enhanced chemical stability via functional group mediation. As demonstrated in representative systems, Ti mesh-supported flexible g-C3N4/Ti3C2/TiO2 nanotube arrays exhibited 85.12% degradation of 10 mg/L tetracycline hydrochloride within 180 min, outperforming conventional g-C3N4/TiO2 composites (36% degradation for 20 mg/L under identical conditions) [89]. This enhancement arises from two synergistic mechanisms: ① Formation of Z-scheme heterojunctions between g-C3N4 and Ti3C2 and ② Strong interfacial coupling between Ti substrate and composite materials.
The MXene surface is rich in functional groups such as -OH, -O, and -F, providing an ideal basis for chemical modification and light-active semiconductor loading as a two-dimensional platform. This enables semiconductor photocatalysts to self-assemble in situ on MXene to form composite materials while also facilitating the loading of metal oxides (such as Fe3O4 [90], Ag3PO4 [91]) to construct high-performance composite photocatalytic systems. Room-temperature synthesized TCT (TiO2 NPs/C-doped amorphous TiOx homojunction with residual Ti3C2Tx MXene cocatalyst) achieved 91.5% tetracycline degradation within 100 min, maintaining 86.9% efficiency after six cycles. The Type-II heterojunction configuration promotes directional charge separation [92]. Magnetically responsive MXene hybrids (Fe2O3 nanoparticles anchored on Ti3C2 surface) designed by combining heat treatment and hydrothermal method. This composite demonstrates enhanced tetracycline removal efficiency (92.4% within 60 min) and rapid magnetic separation capability (saturation magnetization: 15.2 emu·g−1), while the Fe2O3 incorporation effectively mitigates the inherent stability limitations of pristine MXene under oxidative conditions [90]. This method of strategically combining MXene with structurally compatible materials (such as semiconductors and metal oxides) successfully addresses the key shortcomings of single component photocatalysts, namely rapid charge recombination and insufficient active sites. Furthermore, MXene’s transition metal matrix enables in situ growth of semiconductor hybrids (e.g., TiO2-C [89]) through its role as both a structural template and electron mediator. A summary of the photocatalytic-promoting functions of MXene is presented in Table 4.

3.3. The Chemical Structure of MXene Strengthens the Stability of the Material

3.3.1. Regulation Strategies for Strong Metal–Support Interactions in MXene Photocatalytic Systems

The emerging strong metal–support interaction (SMSI) offers a transformative paradigm for designing MXene-based photocatalytic systems. Owing to their lamellar architecture, tunable surface terminations, and metallic-like conductivity, MXenes provide robust platforms for SMSI-type electronic coupling and interfacial configuration control, enabling advances in carrier separation, active-site stabilization, and photo-corrosion resistance [94].
Surface terminations (–O/–F/–OH) participate in p-d orbital hybridization with supported metals and semiconductors, acting as interfacial electron buffers and modifying band alignment. In Au/Ti3C2Tx systems, experiments and calculations show substantial charge redistribution at the interface and improved interfacial transport, consistent with an SMSI-like electronic contact that favors directional charge flow under illumination [95]. Beyond static interfacial states, photo- or thermo-induced SMSI is known to restructure metal/support contacts and form ultrathin encapsulation layers on classical oxide supports; analogous interfacial reorganization pathways (e.g., Ti–O–M linkages, defect-assisted bonding) are increasingly explored on 2D supports and MXenes to stabilize active phases during photocatalysis [96]. For MXene/transition-metal dichalcogenide (TMD) junctions (e.g., V2C//MoS2), interface engineering and covalent/defect-anchored bonding have been shown to strengthen electronic coupling, tune the local work function/d-band center, and optimize adsorption energetics of intermediates—collectively promoting bidirectional charge transfer and durability [97]. In parallel, MXenes exhibit broadband optical/photothermal responses (visible-to-NIR) that intensify local electromagnetic fields and enhance multi-field energy coupling at the metal/MXene or semiconductor/MXene interface, which is beneficial for light harvesting and hot-carrier utilization.
Altogether, these SMSI-mediated strategies—combining termination-guided electronic/geometric effects, adaptive interfacial protection, and broadband light–matter coupling—provide practical design rules for stable, high-efficiency MXene-based photocatalytic systems applicable to solar water splitting, CO2 reduction, and pollutant degradation.

3.3.2. Case Studies of Enhanced Stability in MXene Photocatalytic Systems

MXene, with its unique structure rich in transition metals such as titanium, shows great potential as a multifunctional composite catalyst. The surface-exposed Ti sites on MXene exhibit significantly enhanced redox reactivity compared to conventional carbon materials, facilitating in situ redox cycling of metal oxide catalysts and improving photocatalytic stability. Theoretical studies confirm that these Ti sites accelerate electron transfer kinetics, with measured rate constants exceeding those of noble metal cocatalysts by 2–3 orders of magnitude. Simultaneously, the abundant surface Ti and transition metal sites stabilize metal oxides through dual mechanisms: (1) enhancing synergistic catalytic pathways via optimized charge transfer, and (2) suppressing metal leaching by passivating lattice dissolution channels.
MXene’s excellent interface control capability is the key to its performance improvement. The hydrophilic surface functional groups (-OH, -O) of Ti3C2 enable intimate interfacial contact with semiconductors like Ag3PO4. This interaction, combined with MXene’s redox-active Ti sites, mitigates Ag+ reduction by photoelectrons, reducing photocorrosion by 78% compared to bare Ag3PO4 [41]. In Cu2O/Cu@MXene nanocomposite, MXene substrate simultaneously enhances tetracycline (TC) adsorption and interfacial electron transfer, achieve 99.14% tetracycline removal within 30 min (pseudo-first-order rate constant: 0.1505 min−1), representing a 3.2 times enhancement over standalone Cu2O/Cu, and the degradation rate remains above 82% after five cycles [98].
Photocorrosion resistance remains a critical challenge for chalcogenide semiconductors (e.g., CdS, ZnS [99]), where surface sulfide ions (S2−) oxidize to elemental sulfur or sulfates during photocatalysis, leading to structural degradation. The 1D CdS/2D MXene heterojunction demonstrates exceptional stability, maintaining consistent hydrogen evolution rates over 15 h of continuous operation [100], in stark contrast to pristine CdS, which exhibits an activity loss within 8 h [101]. This highlights MXene’s dual role as both a charge-transfer mediator and protective cocatalyst to suppress photocorrosion. Further stability enhancements are achieved through polymer-MXene nanocomposite engineering. Polyvinylpyrrolidone (PVP)-encapsulated MXene cast onto polyethylene terephthalate (PET) substrates forms PMP photoresponsive films for tetracycline degradation. MXene incorporation modulates interlayer spacing (increasing from 0.98 nm to 1.24 nm) and composite surface roughness (Ra = 12.3 nm vs. 8.7 nm for pure PET), optimizing active site distribution and reducing photocatalytic deactivation risks [15]. More importantly, MXene’s large specific surface area makes it an ideal catalyst growth substrate, effectively preventing active component aggregation and inhibiting deactivation, and its functionality is superior to that of precious metals that require high specific surface area carriers [102].
MXene quantum dots (MQDs) offer unique advantages for constructing efficient photocatalytic systems due to their ultra-high specific surface area (approximately 300 m2/g), strong quantum confinement effect, and excellent metallic conductivity. MQD can act as intermediate electron pathways to form Z-scheme heterojunctions, thereby enhancing photocatalytic reactions. MQDs with lateral dimensions below 20 nm circumvent the shielding effects commonly observed in noble metal nanoparticles [26], while the Schottky junctions formed between MQDs and photocatalysts enhance carrier mobility and interfacial coupling strength [71,100]. Ti3C2Tx MQDs exhibit exceptional antibacterial properties under low-concentration, dark conditions, maintaining efficacy after 4 h of continuous mechanical agitation [103]. Integrating MQDs into photocatalytic systems enables innovative exploration of their bacteriostatic mechanisms. Leveraging strong quantum confinement, metallic conductivity, anisotropic charge transport, and high surface area, MQDs serve as rapid electron highways in Z-scheme heterojunctions, effectively suppressing bacterial adhesion and biofilm formation. This reduces corrosion and fouling from microbial metabolites, extending photocatalyst lifespan. MQDs excel in generating reactive oxygen species (ROS) under UV irradiation, which are crucial for the efficient degradation of antibiotics. For instance, a g-C3N4/BiOBr/MQD composite synthesized via solvothermal methods combines heterojunctions, oxygen vacancies (OVs), and MQD-enhanced charge transfer. OVs promote O2 chemisorption, while MQDs accelerate electron transport and selectively generate 1O2, achieving 99% tetracycline hydrochloride degradation within 30 min under visible light. Long-term stability tests reveal >90% efficiency retention after 8 cycles (40 min each) [104]. By exploiting MXene’s structural, chemical, and reactive versatility, this approach resolves persistent challenges like photocorrosion in conventional semiconductors, enhancing both stability and operational durability.

3.4. Photocatalytic Composites for Co-Catalytic Modification with MXene Co-Catalytic

To further explore MXene as a co-catalyst to enhance photocatalytic performance and accelerate the rate of photocatalytic reactions, the photothermal effect, localized surface plasmon resonance (LSPR) effect generated under the regulation of MXene co-catalysis, as well as the synergistic effects with other modification methods were investigated, to clarify the versatility and practicality of MXene as a co-catalyst.

Innovative Mechanisms for Plasma-Induced Chemical Reactions

MXene exhibits excellent interface compatibility in photocatalytic systems, with advantages in terms of broad spectral response and low defect interfaces (Table 5). Its localized surface plasmon resonance (LSPR) effect enables broadband infrared absorption extending beyond 800 nm, effectively addressing the narrow optical response range characteristic of conventional semiconductors [49]. The atomically smooth surface morphology (root-mean-square roughness < 0.2 nm) and inherent two-dimensional architecture promote robust van der Waals interactions (binding energy ≈ 0.03 eV/atom) with semiconductor substrates, resulting in a 62% reduction in interfacial defect density compared to conventional heterojunctions.
Metal nanostructures generate a plasmonic effect upon photoexcitation and trigger chemical reactions through two core mechanisms. When the energy of the incident light (hv) exceeds the Fermi level (Ef) of the metal, non-equilibrium state hot electrons are generated. Subsequently, the energy is transferred to the lattice via electron–phonon coupling within a picosecond time scale. At the same time, the vibrational energy of the lattice is transferred to the surface molecules through interfacial vibrational coupling. These two mechanisms (hot electron transfer and phonon-mediated heat transfer) work synergistically to enhance the efficiency of MXene-based photocatalytic systems.
MXene, as a new class of two-dimensional transition metal carbide, has entered the advanced plasmonic material system due to its unique electronic structure characteristics. Compared with traditional materials, it exhibits three breakthrough characteristics in the field of light-matter interaction.
First, it shows a broadband plasmonic response. A significant surface plasmon resonance effect is presented in the visible to near-infrared wavelength range (400–2000 nm), and its resonance frequency can be tuned by the number of layers (a single-layer red-shift is about 15%). Compared with noble metals (Au/Ag only cover the visible light region), the plasmonic absorption bandwidth of MXene can be extended by 3–5 times, providing the possibility for multi-wavelength light energy capture. Secondly, the surface termination group engineering can induce significant surface-shell synergistic effects. Under surface engineering modification, terminating groups such as -O and -F regulate the position of the Fermi level through p-d orbital hybridization (the regulation range can reach 0.8 eV). More importantly, the atomic-level flatness of MXene’s surface (roughness < 0.2 nm) and the weak interlayer van der Waals interactions (binding energy ~0.03 eV/atom) work together to give it an exceptional ability to precisely control the direction of molecular adsorption. In addition, it has other advantages compared with traditional photocatalysts in Table 6.
Ti3C2Tx MXene exhibits exceptional photothermal conversion efficiency, enabling direct solar-to-thermal energy transduction within catalytic systems. The localized surface plasmon resonance (LSPR) effect amplifies interfacial electromagnetic fields (field intensity enhancement: 102–103), thereby reducing activation energy barriers (ΔE ≈ 0.35 eV) and enhancing both thermodynamic feasibility and kinetic rates of photocatalytic degradation [30,111]. Thickness-dependent studies reveal that Ti3C2Tx nanosheets with sub-10 nm dimensions achieve 1.8 times stronger LSPR fields than bulk counterparts, positioning MXene as a cost-effective alternative to noble metals for plasmonic applications [112]. Under illumination, plasmonic Ti3C2Tx generates high-energy hot holes that directly oxidize antibiotics and organic pollutants into low-molecular-weight intermediates. Concurrently, delayed electron-hole recombination enables hot electrons to migrate toward adsorbed O2, generating reactive superoxide radicals (O2−) for pollutant mineralization [112,113]. In the photothermal effect, it has been found that the construction of heterojunctions increases the electron cloud density of Ti3C2Tx. Meanwhile, it also enhances the number of hot electrons and the intensity of the local electromagnetic field generated by the localized surface plasmon resonance (LSPR) [114]. During the photocatalytic reaction process, the separation of photogenerated electron-hole pairs and the migration of photogenerated electrons occur under photoexcitation. At the same time, the thermal effect also provides a certain amount of energy for the photoelectrons, which serves as one of the driving forces for their separation and migration.
Surface defects commonly found in solid materials, particularly surface oxygen vacancies (OVs), play a central role in photocatalysis. OVs, as the imperfect lattices with rearranged electron distributions left by the loss of oxygen atoms in the semiconductor crystal structure, are the most common anion vacancies [115]. They are conducive to the formation of active sites that are crucial for the catalytic process. The escape of oxygen atoms from the lattice sites of semiconductors leads to the redistribution of electrons. By promoting the exposure of active sites and narrowing the bandgap, it improves the physical and chemical properties of the material, thus optimizing the kinetics of various reactions, and has attracted extensive attention especially in the field of photocatalysis [116]. Integrating the main catalyst with MXene nanomaterials forms a multicomponent heterojunction. The lattice mismatch and lattice distortion caused by the different crystal structures of each component induce the generation of interfacial oxygen vacancies (OVs), as shown in Table 7. The formation of oxygen vacancies captures many photogenerated electrons produced under photoexcitation and inhibits the recombination of photogenerated electron-hole pairs. The electrons occupying the O 2p orbitals can significantly activate the oxygen-deficient surface as an electron-rich center, thereby providing more oxygen adsorption sites, reducing the bandgap of the semiconductor, promoting the separation of electron-hole pairs, and exerting an enrichment effect on electrons and pollutants, thus enhancing the reaction ability under visible light.
Leveraging MXene’s intrinsic properties (e.g., tunable interlayer spacing, metallic conductivity) and co-catalytic optimization mechanisms, MXene composites with metal oxides [123], metal sulfides, graphitic carbon nitride (g-C3N4) [124], and oxyhalides [125] have emerged as robust platforms for antibiotic degradation. These composite systems effectively address two key challenges: (1) enhancing operational stability and (2) amplifying photoactivity via extended visible-light absorption and charge separation efficiency.
Developed by the Hefei Institutes of Physical Science, this architecture exploits MXene’s biocompatibility and interlayer spacing (0.98–1.25 nm adjustable) to in situ grow Cu2O/Cu nanoparticles. The system achieves 92% tetracycline removal within 60 min by synergizing adsorption-enriched tetracycline at MXene interfaces with accelerated electron transfer to Cu sites (charge transfer resistance reduced by 68%) [126]. Integrating MXene quantum dots (MQDs) with phosphorus-doped graphitic carbon nitride (g-C3N4) creates a 0D-2D heterostructure that enhances photocatalytic efficiency through dual mechanisms: (1) increased specific surface area and (2) suppressed charge recombination. This architecture demonstrates exceptional tetracycline hydrochloride (TC) degradation (88.40% in 60 min) and Cr(VI) reduction efficiency (94.2% in 45 min) under visible light.
The aerosol-assisted self-assembled g-C3N4/MXene/Ag3PO4 (PCN/M/AP) S-scheme heterojunction exhibits MXene-dependent activity enhancement. Increasing MXene content from 3 wt% to 7 wt% elevates TC degradation rates from 53.79% to 88.40% [91]. Under visible light irradiation, the rate constant value of PCN/M/AP is approximately 4.3 times higher than that of the pristine g-C3N4. This is attributed to the formation of a MXene-mediated Z-scheme heterojunction, which promotes the separation and transfer of photogenerated electron-hole pairs. In addition, after 5 consecutive cycles, the photocatalytic performance of PCN/M/AP for tetracycline (TC) still remains at 83.80% [127]. In MXene-based composites, g-C3N4 is widely used mainly due to its unique advantages. On one hand, it has advantages such as suitable bandgap energy, high chemical stability, a unique layered structure, and non-toxicity [80]. Compared with other metal-containing semiconductors, such as metal sulfides and metal oxides, metal-free g-C3N4 will generate a strong C-N conjugate structure after high-temperature calcination, ensuring excellent stability and resulting in the high photocatalytic ability of g-C3N4. On the other hand, the conjugate structure derived from triazine units and the two-dimensional structure caused by the weak van der Waals force between adjacent layers both create abundant active sites for contact with target pollutants. However, the visible light reaction rate of g-C3N4 is relatively low because its bandgap (Eg) is only 2.7 eV, and it only absorbs light with a wavelength less than 470 nm. At the same time, the recombination of electrons and holes easily occurs on g-C3N4, leading to a significant decrease in photocatalytic efficiency [128,129,130]. MXene is the key component that overcomes these limitations. As an efficient intermediate, MXene can not only achieve the rapid transfer of photogenerated electrons to enhance the overall photocatalytic activity but also construct more mature and effective Z-scheme heterojunctions to strongly suppress carrier recombination.

4. Dynamic Regulation of MXene Coordination Structure and Photocatalytic Environmental Adaptability

4.1. Photocatalytic Anti-Interference Mechanism in Complex Water Quality Environment

MXene exhibits multifunctional roles in photocatalytic antibiotic degradation, primarily mediated by its tunable surface/electronic properties governed by microstructure and surface chemistry. To comprehensively evaluate its synergistic catalytic performance, it is necessary to systematically analyse the effects of external environmental factors (pH value, coexisting ions and dissolved organic matter) on the system’s performance and stability.
The initial wastewater pH critically modulates photocatalytic activity through two mechanisms: (1) altering adsorbent surface charge via the point of zero charge and (2) inducing pH-selective antibiotic interactions. Below pH 6.85 (acidic to near-neutral conditions), electrostatic attraction between the positively charged catalyst surface and tetracycline hydrochloride (TCH, pKa ≈ 3.3) enhances adsorption capacity (Qe: 48.7 → 92.4 mg·g−1) [131]. Conversely, at pH > 6.85, reduced surface charge density weakens catalyst-antibiotic affinity, decreasing degradation efficiency (89% → 63% over pH 4–9). The acid dissociation constant (pKa) quantifies the pH-dependent ionization behavior of weakly acidic/basic antibiotics, defining the pH at which 50% dissociation occurs. Antibiotic-specific pKa values (Table 8) govern their ionization states: below pKa, neutral molecular forms dominate (e.g., sulfamethoxazole, pKa = 1.7/5.6), while ionized species prevail at pH > pKa (e.g., tetracycline, pKa = 3.3/7.7/9.7) [132]. This ionization dichotomy critically modulates adsorption selectivity on photocatalytic polymer surfaces. Cationic antibiotics (e.g., ciprofloxacin, pKa = 6.1/8.7) exhibit 3.2 times higher adsorption capacity (Qmax = 128 mg·g−1) on negatively charged MXene-PLA composites at pH 7.4 via electrostatic attraction. Anionic antibiotics (e.g., sulfadiazine, pKa = 2.0/6.5) preferentially bind to protonated amine-functionalized polymers (Qmax = 98 mg·g−1) at pH 4.2. The differential adsorption arises from pH-tunable surface charges (zeta potential range: −35 mV to +28 mV) and chemical affinity mismatches (hydrophobic/hydrophilic balance ΔH ≈ 12–18 kJ·mol−1). Molecular dynamics simulations confirm that ionized antibiotics exhibit 1.8–2.5× stronger binding energies (Ebind = −45 to −68 kJ·mol−1) compared to neutral forms due to enhanced electrostatic/hydrogen-bond interactions [133]. The degradation efficiency of antibiotics is governed by both molecular ionization states and catalyst structural stability. See Table 8 for details. Amphoteric antibiotics like tetracycline and amphotericin B exhibit pH-dependent charge switching, modulating their adsorption affinity to photocatalysts [134].
The presence of inorganic anions (e.g., HCO3, Cl) in real aqueous systems introduces competitive interfacial dynamics that govern antibiotic removal efficiency. HCO3 enhances tetracycline-class antibiotic degradation (e.g., oxytetracycline and chlortetracycline) through •OH radical amplification, whereas Cl exhibits dual inhibitory effects: (1) •OH scavenging and (2) catalyst surface passivation, collectively suppressing oxytetracycline and chlortetracycline degradation by 14.8% at 100 mmol L−1 with concentration-dependent inhibition trends. These anions further modulate antibiotic mobility by altering photocatalyst surface charge, promoting nanoparticle aggregation and reducing environmental transport efficiency. Such ion-specific interfacial engineering highlights the need for tailored photocatalyst design to mitigate competitive anion effects in complex wastewater matrices.
Humic acid is the main component of total organic carbon (TOC) in water (about 50%) and mainly acts as a free radical scavenger in photocatalytic systems. At 100 mg L−1 humic acid, TC-HCl removal decreases marginally from 99% to 94% due to competitive •OH consumption (quenching efficiency: 73.5%). Notably, humic acid preferentially adsorbs on TiO2 surfaces (Qmax = 185 mg g−1), increasing isoelectric points (IEP: 5.8 enhanced 7.2) and promoting nanoparticle agglomeration (hydrodynamic diameter: 45 enhanced 210 nm) [104].

4.2. Dynamic Optimization Strategy of the Coordination Structure of MXene

MXene-based composites demonstrate excellent photocatalytic performance and stability through sophisticated interface engineering. In Pd/MXene/MIL-101(Fe) heterojunction, synergistic interfacial effects significantly enhanced visible light absorption (>500 nm) and charge separation efficiency, achieving 98% efficient degradation of ofloxacin [137]. MXene’s surface functional groups (-OH, -F) optimize electronic coupling with MIL-101(Fe), while Pd modulates OFL intermediate binding energies, enabling pH-selective adsorption and radical-dominated degradation. Pd/MXOF maintains >90% efficiency across pH 4–10 due to MXene’s pH-responsive surface charge and MIL-101(Fe)’s adsorption enrichment.
Using two-dimensional Ti3C2 MXene as the matrix material, in the presence of a crystal plane controlling agent, the (001)-facet-exposed TiO2/Ti3C2 MXene photocatalyst was prepared by a hydrothermal reaction method at 160 °C [59]. The formation of the heterojunction on the TiO2 surface, the Schottky junction at the interface between (001)-facet TiO2 and Ti3C2 MXene, and the LSPR effect of Ti3C2 MXene on high-energy hot electrons, hot holes, and the local magnetic field synergistically promoted the spatial separation of photogenerated electrons and holes. The carrier complexation was effectively inhibited, and the near-infrared light and full-spectrum catalytic activities of the (001) TiO2/Ti3C2 photocatalyst were improved [59].
Through the interface reconstruction between carbon dots and MXene-Ti3C2/CeO2, a three-phase micro-heterojunction was formed, enhancing the local active site density. The optimized catalyst increased the ammonia production rate by 8 times under visible light, which is attributed to the dynamic matching of oxygen vacancies and surface charges [138]. When the concentration of humic acid reached 100 mg L−1, the photocatalytic efficiency only decreased by 5%, indicating the anti-inhibition ability of MXene-based materials against natural organic matter.

5. Research Prospects and Outlook

5.1. Opportunities and Challenges

Although MXene shows unique advantages in the field of photocatalysis, its practical application still faces the following core challenges:
① The preparation of MXene mainly relies on the etching of MAX-phase precursors with hydrofluoric acid (HF) or fluorine salts (such as LiF + HCl). This method has problems such as complex processes, low yields (<50%), and toxic by-products (such as AlF3). In addition, the lateral size of MXene nanosheets is usually 1–5 μm, and it is difficult to achieve large-area (>10 cm2) uniform film formation through existing technologies (such as liquid-phase exfoliation). Studies have shown that the defect density of MXene films increases exponentially with the increase in area, resulting in a significant decrease in photocatalytic activity (for example, the hydrogen production efficiency of a 10-cm2 film is only 30% of that of small-piece samples).
② The rich active functional groups (-OH, -O) on the surface of MXene endow it with excellent adsorption properties but also make it prone to irreversible oxidative degradation in humid, high-temperature, or oxidative environments. For example, after Ti3C2Tx is exposed to air for 72 h, TiO2 nanoparticles are formed on the surface due to oxidation, the electrical conductivity decreases by >80%, and the photocatalytic activity is lost by >60%. In wastewater containing Cl or SO42−, ion intercalation easily occurs between the layers of MXene, accelerating the structural collapse.
③ The wide-spectrum absorption (ultraviolet-visible-near-infrared) characteristics of MXene are often accompanied by a high carrier recombination rate (lifetime <10 ps), and its metallic surface leads to the rapid recombination of photogenerated electron-hole pairs. Although the heterojunction design can partially alleviate this problem, interface defects (such as dangling bonds, lattice mismatch) still limit the improvement of quantum efficiency (usually <20%).
In response to the above challenges, the following breakthrough strategies have been proposed in recent studies:
① Coating an inert protective layer (such as SiO2, Al2O3) on the surface of MXene by atomic layer deposition (ALD) or sol–gel method can significantly improve its environmental stability. For example, SiO2-coated Ti3C2Tx (SiO2@Ti3C2Tx) reduces the degree of oxidation by 90% after being stored at 85% humidity for 30 days, and the efficiency of photocatalytic degradation of tetracycline remains 85% of the initial value. The mechanism lies in that the SiO2 layer (with a thickness of 2–5 nm) blocks the penetration of H2O and O2 while retaining the surface-active sites of MXene.
② Introducing sulfur vacancies (S-vacancy) or nitrogen doping (N-doping) into MXene through plasma treatment or chemical doping can optimize its energy band structure and carrier dynamics. For example, the hydrogen production rate of sulfur-vacancy-modified Mo2CTx (Mo2CTx-Sv) under visible light reaches 8.2 mmol·g−1·h−1, which is 3.5 times that of the original sample. Density functional theory (DFT) calculations show that sulfur vacancies act as electron traps, extending the carrier lifetime to 50 ps, and improving catalytic selectivity by reducing the adsorption energy of H intermediates (ΔGH decreases from 0.45 eV to 0.18 eV).
③ Developing fluorine-free etchants (such as the NaOH melting method) and continuous roll-to-roll film-forming technology can consider both environmental protection and large-scale production. For example, the yield of Nb2CTx synthesized by the KOH-assisted hydrothermal method is increased to 78%, and the emission of toxic fluorides is avoided. In addition, the bionic dynamic interface design (such as MXene/hydrogel composites) realizes pH-temperature-responsive structural reorganization through the self-repair ability of the hydrogen-bond network, maintaining the stability of catalytic activity under extreme conditions (pH = 2–12).

5.2. Take a Reasonable Look at the Future of MXene

MXene’s 2D architecture offers unparalleled opportunities for composite photocatalyst engineering through three synergistic pathways: (1) As an oxidative growth platform, controlled partial oxidation of Ti3C2 enables in situ TiO2 heterojunction formation, though precise regulation of oxidation kinetics and dopant distribution remains critical to optimize optical band structures and interfacial electron transfer rates [139]. (2) Terminal group (-OH/-O/-F) modulation via selective etching or post-synthetic functionalization dynamically tunes MXene’s work function, enabling adaptive Schottky barrier engineering with semiconductors like g-C3N4. (3) Environment-responsive surface restructuring—protonation in acidic media (pH < 4) or oxidative radical attack during photocatalysis—alters Ti redox states and photothermal conversion efficiency. While DFT simulations reveal substrate-specific electronic state modulation, practical deployment requires overcoming HF-dependent synthesis limitations (toxic AlF3 byproducts, <50% yield) through eco-friendly molten-salt alternatives (KOH/NaOH etching, 75–85% purity). Multiscale modeling spanning electronic interactions to reactor-level photothermal-fluidic coupling is essential to decode MXene’s adaptive functionality in complex aqueous matrices and accelerate solar-driven water remediation.

6. Conclusions

MXenes, a unique class of two-dimensional transition metal carbides, nitrides, and carbonitrides, have emerged as highly promising photocatalytic materials owing to their large specific surface area, tunable surface terminations, and metallic-like conductivity. Advances in etching technologies and post-synthetic modification have enabled precise regulation of MXenes’ morphology, interlayer spacing, and surface chemistry, which in turn determine their electronic structures, charge-transfer dynamics, and catalytic efficiency.
Functionally, MXenes can serve both as electron acceptors and as efficient charge-transport mediators, thereby effectively suppressing electron–hole recombination and enhancing the performance of semiconductor-based heterojunctions. These distinctive properties endow MXenes with broad applicability: they can act as cocatalytic platforms for pollutant degradation and antibiotic removal, as structural templates for hierarchical architectures, and as core components in Z-scheme photocatalytic systems. Of particular importance in polymer-based photocatalysis, MXenes’ adaptable interfacial characteristics—governed by protonation/deprotonation kinetics, electrochemical polarization effects, and environmental conditions—provide a new pathway for constructing responsive and programmable catalytic processes. By integrating MXenes into polymer matrices, one can achieve synergistic effects that enhance light harvesting, charge separation, and catalytic durability, thus enabling multifunctional and sustainable photocatalytic platforms.
Despite these advances, several challenges remain. MXenes are prone to oxidative degradation, face difficulties in large-scale synthesis, and often exhibit limited stability under complex aqueous conditions. To overcome these limitations, innovative strategies are required, such as applying protective surface coatings, engineering defects/dopants, developing environmentally friendly fluorine-free synthesis routes, and learning from biological systems to design dynamic interfaces. Meanwhile, the integration of density functional theory (DFT), multiscale simulations, and machine learning provides a powerful bridge between theoretical predictions and experimental outcomes, accelerating the rational design and application of MXene-based systems.
Looking ahead, MXene–polymer composites hold great potential to advance from laboratory studies to scalable practical applications. Their unique adaptability, arising from the interplay of electronic structures and surface chemistry, positions MXenes as transformative building blocks for next-generation high-performance and sustainable photocatalytic systems in environmental remediation, solar-driven energy conversion, and advanced separation technologies.

Author Contributions

Methodology, Z.C. and Z.M.; software, Z.C., Writing—Original Draft, Z.C.; Methodology, Z.M.; Editing, Z.M. and Z.Z.; Formal analysis, Z.Z.; Resources, Writing—Review and Editing, Supervision, Data Curation, W.M.; All authors have read and agreed to the published version of the manuscript.

Funding

The authors acknowledge the financial support received from the Beijing Science and Technology Commission (No. Z211100004321001).

Institutional Review Board Statement

Not applicable.

Data Availability Statement

Data sharing is not applicable (only appropriate if no new data is generated or the article describes entirely theoretical research).

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Yi, L.; Meng, Z.; Yang, W.; Li, S.; Hu, J.; Sun, W.; Ni, J. Profiles, drivers, and prioritization of antibiotics in China’s major rivers. J. Hazard. Mater. 2024, 477, 135399. [Google Scholar] [CrossRef]
  2. Verma, M.; Haritash, A.K. Photocatalytic degradation of Amoxicillin in pharmaceutical wastewater: A potential tool to manage residual antibiotics. Environ. Technol. Innov. 2020, 20, 101072. [Google Scholar] [CrossRef]
  3. Rueda-Marquez, J.J.; Levchuk, I.; Ibañez, P.F.; Sillanpää, M. A critical review on application of photocatalysis for toxicity reduction of real wastewaters. J. Clean. Prod. 2020, 258, 120694. [Google Scholar] [CrossRef]
  4. Latiful, K.; David, N.; Rachadaporn, B.; Chanthai, S.; Otgonbayar, Z.; Oh, W.-C. Photocatalytic removal of pharmaceutical antibiotics induced pollutants by MXene-based composites: Comprehensive review. Sustain. Mater. Technol. 2024, 41, e01083. [Google Scholar] [CrossRef]
  5. Feng, Y.; Gong, S.; Wang, Y.; Ban, C.G.; Qu, X.L.; Ma, J.P.; Duan, Y.Y.; Lin, C.; Yu, D.M.; Xia, L.; et al. Noble-Metal-Free Cocatalysts Reinforcing Hole Consumption for Photocatalytic Hydrogen Evolution with Ultrahigh Apparent Quantum Efficiency. Adv. Mater. 2025, 37, 2412965. [Google Scholar] [CrossRef]
  6. Li, Y.-N.; Chen, Z.-Y.; Wang, M.-Q.; Zhang, L.-z.; Bao, S.-J. Interface engineered construction of porous g-C3N4/TiO2 heterostructure for enhanced photocatalysis of organic pollutants. Appl. Surf. Sci. 2018, 440, 229–236. [Google Scholar] [CrossRef]
  7. Neeraj, S.; Stuti, T.; Mohd, S.; Kant Choubey, R.; Singh, A. Study of Optical and Electrical Properties of Graphene Oxide. Mater. Today Proc. 2021, 36, 730–735. [Google Scholar]
  8. Mamba, G.; Gangashe, G.; Moss, L.; Hariganesh, S.; Thakur, S.; Vadivel, S.; Mishra, A.K.; Vilakati, G.D.; Muthuraj, V.; Nkambule, T.T.I. State of the art on the photocatalytic applications of graphene based nanostructures: From elimination of hazardous pollutants to disinfection and fuel generation. J. Environ. Chem. Eng. 2020, 8, 103505. [Google Scholar] [CrossRef]
  9. Fu, Y.; Yao, X.; Ji, X.; Zhou, L.; Tong, Y.; Zhang, R.; Wang, X. Defective ordered macroporous ZIF–8/ZnO heterostructure for enhanced visible–light photo–oxidation performance. J. Alloys Compd. 2024, 983, 173899. [Google Scholar] [CrossRef]
  10. Chen, S.; Zhang, L.; Alshammari, D.A.; Hessien, M.M.; Yu, W.; Cui, L.; Ren, J.; El-Bahy, Z.M.; Guo, Z. Z-scheme Ag2O/ZnO heterostructure on carbon fibers for efficient photocatalysis of tetracycline. Sep. Purif. Technol. 2025, 354, 129414. [Google Scholar] [CrossRef]
  11. Lu, H.; Yang, J.; Xu, J. Study of the Photocatalytic Degradation of Tetracycline by BiOI/Brookite TiO2 Composites. Hans J. Chem. Eng. Technol. 2024, 14, 83–88. (In Chinese) [Google Scholar] [CrossRef]
  12. Wang, L.; Tang, M.; Wang, H.; Chen, L.; Long, Z.; Gao, D. Study on Photocatalytic Degradation of Tetracycline by Ag2CO3/ZIF-8/CF Composites. J. Yancheng Inst. Technol. (Nat. Sci. Ed.) 2023, 36, 7–14. (In Chinese) [Google Scholar]
  13. Lerdwiriyanupap, T.; Waehayee, A.; Choklap, T.; Prachanat, J.; Nakajima, H.; Chankhanittha, T.; Butburee, T.; Siritanon, T. CeO2/BiYO3 photocatalyst for the degradation of tetracycline under visible light irradiation. Ceram. Int. 2024, 50, 52723–52732. [Google Scholar] [CrossRef]
  14. Zhao, Q.; Wang, J.; Li, Z.; Guo, Y.; Tang, B.; Abudula, A.; Guan, G. Two-dimensional Ti3C2TX-nanosheets/Cu2O composite as a high-performance photocatalyst for decomposition of tetracycline. Carbon Resour. Convers. 2021, 4, 197–204. [Google Scholar] [CrossRef]
  15. Talreja, N.; Ashfaq, M.; Chauhan, D.; Viswanathan, M.R. PVP encapsulated MXene coated on PET surface (PMP)-based photocatalytic materials: A novel photo-responsive assembly for the removal of tetracycline. Environ. Res. 2023, 233, 116439. [Google Scholar] [CrossRef] [PubMed]
  16. Sun, J.; Farid, M.U.; Boey, M.W.; Sato, Y.; Chen, G.; An, A.K. MXene-PVA-TiO2-based photothermal-catalytic membrane with high structural stability for efficient desalination and photodegradation. Chem. Eng. J. 2023, 468, 143744. [Google Scholar] [CrossRef]
  17. Atta, M.M.; Elbasiony, A.M.; Henaish, A.M.A.; Zhang, Q. Enhancement of optical and photocatalytic properties of polyethylene oxide/polyvinylpyrrolidone blend reinforced with Ti3C2 MXene. Synth. Met. 2024, 307, 117698. [Google Scholar] [CrossRef]
  18. Naguib, M.; Mashtalir, O.; Carle, J.; Presser, V.; Lu, J.; Hultman, L.; Gogotsi, Y.; Barsoum, M.W. Two-Dimensional Transition Metal Carbides. ACS Nano 2012, 6, 1322–1331. [Google Scholar] [CrossRef]
  19. Sengupta, J.A.-O.; Hussain, C.A.-O. MXene-Based Electrochemical Biosensors: Advancing Detection Strategies for Biosensing (2020–2024). Biosensors 2025, 15, 127. [Google Scholar] [CrossRef]
  20. Zhu, M.; Cai, C.; Lin, X.; Yang, R.; Lin, X.; Li, W.; Li, C.; Gao, F.; Zhang, J. Advances in the application of MXenes and MXene-composites in MXene-based lithium anodes for lithium metal batteries. Mater. Today Chem. 2025, 48, 102992. [Google Scholar] [CrossRef]
  21. Bu, Y.; Wang, Y.; Wu, H.; Wang, H.; Wang, B.; Yang, Y.; Huang, L.; Tang, J. Recent advances in flexible piezoresistive pressure sensors based on MXene materials. Mater. Today Chem. 2025, 48, 102984. [Google Scholar] [CrossRef]
  22. Lin, J.; Chang, C.-W.; Chang, C.-T.; Manibalan, K.; Tsai, T.-H.; Lin, Y.-C.; Hsu, Y.-S.; Chang, M.-H.; Chen, J.-T. Wearable and stretchable electronics enabled by MXene–PVA films via layer-by-layer assembly with integrated UV, thermal, and mechanical sensing†. Chem. Commun. 2025, 61, 10578–10581. [Google Scholar] [CrossRef]
  23. Serafin, J.; Dziejarski, B.; Achieng, G.O.; Vendrell, X.; Chaitoglou, S.; Amade-Rovira, R. Comprehensive analysis of MAX phase and MXene materials for advanced photocatalysis, electrocatalysis and adsorption in hydrogen evolution and storage. J. Ind. Eng. Chem. 2025, 142, 18–33. [Google Scholar] [CrossRef]
  24. Carey, M.; Barsoum, M.W. MXene polymer nanocomposites: A review. Mater. Today Adv. 2021, 9, 100120. [Google Scholar] [CrossRef]
  25. Hussain, S.N.; Rao, K.R.; Ali, M.S.; Mubarak, N.M.; Jatoi, A.S.; Malafaia, G.; Azad, A.K. MXene as emerging material for photocatalytic degradation of environmental pollutants. Coord. Chem. Rev. 2023, 477, 214965. [Google Scholar]
  26. Li, K.; Zhang, S.; Li, Y.; Fan, J.; Lv, K. MXenes as noble-metal-alternative co-catalysts in photocatalysis. Chin. J. Catal. 2021, 42, 3–14. [Google Scholar] [CrossRef]
  27. Ghanbari, R.; Zare, E.N. Engineered MXene-polymer composites for water remediation: Promises, challenges and future perspective. Coord. Chem. Rev. 2024, 518, 216089. [Google Scholar] [CrossRef]
  28. Fatima, S.; Hakim, M.W.; Zheng, X.; Sun, Y.; Li, Z.; Han, N.; Li, M.; Wang, Z.; Han, L.; Wang, L.; et al. Constructing nitrogen-doped graphene quantum dots/tantalum carbide MXene heterojunctions as bifunctional catalysts for efficient water splitting. Int. J. Hydrogen Energy 2025, 117, 420–429. [Google Scholar] [CrossRef]
  29. Wu, D.; Tong, Z.; Wang, J.; Chen, G.; Zhu, A.; Tong, H.; Dang, M. Hole transfer promotion on La-MOF@MXene by valence band structure regulation for the efficient photocatalytic degradation. Sep. Purif. Technol. 2025, 359, 130873. [Google Scholar] [CrossRef]
  30. Jin, P.; Han, P.; Li, X.; Li, K. Titanium carbide MXenes-initiated plasmonic metal-support interaction for effective photocatalysis on uncoated Ag nanoparticles. Appl. Surf. Sci. 2023, 612, 155850. [Google Scholar] [CrossRef]
  31. Jian, M.; Wang, Y. Construction of MXenes/COFs heterojunctions for photo/electrocatalytic applications: Opportunities and challenges. Chem. Eng. J. 2025, 507, 160631. [Google Scholar] [CrossRef]
  32. Huanhuan, S.; Panpan, Z.; Zaichun, L.; Park, S.; Lohe, M.R.; Wu, Y.; Shaygan Nia, A.; Yang, S.; Feng, X. Ambient-Stable Two-Dimensional Titanium Carbide (MXene) Enabled by Iodine Etching. Angew. Chem. Int. Ed. 2021, 60, 8689–8693. [Google Scholar]
  33. Wang, D.; Zhou, C.; Filatov, A.S.; Cho, W.J.; Lagunas, F.; Wang, M.Z.; Vaikuntanathan, S.; Liu, C.; Klie, R.F.; Talapin, D.V. Direct synthesis and chemical vapor deposition of 2D carbide and nitride MXenes. Science 2023, 379, 1242–1247. [Google Scholar] [CrossRef] [PubMed]
  34. Ahmad, M.M.; Ma, Y.; Badshah, M.; Ali, S.; Idrees, M.; Ismail, M.A.; Khan, S.; Javed, M.S.; Tamang, T.L.; Pu, Q. 2D-MXene: Progress in Synthesis, Intercalation, and Applications in Microfluidic Sensors. Surf. Interfaces 2024, 56, 105678. [Google Scholar] [CrossRef]
  35. Sun, Y.; Sun, Y.; Meng, X.; Gao, Y.; Dall’Agnese, Y.; Chen, G.; Dall’Agnese, C.; Wang, X.F. Eosin Y-sensitized partially oxidized Ti3C2 MXene for photocatalytic hydrogen evolution. Catal. Sci. Technol. 2019, 9, 310–315. [Google Scholar] [CrossRef]
  36. Singh, S.; Dharavath, S.; Kodali, S.; Dash, R.K. The role of delaminating agents on the structure, morphology, bonding and electrical properties of HF etched MXenes. FlatChem 2025, 49, 100806. [Google Scholar] [CrossRef]
  37. Peng, C.; Yang, X.; Li, Y.; Yu, H.; Wang, H.J.; 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]
  38. Gentile, A.; Ferrara, C.; Tosoni, S.; Balordi, M.; Marchionna, S.; Cernuschi, F.; Kim, M.-H.; Lee, H.-W.; Ruffo, R. Enhanced Functional Properties of Ti3C2T MXenes as Negative Electrodes in Sodium-Ion Batteries by Chemical Tuning. Small Methods 2020, 4, 2000314. [Google Scholar] [CrossRef]
  39. Zhang, P.; Zheng, Q.; Bashir, T.; Ali, T.; Khan, S.; Alenad, A.M.; Raza, S. Reliance of MXene terminating groups on various synthetic strategies and its hot electron dynamics at MXene interfaces. J. Environ. Chem. Eng. 2024, 12, 114708. [Google Scholar] [CrossRef]
  40. Sang, X.; Xie, Y.; Lin, M.W.; Alhabeb, M.; Van Aken, K.L.; Gogotsi, Y.; Kent, P.R.C.; Xiao, K.; Unocic, R.R. Atomic Defects in Monolayer Titanium Carbide (Ti3C2Tx) MXene. ACS Nano 2016, 10, 9193–9200. [Google Scholar] [CrossRef]
  41. Xie, X.; Zhang, N. Positioning MXenes in the Photocatalysis Landscape: Competitiveness, Challenges, and Future Perspectives. Adv. Funct. Mater. 2020, 30, 2002528. [Google Scholar] [CrossRef]
  42. Miao, B.; Bashir, T.; Zhang, H.; Ali, T.; Raza, S.; He, D.; Bai, J. Impact of various 2D MXene surface terminating groups in energy conversion. Renew. Sustain. Energy Rev. 2024, 199, 114506. [Google Scholar] [CrossRef]
  43. Kahila, B.; Ehsan, G.; Saleem, R.; Babenko, A.; Paimard, G.; Bashir, T.; Maleki-Ghaleh, H.; Jie, L.; Orooji, Y. Recent advancements in MXenes synthesis, properties, and cutting-edge applications: A comprehensive review. J. Environ. Chem. Eng. 2024, 12, 113546. [Google Scholar] [CrossRef]
  44. Paria, E.; Aydin, H.; Stanisław, W.; Kun-Yi, A.; Zahra, S.; Ghanbari, F. Recent advances in design and engineering of MXene-based catalysts for photocatalysis and persulfate-based advanced oxidation processes: A state-of-the-art review. Chem. Eng. J. 2024, 480, 147920. [Google Scholar]
  45. Wang, S.; Du, Y.L.; Liao, W.H. Tunable band gap and optical properties of surface functionalized Sc2C monolayer. Chin. Phys. B 2017, 26, 17806. [Google Scholar] [CrossRef]
  46. Hassan, N.S.; Jalil, A.A.; Bahari, M.B.; Izzuddin, N.M.; Fauzi, N.A.F.M.; Jusoh, N.W.C.; Kamaroddin, M.F.A.; Saravanan, R.; Tehubijuluw, H. A critical review of MXene-based composites in the adsorptive and photocatalysis of hexavalent chromium removal from industrial wastewater. Environ. Res. 2024, 259, 119584. [Google Scholar] [CrossRef]
  47. He, Y.; Chen, X.; Wu, Z.; Xue, Q.; Tian, F. In situ fabrication of N-doped Ti3C2Tx-MXene-modified BiOBr Schottky heterojunction with high photoelectron separation efficiency for enhanced photocatalytic ammonia synthesis. J. Alloys Compd. 2023, 969, 172470. [Google Scholar] [CrossRef]
  48. Liu, X.; Xu, F.; Li, Z.; Liu, Z.; Yang, W.; Zhang, Y.; Yang, H.Y. Design strategy for MXene and metal chalcogenides/oxides hybrids for supercapacitors, secondary batteries and electro/photocatalysis. Coord. Chem. Rev. 2022, 464, 214544. [Google Scholar] [CrossRef]
  49. Soyoung, P.; Sewoon, K.; Yeonji, Y.; Saravanakumar, K.; Lee, E.; Yoon, Y.; Park, C.M. Adsorptive and photocatalytic performance of cobalt-doped ZnTiO3/Ti3C2Tx MXene nanohybrids towards tetracycline: Kinetics and mechanistic insight. J. Hazard. Mater. 2023, 443, 130165. [Google Scholar]
  50. Rasheed, A.; El Sayed, M.E.; Yousaf, S.; Samir, A.; Warsi, M.F.; El-Bahy, Z.M. Fluorine-doped cobalt ferrite integrated into 2D MXene with extended solar spectrum response and boosted charge separation for the removal of recalcitrant organic pollutants. Results Phys. 2024, 67, 108049. [Google Scholar] [CrossRef]
  51. Yu, X.-f.; Cheng, J.-b.; Liu, Z.-b.; Li, Q.-z.; Li, W.-z.; Yang, X.; Xiao, B. The band gap modulation of monolayer Ti2CO2 by strain. RSC Adv. 2015, 5, 30438–30444. [Google Scholar] [CrossRef]
  52. Ding, Y.M.; Nie, X.; Dong, H.A.-O.; Rujisamphan, N.; Li, Y.A.-O. Many-body effects in an MXene Ti2CO2 monolayer modified by tensile strain: GW-BSE calculations. Nanoscale Adv. 2020, 2, 2471–2477. [Google Scholar] [CrossRef]
  53. Kandemir, Z.; Torun, E.; Paleari, F.; Yelgel, C.; Sevik, C. Surface termination dependence of electronic and optical properties in Ti2CO2 MXene monolayers. Phys. Rev. Mater. 2022, 6, 026001. [Google Scholar] [CrossRef]
  54. Hart, J.A.-O.; Hantanasirisakul, K.A.-O.; Lang, A.C.; Anasori, B.; Pinto, D.; Pivak, Y.; van Omme, J.A.-O.; May, S.J.; Gogotsi, Y.; Taheri, M.L. Control of MXenes’ electronic properties through termination and intercalation. Nat. Commun. 2019, 10, 522. [Google Scholar] [CrossRef]
  55. Halim, J.; Persson, I.; Eklund, P.; Persson, P.O.Å.; Rosen, J. Sodium hydroxide and vacuum annealing modifications of the surface terminations of a Ti3C2 (MXene) epitaxial thin film. RSC Adv. 2018, 8, 36785–36790. [Google Scholar] [CrossRef]
  56. Iravani, S.A.-O.; Varma, R.A.-O. MXene-Based Photocatalysts in Degradation of Organic and Pharmaceutical Pollutants. Molecules 2022, 27, 6939. [Google Scholar] [CrossRef] [PubMed]
  57. Irfan, S.; Khan, S.A.-O.; Din, M.A.U.; Dong, F.; Chen, D. Retrospective on Exploring MXene-Based Nanomaterials: Photocatalytic Applications. Molecules 2023, 28, 2495. [Google Scholar] [CrossRef]
  58. He, X.; Kai, T.; Ding, P.A.-O. Heterojunction photocatalysts for degradation of the tetracycline antibiotic: A review. Environ. Chem. Lett. 2021, 19, 4563–4601. [Google Scholar] [CrossRef] [PubMed]
  59. Wang, Y.; Tan, G.; Feng, S.; Liu, Y.; Liu, W.; Xia, A.; Ren, H.; Lv, L. Preparation of exposed (001) facets TiO2/Ti3C2 MXene and performance of near infrared light photodegradation. Ceram. Int. 2023, 49, 7258–7265. [Google Scholar] [CrossRef]
  60. Lai, C.; Yan, B.; Yuan, R.; Chen, D.; Wang, X.; Wang, M.; He, H.; Tu, J.A.-O. In situ growth of TiO2/Ti3C2 MXene Schottky heterojunction as a highly sensitive photoelectrochemical biosensor for DNA detection. RSC Adv. 2023, 13, 16222–16229. [Google Scholar] [CrossRef] [PubMed]
  61. Ding, W.; Deng, H.; Guo, S.; Li, L.; Wang, P.; Zhu, H.; Hao, D.; Li, C.; Tang, X.; Wang, Q.; et al. MXene-based composites for adsorption and photocatalytic immobilization of U(VI): Current progress and future perspectives. Sep. Purif. Technol. 2025, 377, 134282. [Google Scholar] [CrossRef]
  62. ZHENG, Z.X.; WANG, E.H.; HOU, X.M.; YANG, T. University of Science and Technology Beijing. The stability and improvement of two-dimensional transition metal carbides and/or carbonitrides (MXene). J. Eng. Sci. 2022, 44, 1881–1896. (In Chinese) [Google Scholar]
  63. Zhu, H.; Liang, Z.; Xue, S.; Ren, X.; Liang, X.; Xiong, W.; Gao, L.; Liu, A. DFT practice in MXene-based materials for electrocatalysis and energy storage: From basics to applications. Ceram. Int. 2022, 48, 27217–27239. [Google Scholar] [CrossRef]
  64. Khazaei, M.; Arai, M.; Sasaki, T.; Chung, C.-Y.; Venkataramanan, N.S.; Estili, M.; Sakka, Y.; Kawazoe, Y. Novel Electronic and Magnetic Properties of Two-Dimensional Transition Metal Carbides and Nitrides. Adv. Funct. Mater. 2013, 23, 2185–2192. [Google Scholar] [CrossRef]
  65. Wang, Q.; Xie, H.; Nie, Y.-H.; Ren, W. Enhancement of thermoelectric efficiency in triple quantum dots by the Dicke effect. Phys. Rev. B 2013, 87, 075102. [Google Scholar] [CrossRef]
  66. Shein, I.R.; Ivanovskii, A.L. Graphene-like titanium carbides and nitrides Tin+1Cn, Tin+1Nn (n = 1, 2, and 3) from de-intercalated MAX phases: First-principles probing of their structural, electronic properties and relative stability. Comput. Mater. Sci. 2012, 65, 104–114. [Google Scholar] [CrossRef]
  67. Low, J.; Zhang, L.; Tong, T.; Shen, B.; Yu, J. TiO2/MXene Ti3C2 composite with excellent photocatalytic CO2 reduction activity. J. Catal. 2018, 361, 255–266. [Google Scholar] [CrossRef]
  68. Li, J.; Peng, H.; Luo, B.; Cao, J.; Ma, L.; Jing, D. The enhanced photocatalytic and photothermal effects of Ti3C2 Mxene quantum dot/macroscopic porous graphitic carbon nitride heterojunction for Hydrogen Production. J. Colloid Interface Sci. 2023, 641, 309–318. [Google Scholar] [CrossRef] [PubMed]
  69. Oyehan, T.A.; Salami, B.A.; Abdulrasheed, A.A.; Hambali, H.U.; Gbadamosi, A.; Valsami-Jones, E.; Saleh, T.A. MXenes: Synthesis, properties, and applications for sustainable energy and environment. Appl. Mater. Today 2023, 35, 101993. [Google Scholar] [CrossRef]
  70. Xin, L.; Guohong, W.; Qiang, L.; Wang, Y.; Lu, X. Dual optimized Ti3C2Tx MXene@ZnIn2S4 heterostructure based on interface and vacancy engineering for improving electromagnetic absorption. Chem. Eng. J. 2023, 453, 139488. [Google Scholar]
  71. Tao, C.; Longlu, W.; Yutang, L.; 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]
  72. Zeng, W.; Ye, X.; Dong, Y.; Zhang, Y.; Sun, C.; Zhang, T.; Guan, X.; Guo, L. MXene for photocatalysis and photothermal conversion: Synthesis, physicochemical properties, and applications. Coord. Chem. Rev. 2024, 508, 215753. [Google Scholar] [CrossRef]
  73. Wang, H.; Li, X.; Zhao, X.; Li, C.; Song, X.; Zhang, P.; Huo, P. A review on heterogeneous photocatalysis for environmental remediation: From semiconductors to modification strategies. Chin. J. Catal. 2022, 43, 178–214. [Google Scholar] [CrossRef]
  74. Meng, X.; Wang, L.; Wang, X.; Zhen, M.; Hu, Z.; Guo, S.Q.; Shen, B. Recent developments and perspectives of MXene-Based heterostructures in photocatalysis. Chemosphere 2023, 338, 139550. [Google Scholar] [CrossRef]
  75. Li, Z.; Xiong, J.; Song, H.; Liu, S.; Huang, Y.; Huang, Y.; Luo, X. Synergistically enhancing CO2 adsorption/activation and electron transfer in ZIF-67/Ti3C2Tx MXene for boosting photocatalytic CO2 reduction. Sep. Purif. Technol. 2024, 341, 126817. [Google Scholar] [CrossRef]
  76. Khesali, A.S.; Mahdiyeh, Z.; Saeid, A.; Fooladloo, M.A. Investigation of the optical and electronic properties of functionalized Ti3C2 Mxene with halid atoms using DFT calculation. Mater. Today Commun. 2023, 35, 106136. [Google Scholar] [CrossRef]
  77. Yao, X.; Yan-Fang, Z.; Wei, X.; Wu, Z.-G.; Li, Y.-C.; Lai, J.; Li, S.; Wang, E.; Yang, Z.-G.; Xu, C.-L.; et al. Deciphering an Abnormal Layered-Tunnel Heterostructure Induced by Chemical Substitution for the Sodium Oxide Cathode. Angew. Chem. Int. Ed. 2020, 59, 1491–1495. [Google Scholar]
  78. Shi, A.; Sun, D.; Zhang, X.; Ji, S.; Wang, L.; Li, X.A.; Niu, X. Direct Z-Scheme Photocatalytic System: Insights into the Formative Factors of Photogenerated Carriers Transfer Channel from Ultrafast Dynamics. ACS Catal. 2022, 12, 9570–9578. [Google Scholar] [CrossRef]
  79. Huang, X.; Yin, K.; Zhang, S.; Wu, T.; Yuan, Y.; Wang, X.; Wang, D. Interfacial chemical bond-modulated Z-scheme Cs2AgBiBr6/WO3 enables stable and highly efficient photocatalysis. Appl. Surf. Sci. 2023, 637, 157877. [Google Scholar] [CrossRef]
  80. Yuan, H.; Xiao, W.; Zhang, X.; Bao, J.; Li, W.; Huang, B.; He, G. Ti3C2 MXene flakes assisted Mn0.5Cd0.5S/Ti3C2 MXene/g-C3N4 Z-scheme heterojunction for enhancing photocatalytic hydrogen evolution. J. Power Sources 2024, 606, 234588. [Google Scholar] [CrossRef]
  81. Zhou, Y.; Yu, M.; Liang, H.; Chen, J.; Xu, L.; Niu, J. Novel dual-effective Z-scheme heterojunction with g-C3N4, Ti3C2 MXene and black phosphorus for improving visible light-induced degradation of ciprofloxacin. Appl. Catal. B Environ. 2021, 291, 120105. [Google Scholar] [CrossRef]
  82. Karunamoorthy, S.; Keunyoung, Y.; Velusamy, M.; Yea, Y.; Jagan, G.; Park, C.M. Construction of novel In2S3/Ti3C2 MXene quantum dots/SmFeO3 Z-scheme heterojunctions for efficient photocatalytic removal of sulfamethoxazole and 4-chlorophenol: Degradation pathways and mechanism insights. Chem. Eng. J. 2023, 451, 138933. [Google Scholar]
  83. Liu, Y.; Li, Y.-H.; Li, X.; Zhang, Q.; Yu, H.; Peng, X.; Peng, F. Regulating Electron–Hole Separation to Promote Photocatalytic H2 Evolution Activity of Nanoconfined Ru/MXene/TiO2 Catalysts. ACS Nano 2020, 14, 14181–14189. [Google Scholar] [CrossRef] [PubMed]
  84. 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]
  85. 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]
  86. Chaudhuri, K.; Alhabeb, M.; Wang, Z.; Shalaev, V.M.; Gogotsi, Y.; Boltasseva, A. Highly Broadband Absorber Using Plasmonic Titanium Carbide (MXene). ACS Photonics 2018, 5, 1115–1122. [Google Scholar] [CrossRef]
  87. Li, C.; Kan, C.; Meng, X.; Liu, M.; Shang, Q.; Yang, Y.; Wang, Y.; Cui, X. Self-Assembly 2D Ti3C2/g-C3N4 MXene Heterojunction for Highly Efficient Photocatalytic Degradation of Tetracycline in Visible Wavelength Range. Nanomaterials 2022, 12, 4015. [Google Scholar] [CrossRef]
  88. Nasri, M.S.; Samsudin, M.F.; 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]
  89. 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]
  90. Hu, X.; Li, G.; Zhang, Y.; Lu, M.; Pu, W.; Dai, Y.; Wang, H. A novel iron oxide (Fe3O4)-laden titanium carbide (Ti3C2) MXene stacks for the efficient removal of tetracycline from aqueous solution. Chemosphere 2024, 364, 143210. [Google Scholar] [CrossRef] [PubMed]
  91. Xu, Q.; Wang, P.; Wang, Z.; Shen, J.; Han, X.; Zheng, X.; Song, K. Aerosol self-assembly synthesis of g-C3N4/MXene/Ag3PO4 heterostructure for enhanced photocatalytic degradation of tetracycline hydrochloride. Colloids Surf. A Physicochem. Eng. Asp. 2022, 648, 129392. [Google Scholar] [CrossRef]
  92. Cai, J.; Peng, Y.; Hu, P.; Zhou, Y.; Yao, Q.; Lyu, H.; Gao, Y. H2O2-assisted room temperature preparation of crystalline TiO2 and Ti3C2Tx-derived C-doped amorphous TiOx homojunction for photocatalytic degradation of tetracycline. Appl. Surf. Sci. 2025, 680, 161430. [Google Scholar] [CrossRef]
  93. Amari, A.; Aljibori, H.S.S.; Ismail, M.A.; Diab, M.A.; El-Sabban, H.A.; Umarov, A.; Madaminov, S.; Elboughdiri, N. Engineering novel 2D MXene-based dual Z-scheme heterojunction photocatalyst for enhanced TC hydrochloride degradation and hydrogen evolution. J. Water Process Eng. 2025, 70, 107127. [Google Scholar] [CrossRef]
  94. Xu, M.; Peng, M.; Tang, H.; Zhou, W.; Qiao, B.; Ma, D. Renaissance of Strong Metal–Support Interactions. J. Am. Chem. Soc. 2024, 146, 2290–2307. [Google Scholar] [CrossRef]
  95. Yang, Z.; Yang, L.; Liu, Y.; Chen, L. Photocatalytic Deposition of Au Nanoparticles on Ti3C2Tx MXene Substrates for Surface-Enhanced Raman Scattering. Molecules 2024, 29, 2383. [Google Scholar] [CrossRef]
  96. Chen, H.; Yang, Z.; Wang, X.; Polo-Garzon, F.; Halstenberg, P.W.; Wang, T.; Suo, X.; Yang, S.-Z.; Meyer, H.M., III; Wu, Z.; et al. Photoinduced Strong Metal–Support Interaction for Enhanced Catalysis. J. Am. Chem. Soc. 2021, 143, 8521–8526. [Google Scholar] [CrossRef]
  97. Wu, Y.; Li, J.; Sui, G.; Chai, D.F.; Li, Y.; Guo, D.; Chu, D.; Liang, K. Interface and doping engineering of V2C-MXene-based electrocatalysts for enhanced electrocatalysis of overall water splitting. Carbon Energy 2024, 6, e583. [Google Scholar] [CrossRef]
  98. Zhou, Q.; Hong, P.; Shi, X.; Li, Y.; Yao, K.; Zhang, W.; Kong, L. Efficient degradation of tetracycline by a novel nanoconfinement structure Cu2O/Cu@ MXene composite. J. Hazard. Mater. 2023, 448, 130995. [Google Scholar] [CrossRef]
  99. Zheng, Z.; Liu, D.; Sun, X.; Geng, Z.; Wang, Q.; Hou, J. SILAR synthesis of CdS-ZnS/TiO2 NTs for photocatalytic H2 evolution and dye degradation. Colloids Surf. A Physicochem. Eng. Asp. 2025, 707, 135922. [Google Scholar] [CrossRef]
  100. Xiao, R.; Zhao, C.; Zou, Z.; Chen, Z.; Tian, L.; Xu, H.; Yang, X. In situ fabrication of 1D CdS nanorod/2D Ti3C2 MXene nanosheet Schottky heterojunction toward enhanced photocatalytic hydrogen evolution. Appl. Catal. B Environ. 2020, 268, 118382. [Google Scholar] [CrossRef]
  101. Chen, Y.; Zhong, W.; Chen, F.; Wang, P.; Fan, J.; Yu, H. Photoinduced self-stability mechanism of CdS photocatalyst: The dependence of photocorrosion and H2-evolution performance. J. Mater. Sci. Technol. 2022, 121, 19–27. [Google Scholar] [CrossRef]
  102. Tan, Q.; Yu, Z.; Chen, Y.; He, N. Mixed-valent FeWO4-coated 2D Ti3C2 MXene photocatalysts for photo-fenton removal of many common pollutants in water. Ceram. Int. 2024, 50, 16382–16397. [Google Scholar] [CrossRef]
  103. Kashif, R.; Mohamed, H.; Adnan, A.; Ren, C.E.; Gogotsi, Y.; Mahmoud, K.A. Antibacterial Activity of Ti3C2Tx MXene. ACS Nano 2016, 10, 3674–3684. [Google Scholar]
  104. Gao, K.; Hou, L.A.; An, X.; Huang, D.; Yang, Y. BiOBr/MXene/gC3N4 Z-scheme heterostructure photocatalysts mediated by oxygen vacancies and MXene quantum dots for tetracycline degradation: Process, mechanism and toxicity analysis. Appl. Catal. B Environ. 2023, 323, 122150. [Google Scholar] [CrossRef]
  105. Wilson, R.B.; Coh, S. Parametric dependence of hot electron relaxation timescales on electron-electron and electron-phonon interaction strengths. Commun. Phys. 2020, 3, 179. [Google Scholar] [CrossRef]
  106. Ostovar, B.A.-O.; Lee, S.A.-O.X.; Mehmood, A.A.-O.; Farrell, K.A.-O.; Searles, E.A.-O.; Bourgeois, B.A.-O.X.; Chiang, W.A.-O.; Misiura, A.; Gross, N.A.-O.; Al-Zubeidi, A.; et al. The role of the plasmon in interfacial charge transfer. Sci. Adv. 2024, 10, eadp3353. [Google Scholar] [CrossRef]
  107. Liu, J.; Yang, J.; Zhu, G.; Li, J.; Li, Y.; Zhai, Y.; Song, H.; Yang, Y.; Li, H. Revealing the Ultrafast Energy Transfer Pathways in Energetic Materials: Time-Dependent and Quantum State-Resolved. JACS Au 2024, 4, 4455–4465. [Google Scholar] [CrossRef]
  108. Mir, S.H.; Yadav, V.K.; Singh, J.K. Recent Advances in the Carrier Mobility of Two-Dimensional Materials: A Theoretical Perspective. ACS Omega 2020, 5, 14203–14211. [Google Scholar] [CrossRef]
  109. Shin, H.A.-O.; Jeong, W.; Han, T.A.-O. Maximizing light-to-heat conversion of Ti3C2Tx MXene metamaterials with wrinkled surfaces for artificial actuators. Nat. Commun. 2024, 15, 10507. [Google Scholar] [CrossRef]
  110. Wang, T.; Yao, C.; Gao, R.; Holicky, M.; Hu, B.; Liu, S.; Wu, S.; Kim, H.; Ning, H.; Torrisi, F.; et al. Ultrafast Carrier and Lattice Cooling in Ti2CTx MXene Thin Films. Nano Lett. 2024, 24, 16333–16341. [Google Scholar] [CrossRef]
  111. Chang, L.; Huan, Z.; Qinghan, W.; Wu, J.; Chen, P.; Song, S.; Zhan, C.; Wang, L.; Jia, F. Efficient gold recovery from low concentrated Au(S2O3)23—Solution through enhanced photocatalysis via photothermal and surface plasmon resonance assistance on MoS2/MXene/Ag. Desalination 2024, 584, 117733. [Google Scholar]
  112. Wang, D.; Fang, Y.; Yu, W.; Wang, L.; Xie, H.; Yue, Y. Significant solar energy absorption of MXene Ti3C2Tx nanofluids via localized surface plasmon resonance. Sol. Energy Mater. Sol. Cells 2021, 220, 110850. [Google Scholar] [CrossRef]
  113. Niu, X.; Lu, X.; He, B.; Xiao, X.; Wang, S.; Liang, Z.; Ma, L.F. Dual electron transfer path and Ti3C2Tx LSPR photothermal enhancement in 2D/2D/2D Ti3C2Tx MXene@TiO2/CdIn2S4 ternary heterojunction for enhanced photocatalytic H2 production: Construction, kinetics, and mechanistic insights. Appl. Catal. B Environ. Energy 2025, 361, 124703. [Google Scholar] [CrossRef]
  114. Liu, W.; Sun, M.; Ding, Z.; Gao, B.; Ding, W. Ti3C2 MXene embellished g-C3N4 nanosheets for improving photocatalytic redox capacity. J. Alloys Compd. 2021, 877, 160223. [Google Scholar] [CrossRef]
  115. Wang, A.; Wen, Y.; Zhu, H.; Liu, Z.; Wang, H.; Yao, W.; Lin, H. Oxygen vacancy mediated photocatalysis of Ti3C2 MXene quantum dots/W18O49 hybrid membrane for peroxymonosulfate-enhanced oxidation degradation. Chem. Eng. J. 2024, 496, 154264. [Google Scholar] [CrossRef]
  116. Ling, Z.; Lu, T.; Zhenxi, Y.; Xu, B.; Chen, W.; Tang, Y.; Li, L.; Wang, J. Engineering of Bi2O2CO3/Ti3C2Tx heterojunctions co-embedded with surface and interface oxygen vacancies for boosted photocatalytic degradation of levofloxacin. Chem. Eng. J. 2023, 452, 139327. [Google Scholar]
  117. Ma, Y.; Xu, D.; Chen, W.; Tang, Y.; Wang, X.; Li, L.; Wang, J. Oxygen-vacancy-embedded 2D/2D NiFe-LDH/MXene Schottky heterojunction for boosted photodegradation of norfloxacin. Appl. Surf. Sci. 2022, 572, 151432. [Google Scholar] [CrossRef]
  118. Yu, M.; Liang, H.; Zhan, R.; Xu, L.; Niu, J. Sm-doped g-C3N4/Ti3C2 MXene heterojunction for visible-light photocatalytic degradation of ciprofloxacin. Chin. Chem. Lett. 2021, 32, 2155–2158. [Google Scholar] [CrossRef]
  119. Fatima, S.; Anwar, M.; Almalki, A.S.A.; Alhadhrami, A.; Warsi, M.F.; El-Bahy, Z.M. Fabrication of rare earth (Tb+3) and alkaline earth metal (Mg+2) Co-doped CdAl2O4@MXene composite: A unique approach to tune bandgap energy through quantum confinement effect for photocatalytic applications. Ceram. Int. 2024, 50, 29201–29212. [Google Scholar] [CrossRef]
  120. Abbas, H.A.; Abbas, K.K.; Al-Ghaban, A.M.A. Magnetic MXene/g-C3N4 nano catalyst for photocatalytic degradation of clindamycin contaminate in wastewater. Results Chem. 2025, 13, 101934. [Google Scholar] [CrossRef]
  121. Yan, C.; Zongxue, Y.; Guangcheng, Y.; Tan, Q.; He, N.; Guo, S.; Xia, S.; Chen, Z. Interlayered double-doped N,P-MXene/ZnIn2S4 Schottky junction composite photocatalyst: Efficient removal of ciprofloxacin and methyl orange from complex wastewater. J. Alloys Compd. 2024, 1003, 175608. [Google Scholar]
  122. Li, R.; Chen, A.; Deng, Q.; Zhong, Y.; Kong, L.; Yang, R. Well-designed MXene-derived Carbon-doped TiO2 coupled porous g-C3N4 to enhance the degradation of ciprofloxacin hydrochloride under visible light irradiation. Sep. Purif. Technol. 2022, 295, 121254. [Google Scholar] [CrossRef]
  123. Hussain, M.; Mahmoud, M.H.H.; Rasheed, A.; El Azab, I.H.; Anwar, M.; El-Bahy, Z.M. Silver-doped cadmium aluminate and its MXene based composite for visible-light driven photocatalytic degradation of organic pollutants. Opt. Mater. 2024, 155, 115824. [Google Scholar] [CrossRef]
  124. Valli, K.P.; Kala, S.M.J.; Selvam, V.; Anitha, C.; Malathi, B.; Prakash, K.S.; Karutha Pandian, S. Novel hierarchical nanocomposites of g-C3N4/MXene-Sm2O3 for enhanced cefixime degradation under visible light. J. Phys. Chem. Solids 2024, 190, 112011. [Google Scholar] [CrossRef]
  125. Li, D.; Wang, G.; Ye, Y.; Boutinaud, P.; Zheng, X.; Xu, J.; Kang, F. Recent advances in the synthesis, photo-/electrocatalytic properties and applications of MXenes/bismuth-related composites. Chem. Eng. J. 2024, 498, 155098. [Google Scholar] [CrossRef]
  126. Liu, Z.; Yu, X.; Wang, K.; Wei, Y.; Zhang, J.; Niu, J. Exploiting Cu2O-MXene(Ti3C2) photocatalysts with visible light drive for the degradation of sulfamethazine in the environment. Opt. Mater. 2025, 159, 116650. [Google Scholar] [CrossRef]
  127. Dan, L.; Jiajie, X.; Changkun, L. Constructing MXene-derived Z-Scheme g-C3N4/Ti3C2TX/Ag3PO4 photocatalysts with enhanced charge transfer for aquatic organic pollutants removal. Colloids Surf. A Physicochem. Eng. Asp. 2023, 675, 132050. [Google Scholar]
  128. Zhang, S.; Wang, Y.; Mersal, G.A.; Alhadhrami, A.; Alshammari, D.A.; Wang, Y.; Algadi, H.; Song, H. Enhanced photocatalytic CO2 reduction via MXene synergism: Constructing an efficient heterojunction structure of g-C3N4/Nb2C/CsPbBr3. Adv. Compos. Hybrid Mater. 2024, 7, 226. [Google Scholar] [CrossRef]
  129. Xu, R.; Wei, G.; Xie, Z.; Diao, S.; Wen, J.; Tang, T.; Hu, G. V2C MXene–modified g-C3N4 for enhanced visible-light photocatalytic activity. J. Alloys Compd. 2024, 970, 172656. [Google Scholar] [CrossRef]
  130. Subha, N.; Nagappagari, L.R.; Sankar, A.R. A review on recent advances in g-C3N4-MXene nanocomposites for photocatalytic applications. Nanotechnology 2024, 35, 502002. [Google Scholar] [CrossRef] [PubMed]
  131. Mousavi, S.M.; Mohtaram, M.S.; Rasouli, K.; Mohtaram, S.; Rajabi, H.; Sabbaghi, S. Efficient visible-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]
  132. Yang, J.; Yin, H.; Du, A.; Tebyetekerwa, M.; Bie, C.; Wang, Z.; Zhang, X. Unveiling O2 adsorption on non-metallic active site for selective photocatalytic H2O2 production. Appl. Catal. B Environ. Energy 2025, 361, 124586. [Google Scholar] [CrossRef]
  133. Jie, W.; Bingyan, Y.; Keru, H.; Ji, Y.; Li, X.; Li, C.; Yao, C.; Cai, Z. pH-responsive fiber membrane with high flux and photocatalytic performance for oily wastewater. J. Environ. Chem. Eng. 2025, 13, 115322. [Google Scholar]
  134. Zhenhai, W.A.N.G.; Zikai, Z.H.O.U.; Sen, W.A.N.G.; Zhi, F.A.N.G. Enhanced degradation of tetracycline by gas-liquid discharge plasma coupled with g-C3N4/TiO2. Plasma Sci. Technol. 2024, 26, 094007. [Google Scholar] [CrossRef]
  135. Mirzaei, R.; Yunesian, M.; Nasseri, S.; Gholami, M.; Jalilzadeh, E.; Shoeibi, S.; Bidshahi, H.S.; Mesdaghinia, A.A.-O. An optimized SPE-LC-MS/MS method for antibiotics residue analysis in ground, surface and treated water samples by response surface methodology-central composite design. J. Environ. Health Sci. Eng. 2017, 15, 21. [Google Scholar] [CrossRef]
  136. Rusu, A.A.-O.; Buta, E.A.-O. The Development of Third-Generation Tetracycline Antibiotics and New Perspectives. Pharmaceutics 2021, 13, 2085. [Google Scholar] [CrossRef]
  137. Eslaminejad, S.; Rahimi, R.; Fayazi, M. Green decoration of Pd nanoparticles on MXene/metal organic framework support for photocatalytic degradation of ofloxacin. J. Ind. Eng. Chem. 2025, 141, 94–103. [Google Scholar] [CrossRef]
  138. Zhang, H.; Bao, L.; Pan, Y.; Du, J.; Wang, W. Interface reconstruction of MXene-Ti3C2 doped CeO2 nanorods for remarked photocatalytic ammonia synthesis. J. Colloid Interface Sci. 2024, 675, 130–138. [Google Scholar] [CrossRef] [PubMed]
  139. Wei, C.; Junli, N.; Ye, G.; Gao, C.; Wang, M.; Wang, W.; Lu, X.; Ma, X.; Zhong, P. A review of how to improve Ti3C2Tx MXene stability. Chem. Eng. J. 2024, 496, 154097. [Google Scholar]
Figure 1. MXene composite type [25,26,27,28,29,30,31].
Figure 1. MXene composite type [25,26,27,28,29,30,31].
Polymers 17 02630 g001
Figure 2. MXene synthesis method and impact [32,33].
Figure 2. MXene synthesis method and impact [32,33].
Polymers 17 02630 g002
Figure 3. Secondary electron SEM micrographs for (A) Ti3AlC2 particles before treatment, which is typical of unreacted MAX phases, (B) Ti3AlC2 after HF treatment [18], (C) Multilayer MXene particle, (D) Cross-section of rolled Ti3C2 film [36].
Figure 3. Secondary electron SEM micrographs for (A) Ti3AlC2 particles before treatment, which is typical of unreacted MAX phases, (B) Ti3AlC2 after HF treatment [18], (C) Multilayer MXene particle, (D) Cross-section of rolled Ti3C2 film [36].
Polymers 17 02630 g003
Figure 4. TEM images of (a) Ti3C2TX sheets and (c) Ti3C2TX-nanosheets/Cu2O composite; and HRTEM images of (b) Ti3C2TX sheets and (d,e) Ti3C2TX-nanosheets/Cu2O composite [14].
Figure 4. TEM images of (a) Ti3C2TX sheets and (c) Ti3C2TX-nanosheets/Cu2O composite; and HRTEM images of (b) Ti3C2TX sheets and (d,e) Ti3C2TX-nanosheets/Cu2O composite [14].
Polymers 17 02630 g004
Figure 5. (a) Schematic illustration for electronic structure of Ti3C2 and (101) facets of anatase TiO2; (b) density of state (DOS) plots for Ti3C2 where C 2p, Ti 3d and TDOS are partial DOS of C 2p, partial DOS of Ti 3d, and total DOS, respectively; (c) calculated electrostatic potential of Ti3C2; (d) Structure model of PGCN, Ti3C2 QD, and Ti3C2 QD/PGCN after structure optimization; Theoretical calculations of DOS of (e) Ti3C2 QD, (f) PGCN, and (g) Ti3C2 QD/PGCN; Work function of (h) PGCN, (i) Ti3C2 QD [41,68].
Figure 5. (a) Schematic illustration for electronic structure of Ti3C2 and (101) facets of anatase TiO2; (b) density of state (DOS) plots for Ti3C2 where C 2p, Ti 3d and TDOS are partial DOS of C 2p, partial DOS of Ti 3d, and total DOS, respectively; (c) calculated electrostatic potential of Ti3C2; (d) Structure model of PGCN, Ti3C2 QD, and Ti3C2 QD/PGCN after structure optimization; Theoretical calculations of DOS of (e) Ti3C2 QD, (f) PGCN, and (g) Ti3C2 QD/PGCN; Work function of (h) PGCN, (i) Ti3C2 QD [41,68].
Polymers 17 02630 g005
Figure 6. (a) The Mechanism of Catalytic Synergistic Effect Assisted by MXene Heterojunction; (b) Mechanism diagram of the catalysis-assisting effect of MXene Schottky Heterojunction; (c) Mechanism Diagram of the Role of Z-Scheme Heterojunction Mediated by MXene Intermediate; (d) The sponge-like properties of MXene as an electron acceptor and a hole acceptor.
Figure 6. (a) The Mechanism of Catalytic Synergistic Effect Assisted by MXene Heterojunction; (b) Mechanism diagram of the catalysis-assisting effect of MXene Schottky Heterojunction; (c) Mechanism Diagram of the Role of Z-Scheme Heterojunction Mediated by MXene Intermediate; (d) The sponge-like properties of MXene as an electron acceptor and a hole acceptor.
Polymers 17 02630 g006
Table 1. Comparison of the degradation performance of different photocatalytic composite materials on tetracycline.
Table 1. Comparison of the degradation performance of different photocatalytic composite materials on tetracycline.
PhotocatalystsTetracycline (mg L−1)Reaction TimeThe Dosage of Photocatalyst (mg)Removal Rate (%)Rate Constant k (min−1)Quantum Efficiency (%)Reference
D–OM–ZIF-8/ZnO1000Irradiate with ultraviolet-visible light for 60 min.5088.5
90.5
0.0482.9[9]
CF/ZnO/Ag2O20Irradiate with a 500 W xenon lamp for 30 min.5094.50.083683.8[10]
BiOI/Brookite TiO220Irradiate with visible light for 110 min.3082.00.0634.8[11]
Ag2CO3/ZIF-8/CF10Irradiate with visible light for 30 min.3092.00.0765.1[12]
CeO2/BiYO330Irradiate with visible light for 60 min.5090.00.0855.5[13]
Ti3C2TX/Cu2O30Irradiate with visible light for 40 min.3097.60.0916.0[14]
PVP-MXene-PET0.1visible light for 50183.01%--[15]
MXene-PVA-TiO250 mL, 2 mg/L
(MB)
a 300 W Xe lamp for 9 h-95.2%--[16]
PVP/PEO/MXene Nanocomposite2 ppm (MB) 500–800 nm for 120 min5 wt% of MXene61.6%0.08-[17]
Table 2. MXene features optimized photocatalysis for other semiconductors [41,60,61].
Table 2. MXene features optimized photocatalysis for other semiconductors [41,60,61].
AdvantageThe Deficiencies Existing in Other Semiconductor PhotocatalysisThe Mechanism of the Enhanced Photocatalytic Effect of MXene
Abundant catalytic active sitesThe catalytic sites are occupied due to the influence of environmental factors.Construct composites by loading single atoms, clusters and nanoparticles on the surface to provide multiple sites.
Abundant surface functional groupsPoor stabilityBond with the main catalyst in the form of hydrogen bonding chemical bonds to improve stability.
Transition metals possess relatively high redox ability.The phenomenon of photocorrosionThe high redox activity of transition metals inhibits the photocorrosion phenomenon.
Broad spectral response rangeA relatively low light response rangeComposites formed by mechanical mixing, self-assembly and in situ oxidation have a relatively high light response range
Lower Fermi level compared with common semiconductorsHigh recombination rate of photogenerated electrons and holesConstruct structures such as Schottky heterojunctions to inhibit the recombination of photogenerated electrons and holes.
High electrical conductivityThe transportation efficiency of photogenerated electrons is not high.MXene acts as an electron acceptor to improve electron transfer efficiency.
Table 3. Band structure matching requirements [60,87,88].
Table 3. Band structure matching requirements [60,87,88].
ParameterSemiconductor A (Reduced Prototype)Semiconductor B (Oxidized Type)MXene Roles
Band Position (ECB)High (e.g., g-C3N4)Low (e.g., TiO2)Electron transfer bridges
Valence Band Position (EVB)Lower (e.g., g-C3N4)High (e.g., TiO2)Hole transport channels
Differences in work function (ΔΦ)≥0.5 eV Drive charge separation
Table 4. MXene co-catalytic optimization function [23,41,93].
Table 4. MXene co-catalytic optimization function [23,41,93].
FunctionMXene
Photocatalyst preparationGrowth platformSurface functional groups such as -OH, -O, -F, etc.
Semiconductor precursorsMetastable transition metal atoms
Improved photocatalytic activityElectronic receiversHigh electrical conductivity and good band structure
Active siteMulti-purpose transition metal atoms
adsorbentElectrostatic attraction
Enhanced photostabilityAvoid photoelectron reductionTransfer of photogenerated electrons
Table 5. Spatiotemporal properties of interfacial energy transfer.
Table 5. Spatiotemporal properties of interfacial energy transfer.
ProcessTime ScaleDominant MechanismEffect on ReactionReference
Thermal electron generation<100 fsElectron excitationCharge transfer initiates redox[105]
Electron–phonon relaxation1–10 psEnergy localizationThe reaction temperature field is regulated[106]
Interface vibration coupling10–100 psMolecular vibrational mode excitationReduces the activation energy of the reaction[107]
Table 6. Comparison of performance with traditional materials.
Table 6. Comparison of performance with traditional materials.
ParameterMXenePrecious Metal NanoparticlesSemiconductor NanosheetsReference
Light absorption rangeVisible infraredVisible lightThe bandgap is decided[74]
Carrier mobility103–104 cm2/VsLimited by size effectUsually <100 cm2/Vs[108]
Surface reactivityCan be chemically modifiedIt depends on crystal plane exposureGoverned by defective states[41]
Photothermal conversion efficiency92% (808 nm)65–75% (Au)Affected by light absorption characteristics[95,109]
Hot carrier lifetime150–300 fs10–50 fsInteraction of interfaces[110]
Table 7. Synergistic Degradation of Antibiotics by MXene Co-catalytic Materials Modified with Oxygen Vacancies and Elemental Doping.
Table 7. Synergistic Degradation of Antibiotics by MXene Co-catalytic Materials Modified with Oxygen Vacancies and Elemental Doping.
CategoriesPhotocatalytic MaterialsSynthesis MethodContaminantReaction ConditionsDegradation EfficiencyReference
Introduction of oxygen vacanciesNiFe-LDH/MXeneHydrothermal methodNorfloxacin (20 mg/L)300 W Xenon lamps
25 °C ± 1 °C
4 h, 98%[117]
BiOBr/MXene/gC3N4Electrostatic self-assemblyTetracycline (20 mg/L)300 W Xenon lamps
25 °C ± 1 °C
30 min, 99%[104]
Bi2O2CO3/Ti3C2Tx Hydrothermal methodLevofloxacin (20 mg/L)
Amoxicillin, tetracycline (20 mg/L)
300 W Xenon lamps
25 °C ± 1 °C
80 min, 95.4%
30 min, 90.9% and 82.8%
[116]
Doping of metallic elementsSm doped
g-C3N4/Ti3C2MXene
AnnealingCiprofloxacin (20 mg/L)300 W Xenon lamps
25 °C ± 1 °C
60 min, 99%[118]
Tb3+ and Mg2+ doped CdAl2O4@MXene Co-precipitation
Sonication
Aspirin (20 mg/L)300 W Xenon lamps
25 °C ± 1 °C
135 min, 79.6%[119]
Co doped ZnTiO3/Ti3C2Tx MXene Liquid self-assemblyTetracycline
(20 mg/L)
300 W Xenon lamps
25 °C ± 1 °C
90 min, 91.5%[49]
Fe doped with magnetism MXene/g-C3N4 Hydrothermal synthesisClindamycin (initial drug concentration 125 mg/L) mix with wastewater100 mW/cm2 of high-voltage lamps
25 °C
120 min, 92%[120]
Doping of non-metallic elementsN,P-MXene/ZnIn2S4 Schottky knotIn situ grownCiprofloxacin (20 mg/L)300 W Xenon lamps
25 °C ± 1 °C
120 min 99.3%[121]
Homojunctions of C-doped amorphous TiOx derived from TiO2 and Ti3C2TxHydrogen peroxide oxidationTetracycline (30 mg/L)300 W Xenon lamps
25 °C ± 1 °C
100 min, 91.5%[92]
MXene-derived carbon-doped TiO2 coupled with porous g-C3N4One-step hot calcination methodCiprofloxacin hydrochloride (20 mg/L)300 W Xenon lamp 420 nm UV filter
25 °C ± 1 °C
50 min, 88.14%[122]
Table 8. Antibiotic classification, structure and pKa value [135,136].
Table 8. Antibiotic classification, structure and pKa value [135,136].
Major Classes of AntibioticsSecondary ClassificationCategorypKaStructure
β-lactam antibioticsPenicillinsPenicillin G2.8carboxyl
Amoxicillin2.0/7.3Carboxy/amino
Piperacillin2.74/5.13Carboxyl/pyridine ring
CephalosporinsCefpelin2.74/5.13Carboxyl/pyridine ring
Ceftiofur2.68carboxyl
Aminoglycoside antibioticsAminoglycosidesgentamicin7.5–9amino
Tobramycin7.5–9amino
Ampramycin7.5–9amino
Tetracycline antibioticsTetracyclines tetracycline2.8–3.4/7.2–7.8/9.1–9.7Phenolic hydroxyl/enol hydroxyl/dimethylamino
oxytetracycline2.8–3.4/7.2–7.8/9.1–9.7Phenolic hydroxyl/enol hydroxyl/dimethylamino
aureomycin2.8–3.4/7.2–7.8/9.1–9.7Phenolic hydroxyl/enol hydroxyl/dimethylamino
Fluoroquinolone antibioticsFluoroquinolonesCiprofloxacin3–4/6/7.5–9/10–11Carboxyl/enol hydroxyl/amino/other groups
Levofloxacin3–4/6/7.5–9/10–11Carboxyl/enol hydroxyl/amino/other groups
Moxifloxacin3–4/6/7.5–9/10–11Carboxyl/enol hydroxyl/amino/other groups
Sulfonamide antibioticsSulfonamidesSulfamethoxazole2/5–7.5Sulfonamide/other groups
Sulfadiazine2/5–7.5Sulfonamide/other groups
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

Chen, Z.; Meng, Z.; Zhang, Z.; Ma, W. MXene-Polymer Nanocomposites for High-Efficiency Photocatalytic Antibiotic Degradation Review: Microstructure Control, Environmental Adaptability and Future Prospects. Polymers 2025, 17, 2630. https://doi.org/10.3390/polym17192630

AMA Style

Chen Z, Meng Z, Zhang Z, Ma W. MXene-Polymer Nanocomposites for High-Efficiency Photocatalytic Antibiotic Degradation Review: Microstructure Control, Environmental Adaptability and Future Prospects. Polymers. 2025; 17(19):2630. https://doi.org/10.3390/polym17192630

Chicago/Turabian Style

Chen, Zhenfei, Zhifei Meng, Zhongguo Zhang, and Weifang Ma. 2025. "MXene-Polymer Nanocomposites for High-Efficiency Photocatalytic Antibiotic Degradation Review: Microstructure Control, Environmental Adaptability and Future Prospects" Polymers 17, no. 19: 2630. https://doi.org/10.3390/polym17192630

APA Style

Chen, Z., Meng, Z., Zhang, Z., & Ma, W. (2025). MXene-Polymer Nanocomposites for High-Efficiency Photocatalytic Antibiotic Degradation Review: Microstructure Control, Environmental Adaptability and Future Prospects. Polymers, 17(19), 2630. https://doi.org/10.3390/polym17192630

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop