Next Article in Journal
Fast-Disintegrating Oral Films Containing Nisin-Loaded Niosomes
Previous Article in Journal
UV-Induced Photodegradation of 2-Aminothiazole-4-Carboxylic Acid: Identification of New Carbodiimide Molecules
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Advances in MXene-Based Composites for Their Efficiency in the Degradation of Antibiotics and Water Splitting

1
State Key Laboratory of Environment-Friendly Energy Materials, Southwest University of Science and Technology, Mianyang 621010, China
2
School of Manufacturing Science and Engineering, Key Laboratory of Testing Technology for Manufacturing Process, Ministry of Education, Southwest University of Science and Technology, Mianyang 621010, China
3
Center for Integrative Petroleum Research, King Fahd University of Petroleum and Minerals, Dhahran 31261, Saudi Arabia
4
Information and Computer Science Department, King Fahd University of Petroleum and Minerals, Al Khobar 31261, Saudi Arabia
5
Advanced Functional Materials & Optoelectronic Laboratory (AFMOL), Department of Physics, Faculty of Science, King Khalid University, Abha 61421, Saudi Arabia
*
Author to whom correspondence should be addressed.
Molecules 2025, 30(18), 3712; https://doi.org/10.3390/molecules30183712
Submission received: 21 August 2025 / Revised: 5 September 2025 / Accepted: 8 September 2025 / Published: 12 September 2025
(This article belongs to the Section Photochemistry)

Abstract

The increasing occurrence of antibiotics in water bodies all over the world has raised concerns because of the prospect that they might have genotoxic and antibiotic-resistant consequences in both people and aquatic creatures. In particular, it has been discovered that the construction of hybrid photocatalytic composite materials has greater antibiotic degradation efficiencies. The hybrid photocatalysts deliver improved photoabsorbance, charge separation, transfer, and redox characteristics, as well as enhanced photostability and rapid recovery, due to their optimal characteristic qualities, including superior structural, surface, and interfacial properties. Additionally, metal-based electrocatalysts have garnered notable attention in the field of water splitting as they are low-cost, standard and have the potential to be used in green and clean technology. MXene, a family of two-dimensional transition metal carbides and nitrides, was discovered in 2011 due to its high conductivity, large surface area, and abundance of catalytically active sites. By making hybrid structures of MXene with other materials, which have shown better electrocatalytic activity than pure MXenes. The two half-cell processes involved in water electrolysis are the oxygen generation at the anode site and the hydrogen production at the cathode site. This review paper provides a summary of the latest advancements in the design of several hybrid systems, catalysts and their effectiveness in degrading a range of newly discovered antibiotic pharmaceutical pollutants in aquatic settings, as well as recent developments on the use of MXenes and MXene-based hybrid structures such as OER, HER, and bifunctional electrocatalysts for general water splitting.

Graphical Abstract

1. Introduction

Growing crises that threaten the balance of the environment, human well-being, and economic resilience highlight the worldwide environmental and energy challenges of 2025. With 2023 officially being the warmest year ever measured and global CO2 concentrations surpassing 420 parts per million, climate change continues to pose the greatest threat, pushing global temperatures up 1.6 degrees Celsius above pre-industrial levels [1]. Storms, droughts, and wildfires are examples of extreme weather events intensified by global warming, which is expected to have resulted in $300 billion in damages in 2024 [2]. With estimates indicating that 5–10% of major urban centers may be flooded by 2030, coastal areas such as Miami and Shanghai are at risk due to rising sea levels caused by glacier melt and thermal expansion [3]. Despite a 3% increase in coal use in emerging countries due to energy security concerns, natural gas, which accounts for 75% of global greenhouse gas emissions [4], continues to dominate energy systems. These issues are exacerbated by pollution, as only 9% of plastics are successfully recycled each year, despite the annual output reaching over 380 million tons [5].
Pharmaceutical-activated chemicals are the most significant class of emerging aquatic pollutants, and their emergence has garnered considerable professional and public awareness in recent times. Due to their persistence and adverse effects on marine life, pharmaceutical chemicals are continually escaping into the environment, causing negative impacts [6,7,8]. Pharmaceutical substances that are biologically effective in the human system and have the prospect to exert significant effects on the neurological, immunological, circulatory, and hormonal systems are particularly concerning pollutants [9,10]. River water, irrigation water, and household wastewater all contain the medication carbamazepine (CBZ), which is even impermeable to microbial biodegradation [11]. As a psychotropic and antiepileptic drug, carbamazepine is an extensively used pharmaceutical that poses a risk to the atmosphere and biodegrades slowly in water [12]. To eliminate these antibiotics from the aquatic environment, photocatalysis has become a focus of study, as it is a clean, renewable technology with significant potential for tackling economic and environmental issues [13,14,15,16]. Photo-elimination may utilize sunlight to oxidize or degrade pollutants and trigger electron excitation transfer in a semiconductor photocatalyst [17,18,19,20]. Numerous semiconductors, sulfides, and metal oxides have been used to generate nanostructured compounds that serve as photocatalytic examples when exposed to sunlight [21]. However, single-material photocatalysts are unsuccessful due to the rapid recombination of photo-induced charge transfer, which results in low catalyst efficiency. Thus, a range of methods, including metal or non-metal doping, facet regulation, surface sensitization, and heterojunction creation, have been employed to enhance photocatalytic activity and reduce recombination [21,22,23,24,25].
The hydrogen energy is the most faultless clean energy currently available. It is widely used in the transportation, chemical, and aerospace sectors due to its high calorific value and low weight [26,27,28,29,30]. Water resources comprise a lot of hydrogen. Thus, it is almost essential to investigate technologies for the operational mass synthesis of hydrogen from water [31,32,33,34,35]. Water separates into hydrogen and oxygen when electricity is switched on. At the cathode and anode, respectively, the hydrogen generation reaction and the oxygen production reaction occur. Two electrons are transported during the heterogeneous catalytic procedure known as the hydrogen evolution reaction (HER). Regardless of the reaction medium, HER’s theoretical decomposition potential remains constant. However, distinct HER routes are determined by the reaction medium’s pH value, which has a significant influence on the reaction rate. For the HER, Pt is the perfect catalyst. However, it is not feasible to use on an extensive scale due to Pt’s high cost and incomplete deposits [36,37]. As a result, just 4% of industrial hydrogen generation is now formed using electrolytic water technology. As a result, research on HER catalysts is focused on either creating low-cost and effective non-noble metal catalysts or reducing the dose of noble metals.
When photons are absorbed, electron-hole pairs are created within the material’s bandgap, which starts the process of photocatalytic water splitting. According to the foundational study, these charge carriers then go to the catalyst surface, where they can initiate redox reactions with water, as long as the potential difference is greater than the 1.23 eV threshold. The first known example of water splitting into hydrogen and oxygen was recorded in 1972 [38], showing that the reaction could be triggered by visible light alone without the requirement for an external voltage. In 1980, the use of powdered TiO2 and SrTiO3 photocatalysts for UV-induced water splitting was first reported [39]. Since then, numerous studies have been conducted on the use of TiO2 as a photocatalyst in environmental remediation and water treatment [40,41]. TiO2 has been the subject of the majority of photocatalysis research, with the most significant amount of attention being paid to this material to the extent that TiO2 is frequently regarded as a prototype for transition metal oxide surfaces [42,43]. An organic framework that can perform photocatalysis without the aid of co-catalysts was recently reported by Du et al. [44]. When exposed to visible light, the framework demonstrated hydrogen evolution rates of 15,480 μmol/g.h from saltwater and 22,450 μmol/g.h from pure water. One approach that is seen as promising for meeting the industrial demands of hydrogen production on a larger scale is photocatalytic water splitting. In addition to facilitating effective chemical processes at room temperature when exposed to sunlight, photocatalysis has garnered significant interest due to its potential to create ideal technologies that transform abundant, harmless, and clean solar energy into chemical and electrical energy [45].
The oxygen evolution reaction (OER) occurs concurrently at the anode. Compared to the HER, which requires a larger overpotential to drive, the OER exhibits slower dynamics because it is a heterogeneous catalytic procedure that involves the transfer of four electrons. During water electrolysis, the OER is the phase that uses the most energy. The primary purpose of the overpotential needed for water electrolysis is to overcome the OER’s reaction energy barrier [37,46]. Precious metals like Ru and Ir, however, are expensive and challenging to use extensively. Thus, the formation of affordable and effective substitute catalysts is essential to OER catalyst research. Two-dimensional (2D) materials have gathered considerable attention in the catalysis community because of their distinct structure and electrical properties.
Over the last decade, 2D MXene materials have been widely utilized in various catalytic processes, including photoelectric, thermal, electrochemical, and photocatalytic applications [47,48,49,50]. MXenes have been extensively used in multiple fields in recent years due to their superior electrical, magnetic, chemical, and mechanical properties [51,52,53,54,55]. MXenes’ remarkable potential, which encompasses their high metallic conductivity, electronic properties, ease of functionality, and applications in sensors, energy storage, and environmental protection, makes them most suitable for a diverse range of applications. High surface area, biocompatibility, hydrophilicity, activated metallic hydroxide sites, zation, and tunable bandgap [52,56,57].
This review discusses the use of MXene composites, highlighting their flow, anti-fouling, and antibacterial properties in functional devices, such as catalyst membranes for continuous water treatment. The characterization methods (XRD, SEM, EDS) used to link structural characteristics to catalytic performance are covered. The paper concludes by outlining current constraints, including oxidation susceptibility, scalability issues, and restacking of MXene layers. It suggests future research avenues to overcome these obstacles and develop practical applications in green hydrogen production and wastewater treatment.

2. Mxene and Its Unique Characteristics

MXenes are a type of 2D material, mainly containing transition metal carbides and nitrides. They are usually obtained from layered ternary compounds called MAX phases, having the general formula Mn+1AXn, Here M stands for early transitional metal (for example, Nb, Ti, Cr, V, Mo, etc.), A denotes the group 13 or 14 elements (such as Si, Al, Ga, etc.), while X is carbon and nitrogen [58]. The synthesis of MXenes requires selectively etching out layers of elements from the MAX phases, typically using fluorine-containing etchants, which yields ultra-thin layered nanosheets of the form Mn+1XnTx, where Tx denotes surface terminations of hydroxyl (-OH) or oxygen (=O) groups, or fluorine (-F) groups [59]. These surface terminations are crucial in tuning the properties of MXenes, both chemically and physically, and thus, MXenes become highly adaptable for various applications. All the unique properties are shown in Figure 1.
Figure 1. Unique characteristics of MXene materials. Reproduced with permission [60].
Figure 1. Unique characteristics of MXene materials. Reproduced with permission [60].
Molecules 30 03712 g001
One of the defining characteristics of MXenes is their outstanding electrical conductivity, which is comparable to metals. For example, some MXenes have conductivities as high as approximately 11,668 S/cm, making them the best candidates to use in energy conversion and storage, sensing, and electronics [61]. The reason they have such high conductivity is that metal layers are transferred to a 2D geometry, helping to transport electrons effectively. In addition, due to the structure of the MXene layers, they possess a significant specific surface area, providing multiple operating positions for chemical and adsorbent reactions [62]. This feature is especially beneficial in aspects of catalysis, supercapacitors, and batteries, where surface interactions dictate performance.
In addition to electrical properties, MXenes have significant tolerance and flexibility. The 2D panels of these materials are powerful yet flexible, allowing for their combination in composite materials or production into films without structural collapse. Another important property is thermal stability; MXenes retain their structure and function at high temperatures, thereby expanding their applications in severe operating conditions [63]. Additionally, MXenes have hydrophilic surfaces due to their functional groups, thereby increasing their dispersibility in water and their ability to interact with biological molecules, which is beneficial for biomedical applications. MXenes are a highly chemically adjustable material. By modifying the type of transition metal composition, carbon-to-nitrogen ratio, and surface termination, scientists can alter the structure of MXE and its electronic, optical, and catalytic properties. For instance, different metals or heteroatoms can be used to control conductivity or catalytic activity, while maintaining surface terminations alters hydrophilicity and chemical reactivity [64]. Such tunability enables the use of MXenes in a variety of applications, including energy storage systems (such as batteries and supercapacitors), environmental remediation, and biomedical technologies (such as drug delivery and antibacterial agents).
The SEM images of pure MXene and MX-TiO2 composites, prepared with and without NaCl at various hydrothermal temperatures. As shown in Figure 2, Pristine MXene exhibits a characteristic accordion-like layer structure. After hydrothermal treatment at 175 °C and 200 °C, several nanoparticles were detected to develop on the surface and between the layers of the MXene structure, as shown in Figure 2b,c [65]. The various synthesis techniques are shown in Figure 3.The morphological characteristics of Ti3AlC2, Ti3C2Tx, and 001-T/MX were scrutinized using field-emission scanning electron microscopy. A compact, layered ternary carbide structure was perceived in Ti3AlC2. Following HF treatment, the compact symmetry of Ti3AlC2 transformed into a 2D structure, as shown in Figure 4d,e. When HF underwent an exothermic reaction, H2 gas was released, forming the accordion-like structure of MXene as shown in Figure 4b The SEM pictures, also shown in Figure 4c revealed the creation of heterostructures in nanosheets, with a significant percentage of the (001) planes of the anatase phase visible, and the emergence of (001)-TiO2 crystals in sheets. As anticipated, the Ti3C2Tx surface retained its hexagonal shape but became rough during the hydrothermal process. The achievement of the intended materials may be foreseen using the morphological characteristics of several material phases [66].
Figure 2. Scanning electron microscopy of (a) MXene, (b) MX-TiO2 (175 °C), and (c) MX-TiO2 (200 °C). Reproduced with permission [65].
Figure 2. Scanning electron microscopy of (a) MXene, (b) MX-TiO2 (175 °C), and (c) MX-TiO2 (200 °C). Reproduced with permission [65].
Molecules 30 03712 g002
Despite the advantageous properties, functions, and applications of MXenes, several outstanding issues remain regarding their production and stability. Many MXenes’ properties have yet to be explored; some of their characteristics, such as ease of scalability, remain uncertain. Other synthesis techniques also use hazardous reagents, such as hydrofluoric acid, which poses safety and environmental hazards [67]. Furthermore, MXenes are susceptible to oxidation when exposed to air or liquids, which can deteriorate their performance over time. Current developments focus on creating safer and more affordable synthesis pathways, such as fluorine-free techniques and molten salt etching, in addition to surface modifications to enhance stability and prevent layer self-restacking, which can impede ion transit.

2.1. Structural Components

2.1.1. Transition Metal Layers

The tightly packed layers of atoms make up the backbone of MXenes. Electronic and mechanical properties are determined by these transition metals (Groups 3–6). Titanium (Ti) is used in Ti3C2Tx, the first MXene discovered, due to its excellent mechanical toughness and electrical conductivity (~11,000 S/cm) [68]. By introducing solid compounds or ordered double-metal structures, such as multi-metal MXenes like (Ti,V)2CTx or (Mo2Sc)C3Tx, customized features can be achieved, including improved catalytic efficiency and magnetic behavior.

2.1.2. Carbon/Nitrogen Layers

Strong covalent and metallic connections are formed by the X atoms’ occupation of octahedral interstitial spaces between M layers. Recent developments include oxygen-containing MXenes (such as Ti2COTx), which replace some X sites to form oxy-carbides with adjustable band gaps. Still, these are generally C or N. Stability and reactivity are affected by the X composition; nitrides, such as Ti4N3Tx, frequently show greater resistance to oxidation than carbides [64].

2.1.3. Layer Thickness

The number of M layers stacked between X layers depends on the value of n. For instance, M2XTx (such as Ti2CTx) is a single M layer that provides a high degree of flexibility. n = 3: Thicker sheets with increased mechanical strength (M4X3Tx, for example, Ta4C3Tx) [69]. Although increasing n tends to increase structural stability, it may decrease reaction surface accessibility.

2.1.4. Surface Terminations

During synthesis, Tx groups are introduced by etching agents (such as HF and HCl/LiF). These terminations hinder restacking and modifying characteristics by capping exposed M atoms: O-O/-OH: Increase catalytic activity and hydrophilicity. O-F: May decrease conductivity but increase electrochemical stability. Tx may be precisely controlled by post-synthesis procedures (such as thermal annealing or chemical modification), which allows for property optimization for uses such as water splitting or antibiotic degradation [70].

2.1.5. Layered Architecture

MXenes inherit a hexagonal close-packed structure based on MAX phases. Etching removes the parent MAX phase, yielding 2D sheets with an “accordion-like” morphology. Every layer is made up of n + 1 M layers intercalated between n X layers, capped with Tx groups on both termini. This structure offers a high surface-to-volume ratio (~300 m2/g) and numerous active sites, which are essential for adsorption and catalysis [71].

2.2. Synthetic Methods

Top-down etching techniques and bottom-up fabrication procedures are the two primary methods used to synthesize MXenes. The most well-known and popular technique is the top-down technique, which produces the 2D sheets of MXene (Mn+1XnTx) by selectively eliminating the layer of “A” element from the original MAX phases (Mn+1AXn) [72]. Usually, chemical etching procedures employing hydrofluoric acid (HF) or salts containing fluoride are used to accomplish this. The conventional top-down etching approach dissolves the “A” layers, such as silicon or aluminum, by directly applying strong HF acid to the powdered MAX phase [73]. The “A” atoms are eliminated by the HF etching reaction, which also functionalizes the exposed MXene surfaces with ending groups like -O, -OH, and -F, which affect the MXene’s characteristics. The quality, thickness of the layer, and surface chemical composition of the resultant MXenes are significantly impacted by parameters including HF temperature, concentration, and etching time [74]. However, because it is poisonous and corrosive, the direct use of HF raises serious safety and environmental concerns.
Figure 3. Different synthetic techniques of MXene materials. Reproduced with permission [75].
Figure 3. Different synthetic techniques of MXene materials. Reproduced with permission [75].
Molecules 30 03712 g003
To overcome these limitations, modified acid etching techniques have been created that combine fluoride salts (such as LiF and NH4HF2) with hydrochloric acid (HCl) to produce HF in situ. By producing HF slowly in the reaction compound, this technique, often referred to as the “MILD” approach, reduces risks and improves the regulation of etching kinetics [76]. For instance, adding the MAX phase after breaking down LiF in HCl creates a safer etching condition while still effectively removing the “A” layers. After sonication, the in situ HF technique facilitates the delamination process. It creates few-layer or single-layer nanosheets by facilitating the intercalation of cations (such as Li+) between MXene layers. Systems that have been nitrogen-purged to reduce oxidation during synthesis are variations in this strategy [77].
Molten salt etching, another top-down method gaining popularity, involves heating MAX stages with molten fluoride salts, including NaF, LiF, and KF combinations at high temperatures, of approximately 550 °C [78]. Through chemical processes that result in volatile byproducts, such as AlCl3 gas, this technique quickly eliminates the “A” layers, leaving MXene layers that end with chlorine or other groups. Although the method eliminates liquid HF and permits quicker processing periods (minutes as opposed to hours), molten salt etching usually produces MXenes with lower crystallinity and more surface imperfections [79]. The MXene sheets must be purified and separated using post-synthesis rinsing and delamination procedures. Although it shows promise for large-scale production, this method consumes a significant amount of energy and requires optimization to enhance the quality of MXene.
In contrast to top-down approaches, the bottom-up formation of MXenes involves constructing MXene structures from smaller precursors atom by atom or molecule by molecule, often using solution-based crystal growth, molecular beam epitaxy, or chemical vapor deposition (CVD) [67]. The size, form, structure, and surface terminations of MXene can be exquisitely controlled through these approaches. For electrical and catalytic applications, bottom-up approaches can produce MXene films or heterogeneous structures with fewer defects and customized properties [80]. Unlike decreasing engraving, increasing production is still in the early stages for MXenes, and it faces the difficulties of such complicated precursors, prolonged processing times, and expansion issues.

3. MXene in the Degradation of Antibiotics

Concern about the presence of antibiotics in industrial discharges, drinking water, and wastewater has grown in recent years [81,82]. One of the most significant classes of synthetic antibiotics, fluoroquinolones (FQs), is widely used to treat diseases caused by a variety of species of bacteria [83,84]. Traditional sewage treatment measures cannot efficiently eliminate the majority of FQs due to their stable chemical structures and resistance to biological breakdown [85]. Thus, in aquatic environments, FQs have often been detected at detectable concentrations (for example, 0.6 to 5.6 μg/L) [86,87]. It has been revealed that environmental antibiotic residues can lead to the emergence of new, antibiotic-resistant bacteria, ultimately posing a threat to public health [88,89]. Therefore, new methods for giving FQs that remain in water are needed. FQs may now be eliminated from effluent using advanced oxidation procedures (AOPs) [90,91]. Photocatalytic expertise has garnered considerable attention as an AOP type due to its long-lasting stability and excellent efficiency. When exposed to UV light, a variety of conservative photocatalysts, such as TiO2, have shown the aptitude to break down these FQs by producing reactive oxygen species (ROS) [92,93]. The expansion of environmentally friendly and highly pH-tolerant sunlight/light-driven photocatalysts is necessary because the application of these photocatalysts has been hindered by their low light-harvesting ability, narrow pH range, difficulty in separation, and potential toxicity [94,95]. Both human health and aquatic ecology are seriously threatened by the presence of pharmaceuticals in marine environments [96,97].
Therefore, effective techniques are necessary to remove these harmful substances from water. An established and effective method for breaking down organic molecules is catalytic degradation. Pharmaceutical substances have been broken down using a variety of catalysts [98]. However, materials with improved catalytic activity are also sought after by researchers. In recent years, MXene-containing catalysts have also been shown to be an adequate substitute for conventional catalysts [99,100,101,102,103]. High surface area, a large number of surface terminations, semiconductor behavior with a tunable bandgap, and quick photogenerated electron-hole separation ability are the main characteristics of MXenes as co-catalysts [104]. These special qualities have made MXene-based photocatalysts a viable alternative to other 2D materials, such as graphitic carbon nitride (g-C3N) and transition metal dichalcogenides (TMDs).

3.1. Degradation of Antiepileptic Drug

The creation of a hybrid photocatalyst based on Ti3C2Tx nanosheets was achieved through an initial hydrothermal treatment procedure. The structural characteristics and morphology of the as-prepared photocatalyst, as well as the chemical makeup of MXene and its TiO2 derivatives in Ti3C2Tx, represented as 001-T/MX, were systematically described. Through the Schottky circuit created between the TiO2-MXene surfaces, a controlled oxidation activity formed the heterojunction of the prepared photocatalyst. The pure MXene and the as-fabricated 001-T/MX nanohybrid for a drug of carbamazepine (CBZ) were scrutinized for their capacities in adsorption and photocatalytic degradation. The degradation measurements were significantly regulated in acidic conditions with a pH range of 3.0–5.0, and the measured Kapp value of CBZ under ultraviolet light was about 0.0304 per minute, greater than that under natural sunlight. The CBZ molecule was mixed by %OH and %O2 throughout the photocatalytic degradation process; specific degradation routes were proposed in response. For the CBZ breakdown, the new heterostructure 001-T/MX showed a lack of pertinence. Ti3AlC2, Ti3C2Tx, and 001-T/MX morphological appearances were scrutinized using field-emission scanning electron microscopy. A compact, layered ternary carbide structure was obtained in Ti3AlC2. Following HF treatment, the compressed symmetry of Ti3AlC2 converted into a 2D structure, as shown in Figure 4a,b [105]. When HF underwent an exothermic reaction, H2 gas was released, forming according to the structure of MXene as shown in Figure 4c [106].
Figure 4d also shows the formation of heterostructures in nanosheets, with a noticeable percentage of the anatase phase planes visible, and the emergence of (001)-TiO2 crystals within the nanosheets. As anticipated, the Ti3C2Tx surface retained its hexagonal shape but became irregular during the hydrothermal process. The successful completion of the proposed materials may be predicted using the morphological characteristics of several material phases. Furthermore, the adsorption of CBZ onto the surface of the photocatalyst is considered an essential step before the photocatalytic degradation of CBZ [33]. Thus, within the reactor, in a dark situation, the adsorption capability of CBZ on the Ti3C2Tx and the 001-T/MX photocatalyst was examined. At around 5% and 7%, respectively, the adsorptive removal of pure MXene and 001-T/MX was very low, as shown in Figure 4e. Non-electrostatic interactions between the neutral CBZ and the negatively charged adsorbent may be the cause of the poor adsorption of CBZ. Since CBZ is a neutral molecule with a pKa (logarithmic value of the acidity constant) of 13.9, the pH of the system under study is unlikely to have an impact on the adsorption process [66].
Figure 4. FE-SEM pictures of the (a) Ti3AlC2 MAX phase, (b) Ti3C2Tx, (c) 001-T/MX photocatalyst, (d) after HF-30% etching, and (e) comparison of the CBZ photodegradation ability of 001-T/MX and Ti3C2Tx MXene photocatalysts. Reproduced with permission [66].
Figure 4. FE-SEM pictures of the (a) Ti3AlC2 MAX phase, (b) Ti3C2Tx, (c) 001-T/MX photocatalyst, (d) after HF-30% etching, and (e) comparison of the CBZ photodegradation ability of 001-T/MX and Ti3C2Tx MXene photocatalysts. Reproduced with permission [66].
Molecules 30 03712 g004
Liquid chromatographytandem mass spectrometry (LC-MS/MS) was applied to recognize the primary intermediate products, thereby determining the breakdown routes of CBZ utilizing the 001-T/MX photocatalyst. Acridine, 2-aminobenzoic acid, 2-hydroxybenzoic acid, and formaldehyde acridine were recognized as the peaks. %OH free radicals produced from the photocatalyst’s heterostructure confronted the aromatic ring of CBZ, succeeding in intermediate and H-abstraction.
Additionally, two distinct degradation pathways designated as route 1 and route 2, were selected based on the intermediates found, as shown in Figure 5 [107]. Formaldehyde-acridine, or intermediate D, was produced when acridine experienced another %OH substitution. Additionally, process 1’s additional transformation may provide benzoic acid (122 m/z) and aniline (93 m/z), which are then broken down into CO2 and H2O as a consequence of further oxidation and ring cleavage [108]. Accordingly, the ring breakage mechanisms in way 2 transformed the formaldehyde-acridine (intermediate D) into CO2 and H2O [66].
Figure 5. Potential paths for CBZ photocatalytic degradation by the 001-T/MX photocatalyst. Reproduced with permission [66].
Figure 5. Potential paths for CBZ photocatalytic degradation by the 001-T/MX photocatalyst. Reproduced with permission [66].
Molecules 30 03712 g005

3.2. Degradation of Fluoroquinolone

Serious concerns have been raised by the presence of fluoroquinolones (FQs) in the atmosphere. This work examined the mechanism and photocatalytic degradation kinetics of ciprofloxacin (CIP) in ordered mesoporous g-C3N4 (ompg-C3N4). When exposed to sunshine radiation, ompg-C3N4 reacted to CIP deterioration 2.9 times faster than the bulk form of g-C3N4. This improvement might be clarified by ompg-C3N4′s high surface area and efficient charge carrier separation. The Langmuir-Hinshelwood kinetics model was employed to eliminate CIP, highlighting the significant role of surface reactions in the photocatalytic mechanism. Succeeding research on reactive species (RSs) using both ESR knowledge and RS scavenging studies has exhibited that the breakdown of CIP is instigated mainly by the photo hole (h+) and the consequent peroxide anion radical (O2). The degradation ways of CIP were suggested based on the detection of mediates using liquid chromatography with tandem mass spectrometry and the estimation of reactive sites using Frontier Electron Densities. The degradation of FQs during the photocatalysis procedure was significantly affected by the piperazine mediator, as indicated by a comparison of the FQ degradation rates. The ompg-C3N4 demonstrated a highly desirable performance in decreasing antibiotic activity, as observed in an experiment on residual antibiotic activity. A ompg-C3N4 photocatalytic technique driven by sunlight may be effectively utilized to clean up the CIP-contaminated natural waterways, according to the tolerable photogenerated degradation of CIP in ambient water [109].
The concentration of CIP remained unchanged in the absence of the photocatalyst, as shown in Figure 6a, verifying the photostability of CIP during simulated exposure to sunshine. According to the adsorption experiment, under dark conditions, ompg-C3N4 absorbed 24.1% of CIP. In comparison, after 50 min of simulated sunshine irradiation, bulk C3N4 exhibited a degradation competence of 61.2%. Prominently, ompg-C3N4 significantly enhanced the degradation of CIP. For example, during the same period, 92.3% of CIP could be broken down by ompg-C3N4, which was around 2.9 times more than that of bulk g-C3N4. The impact of the ompg-C3N4 concentration on the photoinduced degradation of CIP is seen in Figure 6b. However, when the excess catalyst responds by scattering or reflecting photons to avert the photocatalyst’s excitation, subsequent increases in catalyst loading cause an undetectable decline in degradation efficiency.
Additionally, it was found that the number of active sites decreased when photocatalyst particles aggregated with increasing photocatalyst concentrations [110]. The influences of varying starting concentrations of CIP by a simulated sunlight ompg-C3N4 process are shown in Figure 6c. An increase in final substrate concentrations was found to cause the rate constants of CIP to drop. One description might be that the low breakdown rate of CIP was caused by larger concentrations of CIP absorbing incoming light, which decreased the photons needed to activate ompg-C3N4 [111].
Figure 6. (a) Absorption, photocatalysis, and photolysis of CIP under simulated sunlight exposure are compared; (b) impact of ompg-C3N4 dose on CIP’s photocatalytic degradation kinetics, (c) impact of the initial concentration of CIP on the kinetics of its photocatalytic breakdown, and (d) CIP degradation kinetic rate constants at various pH levels. Reproduced with permission [109].
Figure 6. (a) Absorption, photocatalysis, and photolysis of CIP under simulated sunlight exposure are compared; (b) impact of ompg-C3N4 dose on CIP’s photocatalytic degradation kinetics, (c) impact of the initial concentration of CIP on the kinetics of its photocatalytic breakdown, and (d) CIP degradation kinetic rate constants at various pH levels. Reproduced with permission [109].
Molecules 30 03712 g006
As shown in Figure 6d, the impact of pH level on the ompg-C3N4-mediated photo-induced degradation of CIP was assessed. It was discovered that raising the pH caused the rate constants of CIP to increase. However, the degradation efficiency decreased when the pH increased further [109].

3.3. Degradation of Amoxicillin

Antimicrobial resistance caused by antibiotic misuse is a global problem that jeopardizes both human health and the ecosystem. Dingxin et al. described a photocatalytic method that is both operative and environmentally benign for breaking down the most commonly used antibiotic, amoxicillin. Better than the individual components, a special 2D/2D Bi2WO6/Ti3C2 MXene heterojunction with facet-to-face contact was carefully designed to enable total amoxicillin removal in less than 40 min. It was discovered that the composite photocatalyst enhanced the capacity for light adsorption, facilitated rapid electron-hole transfer, and exhibited improved photothermal adaptation characteristics, which are key components of its photocatalytic mechanism. These properties essentially facilitate the generation of reactive oxygen species, particularly photogenerated holes (h+) and superoxide anion radicals (•O2). Advances in nanostructured catalyst design and photocatalytic methods are enabling innovative strategies for environmental remediation, including more effective wastewater treatment and the degradation of persistent organic contaminants [112].
The samples were exposed to full-spectrum sun radiation for one hour to assess their amoxicillin-degrading competence. The BT0.5, BT1, and BT1.5 hybrids all display better photocatalytic activity against amoxicillin than bare Ti3C2 and bare Bi2WO6, as seen in Figure 7a. With over 100% amoxicillin degradation after 40 min, BT1 in particular has the most excellent photocatalytic activity of any sample and is on par with other photocatalysts that have been described. Since additional Ti3C2 speeds up electron transport during the photocatalytic process, growing the Ti3C2 level increases the rate of deterioration. The redundant Ti3C2, which partly hinders light absorption and protects the active sites on the Bi2WO6 surface, is the reason for the modest performance drop caused by an excessive loading quantity of 1.5%. The pseudo-first-order approximation, as shown in Figure 7b, is followed by all photocatalytic processes, where the introduction of Ti3C2 results in a rise in apparent rate constants (k).
Figure 7. (a) Amoxicillin’s photodegradation and (b) the various samples’ degradation rate constants, and (c) Three amoxicillin degradation cycling tests over BT1. Reproduced with permission [112].
Figure 7. (a) Amoxicillin’s photodegradation and (b) the various samples’ degradation rate constants, and (c) Three amoxicillin degradation cycling tests over BT1. Reproduced with permission [112].
Molecules 30 03712 g007
The greatest deterioration rate constants are found in BT1, which is five times more than that of pristine Bi2WO6. Furthermore, to investigate the stability of the synthesized hybrids, cyclic tests have been carried out under the same conditions, as illustrated in Figure 7c. After three successive loops, the degradation rate for BT1 remains unchanged, mainly, showing the hybrid catalysts’ outstanding durability and reusability for real-world environmental requests. The LC-MS analysis identified the intermediates produced during the BT1-driven photodegradation of amoxicillin. Based on these intermediates, three potential routes for the breakdown of amoxicillin are recommended, as shown in Figure 8. Blue arrows point to Pathway 1, where the β-lactam ring is first lipidated to produce amoxicillin penicillanic acid [112].
Figure 8. Amoxicillin’s suggested photodegradation paths across BT1. Reproduced with permission [112].
Figure 8. Amoxicillin’s suggested photodegradation paths across BT1. Reproduced with permission [112].
Molecules 30 03712 g008

3.4. Degradation of Enrofloxacin

MXenes, a new type of two-dimensional transition metal carbides or nitrides, have gathered remarkable attention for a range of uses because of their special properties, which include ion intractability, hydrophilicity, and high electrical conductivity. Using a rapid and easy microwave hydrothermal method by applying heat treatment on the HCl/NaCl mixture solution, Ti3C2 MXene, or MX, is converted into MX-TiO2 composites. This process creates fine TiO2 nanoparticles on the M-X parent structure and gives the resultant MX-TiO2 composites photocatalytic activity, as shown in Figure 9a. The composites were used to remove the common contaminated antibiotic enrofloxacin (ENR) from water. By adjusting the temperature of hydrothermal, the proportion of M-X and TiO2 may be regulated, producing nanocomposites with adjustable adsorption and photocatalytic capabilities. Since composites made without NaCl were unable to adsorb enrofloxacin effectively, it was shown that the inclusion of NaCl had a significant impact. Sodium ions are concurrently intercalated into the nanocomposite structure when NaCl is added to the hydrothermal process, meaningfully increasing ENR adsorption from 1 to 6 mg ENR/g composite.
Additionally, it decreases the speed of the conversion of MX to TiO2, which results in a more uniform and smaller distribution of TiO2 particles on the structure. Even while composites without NaCl had a greater TiO2 content, MX-TiO2/NaCl nanocomposites, which had sodium intercalated in their structures, demonstrated stronger ENR adsorption and photocatalytic activity. High photocatalytic degradation efficiency results from the effective destruction of adsorbed ENR on the composites by free radicals produced by the photoexcited TiO2 particles. Sukidpaneenid et al. illustrated how adsorption and photocatalytic degradation of the produced chemicals work in concert [65]. The SEM of pure MX and MX-TiO2 composites made with and without NaCl at various hydrothermal temperatures are shown in Figure 9. As shown in Figure 9a, Pure MX exhibits a characteristic accordion-like layer structure. Several nanoparticles were seen to develop on the surface and between the layers of the MXene structure after hydrothermal treatment at 175 °C and 200 °C, as seen in Figure 9b,c. The particles that progress in systems with NaCl addition, Figure 9d seem to be smaller than those in the system without NaCl.
Even though they are now hardly discernible, the composites’ sheet-like and layered structure is still noticeable. During this hydrothermal process, the sodium cations have inhibited the development of the TiO2 particles on the MXene nanosheets, or NaCl, intercalated between the MXene layers. Several fine TiO2 particles were observed to form on the Ti3C2 nanosheets upon thermal oxidation in air [113] or solvothermal in liquid media of ethylene glycol, HF, and isopropyl alcohol [114]. These structures of the MX-TiO2 nanocomposite are reasonably similar to those reported in the literature.
Figure 9. (a) Schematic representation of the production of Ti3C2-TiO2 (MX-TiO2) composites, with or without NaCl. SEM images of (b) MX-TiO2 at 175 °C/NaCl, (c) MX-TiO2 at 200 °C/NaCl, and (d) MX-TiO2 at 225 °C/NaCl. Reproduced with permission [65].
Figure 9. (a) Schematic representation of the production of Ti3C2-TiO2 (MX-TiO2) composites, with or without NaCl. SEM images of (b) MX-TiO2 at 175 °C/NaCl, (c) MX-TiO2 at 200 °C/NaCl, and (d) MX-TiO2 at 225 °C/NaCl. Reproduced with permission [65].
Molecules 30 03712 g009
The ENR was eliminated using MX and MX-TiO2 composites in a dark environment with UVA irradiation. After the materials had been incubated with ENR for eight hours in the dark, the UVA irradiation was carried out. Five hours later, the ENR solution was poured and replaced with a new ENR solution. The ENR elimination effectiveness in a dark environment is shown in Figure 10a. The tidy MX has a removal effectiveness of about 92% because of its excellent ability to absorb ENR. The tidy MX’s cation exchange is most probably the leading cause of this. It has been revealed that MX etched by LiF contains Li cations in its structure, which may help MX adsorb cations [115].
Figure 10. Enrofloxacin (ENR) removal by MX and MX-TiO2 composites as a function of time under (a) 8 + 5 h of dark conditions and (b) 8 h of dark conditions plus 5 h of UVA irradiation. Reproduced with permission [65].
Figure 10. Enrofloxacin (ENR) removal by MX and MX-TiO2 composites as a function of time under (a) 8 + 5 h of dark conditions and (b) 8 h of dark conditions plus 5 h of UVA irradiation. Reproduced with permission [65].
Molecules 30 03712 g010
On the other hand, the MX-TiO2 that has been hydrothermally treated with 1 mol/L HCl has a minimal quantity of Li in its structure and can thus only remove 11.2% of the ENR. The sharp decline in the MX-TiO’s Li concentration. The effectiveness of ENR elimination under UVA irradiation is seen in Figure 10b. To characterize the ENR removal efficiency of the materials that is attributable to the UVA irradiation effect alone, the curves under UVA irradiation were previously normalized by the ENR removal efficiency under the dark conditions at the same respective time in Figure 10a. For comparison, the ENR photolysis under UVA was also shown. MX had very little UV photodegradation activity against ENR, as shown by the fact that its ENR removal effectiveness under UVA irradiation was comparable to the drop in ENR concentration brought on by photolysis itself [65].

3.5. Degradation of Norfloxacin

It belongs to the class of fluoroquinolones, used to treat urinary tract infections [116]. Norfloxacin is now detected in wastewater at high concentrations, particularly in hospitals, and is considered a potential contaminant in the aquatic environment. Because of their extensive usage and environmental toxicity, fluoroquinolone antibiotics have caused a great deal of worry in recent years [116,117]. Numerous studies have shown the photocatalytic degradation of norfloxacin utilizing various materials [118]. The experimental results showed excellent photocatalytic performance toward the degradation of norfloxacin across multiple water matrices. Sayed et al. [119] prepared a novel immobilized TiO2/Ti film with exposed facets via a simple one-pot hydrothermal route to be used in the degradation of norfloxacin from aqueous media. It was observed that OH is primarily involved in the photocatalytic degradation of norfloxacin by faceted TiO2/Ti film, as shown in Figure 11a. Furthermore, Tang et al. [120] observed outstanding performance of visible light-driven photocatalysis for the sol–gel approach, followed by a photo reduction procedure to break down norfloxacin utilizing a unique Z-scheme Ag/FeTiO3/Ag/BiFeO3 as produced. When employing Ag/FeTiO3/Ag/BiFeO3 at 2.0 weight percent Ag, the findings show that the photocatalytic degradation extent reaches 96.5% after 150 min and can be reprocessed with outstanding photocatalytic stability, as shown in Figure 11b–d.
Additionally, Lv et al. [121] used a solvothermal technique to create copper-doped bismuth oxybromide and evaluated its capacity to break down norfloxacin in the presence of visible light. Cu-doped BiOBr, as prepared, demonstrated high activity in the photocatalytic degradation of norfloxacin under visible-light irradiation, with a photocatalytic degradation constant. This was due to its improved light-harvesting properties, enhanced charge separation, and interfacial charge transfer. Additionally, it retained 95% of its initial activity even after five continuous catalytic cycles. Similarly, Tang et al. [122] proposed Bi2WO6, another bismuth-based catalyst, and used it to photodegrade norfloxacin in a nonionic surfactant Triton-X100 dispersion when exposed to visible light. The findings showed that adding TX100 might significantly increase the breakdown of scarcely soluble norfloxacin. At the critical micelle concentration, TX100 stimulated norfloxacin photodegradation and was highly adsorbed on the Bi2WO6 surface. The degradation of antibiotics by MXene based materials are summarized in Table 1.
Figure 11. (a) Faceted TiO2/Ti film photocatalysis mechanism. Reproduced with permission [111]. (b) A potential Z-scheme mechanism diagram System Ag/FeTiO3/Ag/BiFeO3, (c) The impact of FeTiO3 and BiFeO3 mass ratios, and (d) the kinetics of degradation reactions on photocatalytic activity. Reproduced with permission [120].
Figure 11. (a) Faceted TiO2/Ti film photocatalysis mechanism. Reproduced with permission [111]. (b) A potential Z-scheme mechanism diagram System Ag/FeTiO3/Ag/BiFeO3, (c) The impact of FeTiO3 and BiFeO3 mass ratios, and (d) the kinetics of degradation reactions on photocatalytic activity. Reproduced with permission [120].
Molecules 30 03712 g011
Table 1. Degradation of antibiotics by MXene-based material.
Table 1. Degradation of antibiotics by MXene-based material.
MXene-Based MembranesMechanismAntibioticsReference
Ti3C2Tx/CNFsElectrostatic repulsion/sievingazithromycin[123]
Ti3C2TxElectrostatic repulsion/molecular sievingErythromycin[124]
Ti3C2TxElectrostatic repulsion/molecular sievingPenicillin[124]
Ti3C2TxElectrostatic repulsion/molecular sievingRifampicin[124]
Ti3C2TxElectrostatic repulsion/molecular sievingBacitracin[124]
Ti3C2TxPhotocatalytic inactivationE. coli[125]
g-C3N4@MXeneElectrostatic repulsion/molecular sievingTetracycline hydrochloride[126]
g-C3N4/TiO2/kaoliniteChemical stripping and self-assemblyciprofloxacin[127]
g-C3N4/C-TiO2S-scheme heterojunctionciprofloxacin[128]
sepiolite/g-C3N4/PdPhotocatalytic mechanismciprofloxacin[129]
Au@ZnONPs-rGO-gC3N4photodegradation mechanismCiprofloxacin,
Levofloxacin
[130]

4. MXene in Water Splitting

Electrocatalytic water splitting holds considerable promise as a sustainable and eco-friendly method for producing hydrogen and oxygen, with the latter serving as a renewable alternative to conventional fossil fuels. As an electrocatalyst for water splitting, MXene, a unique two-dimensional (2D) layered family of materials, has drawn a lot of interest. This adaptable substance can be modified to enhance its stability and electroactive surface properties for improved electrocatalytic activity [131]. The development of inexpensive, non-noble metal-based electrocatalysts for water splitting has garnered significant interest due to their potential to provide clean, green hydrogen fuel. A family of two-dimensional transition metal nitrides, carbides, and carbonitrides was discovered in 2011. Because of their high electrical conductivity, huge surface area, and abundance of catalytically active sites, these metals have shown potential performance as electrocatalysts in the water splitting process. However, the agglomeration and restacking of MXene flakes restrict their long-term stability and recyclability. By mixing MXene with other materials to generate hybrid structures, which have shown better electrocatalytic activity than pure MXenes, this issue may be addressed. The two half-cell processes involved in electrolysis of water are the OER at the anode and the HER at the cathode [132].

4.1. Hydrogen Evolution Reaction

Because of its immediate use in a variety of sectors, including water electrolysis, and relative simplicity, the hydrogen evolution process HER is one of the electrochemical reactions that has been considered the most. The HER exhibits noticeably quicker kinetics when it takes place on noble metal electrodes, especially those made of platinum group metals (PGM), as opposed to the sluggish kinetics seen in the OER and oxygen reduction reaction (ORR). Practical current densities may be achieved at a few tens of millivolts of overpotential due to these improved kinetics [133,134]. Inquiries into the mechanism of HER on metallic surfaces began in the early 1950s and were primarily focused on nickel [133]. Two electrons are transferred during the two-step hydrogen evolution reaction at the cathode, which also includes the adsorption and desorption of intermediate molecules. The Volmer step is the first stage of HER in an acidic environment, where hydrogen is reduced by one electron and then adsorbed at the cathode surface (H*) [135,136]. Depending on the quantity of adsorbed hydrogen atoms, there are two possible routes for the production of hydrogen molecules after the preceding step. When the H* atom’s concentration is very low, it takes an electron and a proton from the water, which causes H2 to desorb from the cathode surface. The sequence is known as the Volmer–Heyrovsky mechanism, and the process is known as the “Heyrovsky step.” Additionally, when the amount of H* is considerable, two nearby H* atoms on the cathode’s surface combine to form a hydrogen molecule. According to HER diagrams in acidic and alkaline environments, this process is referred to as the “Tafel step,” and the sequence of stages is known as the Volmer–Tafel mechanism, as shown in Figure 12 [133,137,138], which shows the schematics of HER in acidic and alkaline environments. The primary mechanism of the HER for different catalysts can often be identified by comparing the Tafel slope values [139].
Figure 12. HER schematics in alkaline and acidic environments. Reproduced with permission [138].
Figure 12. HER schematics in alkaline and acidic environments. Reproduced with permission [138].
Molecules 30 03712 g012

MXene-Based Nanocomposites for HER

The surface shape of MXenes has a significant impact on their electrocatalytic activity. Yuan et al. [140] used a two-step procedure that included hydrolyzing bulk MAX ceramics and then HF etching to create an electrocatalyst based on highly active MXene nanofibers with a sizable specific surface area. MXene nanofibers have a Tafel slope of 97 mV/dec and a low overpotential of 169 mV at 10 mA/cm2, which are noticeably greater than those of Ti3C2 flakes. Ti3C2 Mxene nanofibers perform better than conventional MXene nanosheets because of their larger specific surface area, which exposes more active spots on their cross-section. Jia-Jun et al. effectively synthesized 3D-MoSe2@Ti-MXene nanoflowers with exclusive nanoscale petal morphologies using a simple hydrothermal technique. The synthesized MoSe2@Ti-MXene’s electrocatalytic properties in a 0.5 M H2SO4 solution for HER were examined. According to Figure 13a, which compares the electrocatalytic performance of Ti3C2 MXene and MoSe2 as well as various hybrid MoSe2/MXene nanoflowers, the MoSe2@Ti-MXene hybrid performed noticeably better than bare MoSe2. It showed an overpotential of 180 mV and the lowest onset potential, at 61 mV, at 10 mA/cm2. In comparison, the bare MoSe2 needed an overpotential of 300 mV to achieve the same current density of 10 mA/cm2. A slower charge transfer rate is shown by the Tafel slope of 200 mV/dec for bare MoSe2, as seen in Figure 13b. The Tafel slope, however, was significantly dropped to only 91 mV/dec for the MoSe2@Ti-MXene composite, suggesting a quicker charge transfer between MoSe2 and Ti-MXene. The study’s findings indicate that the hybrid structure’s enhanced electrocatalytic performance is a result of its increased conductivity and effective electron transport [141].
Figure 13. (a) LSV curves and (b) Tafel plots of Ti3C2 MXene, MoSe2, and various MoSe2/MXene hybrid nanoflowers. Reproduced with the permission [141].
Figure 13. (a) LSV curves and (b) Tafel plots of Ti3C2 MXene, MoSe2, and various MoSe2/MXene hybrid nanoflowers. Reproduced with the permission [141].
Molecules 30 03712 g013
Reghunath and his co-researchers have formed and evaluated a flexible hierarchical composite of Bismuth ferrite/chromium carbide as BiFeO3/Cr2CTx MXene as an electrocatalyst for water splitting. This strategy suggested a simple way to create Cr2CTx MXene from the chromium aluminum carbide MAX Phase. The elimination of aluminum atomic layers from the Cr2AlC MAX structure was confirmed by high-resolution transmission electron microscopy, scanning electron microscopy, and X-ray diffraction investigations. With a low overpotential of 128 mV in 1 M KOH, electrochemical experiments showed that the BiFeO3/Cr2CTx MXene composite, which was made with a decreased Al2O3 composition, performed very well in the hydrogen evolution process. The Tafel slope and charge transfer resistance have been calculated to be 53.3 mV/dec and 0.16Ω, respectively. The BiFeO3/Cr2CTx MXene electrode used only 1.81 V of cell potential in a di-electrode electrolysis system to generate 10 mA/cm2 with sustained stability. This work proposed a simple method for creating Cr2CTx MXene composites for HER applications while highlighting the possible use of BiFeO3/Cr2CTx MXene in HER [142].
The electrical characteristics of the various transition metal components prejudice the HER activity of MXenes. According to Zhi et al., Mo2CTx MXenes have superior HER activity compared to Ti2CTx and may be used as an active and stable catalyst for HER in acid [143]. In this study, Mo2Ga2C and Ti2AlC are used as raw materials for HF etching, which yields Mo2CTx and Ti2CTx MXene, respectively. At 10 mAcm2, Mo2CTX had a lower initial overpotential of 283 mV than Ti2CTX, at 609 mV. The catalyst was unstable, as shown by the additional reduction in Ti2CTx’s HER activity during continuous cycling. After around 30 cycles, Mo2CTx achieves an overpotential of 305 mV at 10 mA/cm2, exhibiting only a modest initial drop in HER activity as compared to Ti2CTx, Mo1.33CTx, and Mo2CTx, which are two Mo-based MXenes that were created by Intikhab et al. [144]. They discovered that.
In contrast, operational durability had little to no effect on HER activity; stoichiometry and atomic surface structure had a significant influence. The HER activity of Mo1.33CTx is lower than that of Mo2CTx. The materials’ over potentials are 422 and 239 mV at 10 mA/cm2, respectively. The variance in surface electronic characteristics and the loss of hcp sites with a core six-fold coordinated carbon atom, which lowers the fraction of the ideal -O terminated, are the causes of Mo1.33CTx poor HER activity. These discoveries have the potential to be applied to other technology-related electrochemical processes and provide essential insights into the creation and advancement of molybdenum-based 2D materials [145,146,147].
In the HER process, the conductive Cr2C and Cr2CO2 are advantageous for electron transport [148]. Transition metal modification or hydrogen coverage may modify the free energy of hydrogen adsorption. The creation of carbon vacancy engineering for Cr2CO2 MXene may also improve HER results. The preparation of 2D Cr-based MXene for potential electrocatalytic water splitting is made possible by these first findings. A new structure of MXene nanofibers as Ti3C2 NFs was created by Yuan et al. [140], and the nanofiber morphology displayed a thickness of 40–60 nm. Additionally, the MXene NFs synthetic approach can be expanded to include the synthesis of other MXenes and applied to promising applications, such as catalysis, batteries, and supercapacitors. Nearly every prior study up to this point has referred to MXene as metal. But according to Abdoulaye et al., the 2D-Ti3N4 MXene has both semiconductor and metallic characteristics [149]. Additionally, this work demonstrates the occurrence of Tx = O and/or OH, which is susceptible to surface oxidation, and offers an integrated structural characterization of exfoliated Ti4N3Tx MXene. With a Tafel slope of around 0.19 V/dec and an over potential of 0.3 V at 10 mA/cm2, the exfoliated Ti4N3Tx shows excellent HER performance.
With its hexagonal structure and sandwich-like arrangement of sulfur and molybdenum atoms, molybdenum disulfide (MoS2) is a well-known member of the transition metal dichalcogenides (TMDCs) family. Using MoS2, Ti3C2, and carbon nanofibers (CNFs), Ma et al. created a very effective and stable HER electrocatalyst, as shown in Figure 14a. Ti3C2 and CNFs were used to develop a plane-like skeletal structure in the first stage, which helped to stop the MXene flakes from restacking. A stereostructured MoS2/Ti3C2@CNFs was then achieved by the spontaneous development of MoS2 on the fiber skeleton. Multiple magnification SEM images of MoS2/Ti3C2@CNFs are shown in Figure 14b,c, respectively. This special structure led to a notable increase in electrical conductivity and a more extensible layered structure.
Figure 14d,e displays the HER polarization curves for the catalysts MoS2, MoS2/Ti3C2@CNFs, MoS2/Ti3C2, and Ti3C2@CNFs in a 0.5 M sulfuric acid electrolyte and Tafel plots of different catalysts, respectively. For comparison, the catalytic performance of commercial Pt/C was also examined. At a HER current of 10 mA/cm2, the MoS2/Ti3C2@CNFs catalyst demonstrated a much-reduced over potential of 142 mV vs. RHE in contrast to 471 mV for MoS2/Ti3C2 and 596 mV for MoS2 alone. This suggests that the fiber skeleton structure of MoS2/Ti3C2@CNFs significantly enhanced the HER’s catalytic performance, demonstrating that CNFs successfully gave the catalyst a solid framework that enabled efficient MoS2 loading and demonstrated outstanding HER activity. Moreover, the HER performance of Mxene based materials are summarized in Table 2.
Figure 14. (a) MoS2/Ti3C2@CNFs synthetic schematic. MoS2/Ti3C2@CNFs SEM pictures at varying magnifications; (b) 12,000×, (c) 1800×, (d) the polarization curves of MoS2, MoS2/Ti3C2, Ti3C2@CNFs, MoS2/Ti3C2@CNFs, and Pt/C catalysts, and (e) Tafel plots of several catalysts. Reproduced with permission [150].
Figure 14. (a) MoS2/Ti3C2@CNFs synthetic schematic. MoS2/Ti3C2@CNFs SEM pictures at varying magnifications; (b) 12,000×, (c) 1800×, (d) the polarization curves of MoS2, MoS2/Ti3C2, Ti3C2@CNFs, MoS2/Ti3C2@CNFs, and Pt/C catalysts, and (e) Tafel plots of several catalysts. Reproduced with permission [150].
Molecules 30 03712 g014
Table 2. HER performance of MXene-based materials as electrocatalysts.
Table 2. HER performance of MXene-based materials as electrocatalysts.
ElectrocatalystElectrolyteTafel SlopeOverpotentialReferences
Ti2CTx0.5M H2SO4100 mV/dec75 mV[151]
Ti3C2Tx0.5M H2SO497 mV/dec169 mV[151]
Pd/Ti3C2Tx–CNT0.1M KOH50 mV/dec158 mV[152]
Ru@Ti3C2Tx-Vc1M KOH32 mV/dec35 mV[153]
Pt/Ti3C2Tx0.5M H2SO429.734[154]
Ti3CN(OH)x@MoS20.5M H2SO464120[155]
Rh–CoNi LDH/MXene1.0M KOH43.974.6[156]
Co-NCNT/Ti3C2Tx1.0M KOH78190[157]
BiFeO3/Cr2CTx1.0 M KOH53.3128[142]
WS2@MXene/GO1.0M KOH5843[158]
Ti3C2Tx: Co1.0M KOH103.3103.6[159]

4.2. Oxygen Evolution Reaction

The studies by Conway et al. have resulted in a strong link between the redox potential of the metal/metal oxide pair and the voltage needed for oxygen production on a metal surface. Even for noble metals, the release of oxygen from the metal surface requires the creation of the appropriate metal oxide. The OER mostly occurs on the hydroxide, oxyhydroxide, or oxide layer that spontaneously develops on the electrocatalyst’s surface, according to recent research that has confirmed this conclusion [133,160]. In contrast to AEM, the lattice oxygen mechanism (LOM) releases oxygen by oxidizing the catalyst’s lattice oxygen. In particular, the lattice oxygen itself tends to participate in the OER process to produce oxygen after undergoing oxidation under the OER potential via lattice oxygen redox chemistry. The LOM often outperforms the traditional AEM in terms of explaining the underlying cause of the exceptional catalytic activity seen in specific solid-phase catalysts. As a result, building more effective OER electrocatalysts may be guided by a thorough knowledge of the reaction process and pertinent parameters [161]. The fundamental working of OER is shown in Figure 15.
Figure 15. The overall OER’s fundamental mechanism in which (a) AEM, (b) LOM of SMSM, (c) LOM of OVSM and (d) LOM of DMSM. Reproduced with permission [161].
Figure 15. The overall OER’s fundamental mechanism in which (a) AEM, (b) LOM of SMSM, (c) LOM of OVSM and (d) LOM of DMSM. Reproduced with permission [161].
Molecules 30 03712 g015

MXene-Based Nanocomposites for OER

Numerous research teams have documented the coupling of sulfides with MXene, which are frequently used as effective catalysts for OER. A FeS2@MXene catalyst was made by Xie et al., who also looked at how well it performed for OER. In an alkaline solution, the resulting catalyst exhibited a minor over potential of 240 mV at 10 mA/cm2. The scientists discovered that by successfully inhibiting aggregation and modifying the electron density and electrophilicity of the active center, the addition of MXene may significantly enhance the OER activity and stability of FeS2 [162].
A CoS2@MXene catalyst with spatial electron rearrangement, ordered structures, and strong anchoring was described by Han et al. It was discovered that the inclusion of MXene enhanced the intrinsic activity and solidity of CoS2 active sites in addition to enriching their density [163]. The design of MXene@Co3S4/Ni3S2 on nickel foam, as shown in Figure 16A, and its performance for OER in alkaline (1M KOH) were reported by Li et al. [164]. The synthesis consists of etching and delamination of Ti3C2Tx MXene, heterogeneous development of CoNi-LDH on MXene, represented in Figure 16B, and wet chemical sulfuration Figure 16C–E, to convert it to CoNi sulfide. With an ultralow over potential of 186 mV, long-term stability for 24 h, and exceptional voltammetric cycling stability for 1000 cycles, the hybrid catalyst exhibits exceptional OER performance. According to DFT calculations, the active center with the lowest barrier for *O to *OOH conversion is the Ni atom rather than the Co atom. A porous NiCoS on Ti3C2Tx was reported by Zou et al. as an effective OER catalyst in a related study [165]. Except for using ZIF-67 as the LDH precursor and avoiding nickel foam, the synthesis is identical to that of Li et al., as shown in Figure 16F [164]. NiCoS and Ti3C2Tx MXene were detected by Raman spectroscopy.
The peak shift in Figure 16G of NiCoS demonstrated the significant electrical interaction between NiCoS and Ti3C2Tx-MXene. MXene’s two-dimensional structure enables a high surface area, as illustrated in Figure 16H for the hybrid catalyst, making it a viable option for zinc-air battery cathode applications that require long-term durability and minimal charging/discharging over potential. In comparison to the pure phosphide equivalent, the FeP-CoP/Ti3C2Tx composite shows better electrochemical performance for the OER in alkaline media, with a smaller Tafel slope and a lower over potential. Additionally, the combination outperformed several different phosphides made from Prussian blue analogs (PBAs) as well as commercial RuO2. More active sites, increased intrinsic activity, and quicker charge transfer kinetics made possible by the addition of Ti3C2Tx MXene were cited as the reasons for the FeP-CoP/Ti3C2Tx composite’s improved performance [166]. The Table 3 represents the OER performance of various MXene based electocatalysts.
Figure 16. (A) NiFeCoP/MXene preparation, (B) XRD, (C) SEM, (D) TEM, and (E) EDS mapping, (F) LSV curves and (G) OER over potential and (H) Large surface area. Reproduced with permission [167].
Figure 16. (A) NiFeCoP/MXene preparation, (B) XRD, (C) SEM, (D) TEM, and (E) EDS mapping, (F) LSV curves and (G) OER over potential and (H) Large surface area. Reproduced with permission [167].
Molecules 30 03712 g016aMolecules 30 03712 g016b
Table 3. OER performance of MXene-based materials as electrocatalysts.
Table 3. OER performance of MXene-based materials as electrocatalysts.
ElectrocatalystTafel slopeOverpotentialReferences
IrCo@ac-Ti3C260 mV/dec220 mV[168]
Ru–FeOOH@ Ti3C2Tx67.7 mV/dec230 mV[169]
Cr–FeNi LDH/MXene54.4232[170]
Co–B@Ti3C2Tx53250[171]
Co3O4@Ti3C2Tx118300[172]
H2PO2-/FeNiLDH-V2C46.5250[173]
CoFe-LDH/Ti3C250319[174]
NiCo-LDH/Ti3C2Tx/NF47.2223[175]
CDs@(PdFeNiCo)Nbx62240[170]

5. Conclusions

Environmental pollution and environmentally friendly power generation are two of the world’s most urgent problems. The MXene-based composites have become a revolutionary class of materials in recent years. They are particularly well-suited for use in the breakdown of antimicrobial pollutants as well as water splitting to obtain hydrogen due to their distinctive mix of excellent electrical conductivity, huge surface area, customizable surface chemistry, and structural plasticity. The review demonstrates how MXenes exhibit synergistic effects that significantly enhance their catalytic and adsorptive properties when incorporated into composite materials with metal oxides, sulfides, polymers, carbon nanostructures, and other 2D materials. MXene-based photocatalysts excel in producing reactive oxygen species, facilitating charge separation, and enabling redox reactions to degrade persistent pharmaceutical residues, such as fluoroquinolones and tetracyclines. These composites are promising candidates for long-term water purification technologies, as they not only demonstrate high degradation efficiency but also excellent recyclability and structural stability after multiple treatment cycles. Their surface terminations and conductive nature reduce over potential, speed up electron transfer, and increase accessibility to the active site. Higher photocurrent densities and steady hydrogen production rates result from the capacity to create heterogeneous structures with semiconductors or metal oxides, which also improve charge carrier mobility and inhibit recombination. Significant safety and environmental risks are associated with the traditional synthesis of MXenes, which involves hazardous chemicals, including hydrofluoric acid. For practical implementation, scalability, repeatability, and surface stability remain major obstacles that must be addressed.
Moreover, the combination of fluorine-free etching and composite design methods holds excellent prospects for overcoming the in-built challenges related to MXene-based materials. These innovations provide the development of environmentally friendly fabrication techniques and the formation of MXene composites with tailored properties, thereby enhancing stability, performance, and functionality. As research in these areas progresses, we can expect more efficient and sustainable MXene materials to emerge, paving the way for broader applications in energy, electronics, environmental, and catalytic industries.

Author Contributions

S.I.: Writing—review and editing, Writing—original draft, S.L. and S.B.K.: review and editing; S.A. and S.I.; Conceptualization and Funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This work was partly financially supported by the Project of State Key Laboratory of Environment-Friendly Energy Materials (227/35100011) and the authors extend their appreciation to the Deanship of Research and Graduate Studies at King Khalid University for funding this work through Large Research Project under grant number RGP2/643/46.

Data Availability Statement

The data will be available on request.

Conflicts of Interest

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

References

  1. Babin, A.; Vaneeckhaute, C.; Iliuta, M.C. Potential and challenges of bioenergy with carbon capture and storage as a carbon-negative energy source: A review. Biomass Bioenergy 2021, 146, 105968. [Google Scholar] [CrossRef]
  2. Wang, P.; Yang, Y.; Xue, D.; Ren, L.; Tang, J.; Leung, L.R.; Liao, H. Aerosols overtake greenhouse gases causing a warmer climate and more weather extremes toward carbon neutrality. Nat. Commun. 2023, 14, 7257. [Google Scholar] [CrossRef] [PubMed]
  3. Safiah Yusmah, M.; Bracken, L.J.; Sahdan, Z.; Norhaslina, H.; Melasutra, M.; Ghaffarianhoseini, A.; Sumiliana, S.; Shereen Farisha, A. Understanding urban flood vulnerability and resilience: A case study of Kuantan, Pahang, Malaysia. Nat. Hazards 2020, 101, 551–571. [Google Scholar] [CrossRef]
  4. Morrison, T.H.; Adger, W.N.; Agrawal, A.; Brown, K.; Hornsey, M.J.; Hughes, T.P.; Jain, M.; Lemos, M.C.; McHugh, L.H.; O’Neill, S.; et al. Radical interventions for climate-impacted systems. Nat. Clim. Change 2022, 12, 1100–1106. [Google Scholar] [CrossRef]
  5. Yates, J.; Deeney, M.; Muncke, J.; Carney Almroth, B.; Dignac, M.-F.; Castillo, A.C.; Courtene-Jones, W.; Kadiyala, S.; Kumar, E.; Stoett, P.; et al. Plastics matter in the food system. Commun. Earth Environ. 2025, 6, 176. [Google Scholar] [CrossRef]
  6. Zhou, S.; Di Paolo, C.; Wu, X.; Shao, Y.; Seiler, T.-B.; Hollert, H. Optimization of screening-level risk assessment and priority selection of emerging pollutants–The case of pharmaceuticals in European surface waters. Environ. Int. 2019, 128, 1–10. [Google Scholar] [CrossRef]
  7. Kosma, C.I.; Nannou, C.I.; Boti, V.I.; Albanis, T.A. Psychiatrics and selected metabolites in hospital and urban wastewaters: Occurrence, removal, mass loading, seasonal influence and risk assessment. Sci. Total Environ. 2019, 659, 1473–1483. [Google Scholar] [CrossRef]
  8. Vickers, N.J. Animal communication: When i’m calling you, will you answer too? Curr. Biol. 2017, 27, R713–R715. [Google Scholar] [CrossRef]
  9. Hofman-Caris, C.; Siegers, W.; van de Merlen, K.; De Man, A.; Hofman, J. Removal of pharmaceuticals from WWTP effluent: Removal of EfOM followed by advanced oxidation. Chem. Eng. J. 2017, 327, 514–521. [Google Scholar] [CrossRef]
  10. Aukema, K.G.; Escalante, D.E.; Maltby, M.M.; Bera, A.K.; Aksan, A.; Wackett, L.P. In silico identification of bioremediation potential: Carbamazepine and other recalcitrant personal care products. Environ. Sci. Technol. 2017, 51, 880–888. [Google Scholar] [CrossRef]
  11. Ha, H.; Mahanty, B.; Yoon, S.; Kim, C.-G. Degradation of the long-resistant pharmaceutical compounds carbamazepine and diatrizoate using mixed microbial culture. J. Environ. Sci. Health Part A 2016, 51, 467–471. [Google Scholar] [CrossRef]
  12. Duan, Y.; Deng, L.; Shi, Z.; Zhu, L.; Li, G. Assembly of graphene on Ag3PO4/AgI for effective degradation of carbamazepine under visible-light irradiation: Mechanism and degradation pathways. Chem. Eng. J. 2019, 359, 1379–1390. [Google Scholar] [CrossRef]
  13. Fu, Y.; Li, Z.; Liu, Q.; Yang, X.; Tang, H. Construction of carbon nitride and MoS2 quantum dot 2D/0D hybrid photocatalyst: Direct Z-scheme mechanism for improved photocatalytic activity. Chin. J. Catal. 2017, 38, 2160–2170. [Google Scholar] [CrossRef]
  14. Eswar, N.K.; Singh, S.A.; Madras, G. Photoconductive network structured copper oxide for simultaneous photoelectrocatalytic degradation of antibiotic (tetracycline) and bacteria (E. coli). Chem. Eng. J. 2018, 332, 757–774. [Google Scholar] [CrossRef]
  15. Li, Z.; Zhu, L.; Wu, W.; Wang, S.; Qiang, L. Highly efficient photocatalysis toward tetracycline under simulated solar-light by Ag+-CDs-Bi2WO6: Synergistic effects of silver ions and carbon dots. Appl. Catal. B Environ. 2016, 192, 277–285. [Google Scholar] [CrossRef]
  16. Zhu, M.; Kim, S.; Mao, L.; Fujitsuka, M.; Zhang, J.; Wang, X.; Majima, T. Metal-free photocatalyst for H2 evolution in visible to near-infrared region: Black phosphorus/graphitic carbon nitride. J. Am. Chem. Soc. 2017, 139, 13234–13242. [Google Scholar] [CrossRef]
  17. Mousavi, M.; Habibi-Yangjeh, A.; Pouran, S.R. Review on magnetically separable graphitic carbon nitride-based nanocomposites as promising visible-light-driven photocatalysts. J. Mater. Sci. Mater. Electron. 2018, 29, 1719–1747. [Google Scholar] [CrossRef]
  18. Akhundi, A.; Habibi-Yangjeh, A. Graphitic carbon nitride nanosheets decorated with CuCr2O4 nanoparticles: Novel photocatalysts with high performances in visible light degradation of water pollutants. J. Colloid Interface Sci. 2017, 504, 697–710. [Google Scholar] [CrossRef]
  19. Asadzadeh-Khaneghah, S.; Habibi-Yangjeh, A.; Abedi, M. Decoration of carbon dots and AgCl over g-C3N4 nanosheets: Novel photocatalysts with substantially improved activity under visible light. Sep. Purif. Technol. 2018, 199, 64–77. [Google Scholar] [CrossRef]
  20. Pirhashemi, M.; Habibi-Yangjeh, A. ZnO/NiWO4/Ag2CrO4 nanocomposites with pnn heterojunctions: Highly improved activity for degradations of water contaminants under visible light. Sep. Purif. Technol. 2018, 193, 69–80. [Google Scholar] [CrossRef]
  21. Zhang, T.; Wang, X.; Zhang, X. Recent progress in TiO2-mediated solar photocatalysis for industrial wastewater treatment. Int. J. Photoenergy 2014, 2014, 607954. [Google Scholar] [CrossRef]
  22. Wu, C.; Huang, Q. Synthesis of Na-doped ZnO nanowires and their photocatalytic properties. J. Lumin. 2010, 130, 2136–2141. [Google Scholar] [CrossRef]
  23. Li, X.; Zhang, W.; Li, J.; Jiang, G.; Zhou, Y.; Lee, S.; Dong, F. Transformation pathway and toxic intermediates inhibition of photocatalytic NO removal on designed Bi metal@ defective Bi2O2SiO3. Appl. Catal. B Environ. 2019, 241, 187–195. [Google Scholar] [CrossRef]
  24. Liao, J.; Cui, W.; Li, J.; Sheng, J.; Wang, H.; Dong, X.a.; Chen, P.; Jiang, G.; Wang, Z.; Dong, F. Nitrogen defect structure and NO+ intermediate promoted photocatalytic NO removal on H2 treated g-C3N4. Chem. Eng. J. 2020, 379, 122282. [Google Scholar] [CrossRef]
  25. Martin, D.J.; Umezawa, N.; Chen, X.; Ye, J.; Tang, J. Facet engineered Ag3PO4 for efficient water photooxidation. Energy Environ. Sci. 2013, 6, 3380–3386. [Google Scholar] [CrossRef]
  26. Midilli, A.; Ay, M.; Dincer, I.; Rosen, M.A. On hydrogen and hydrogen energy strategies: I: Current status and needs. Renew. Sustain. Energy Rev. 2005, 9, 255–271. [Google Scholar] [CrossRef]
  27. Cui, C.; Cheng, R.; Zhang, H.; Zhang, C.; Ma, Y.; Shi, C.; Fan, B.; Wang, H.; Wang, X. Ultrastable MXene@ Pt/SWCNTs’ nanocatalysts for hydrogen evolution reaction. Adv. Funct. Mater. 2020, 30, 2000693. [Google Scholar] [CrossRef]
  28. An, C.-H.; Kang, W.; Deng, Q.-B.; Hu, N. Pt and Te codoped ultrathin MoS2 nanosheets for enhanced hydrogen evolution reaction with wide pH range. Rare Met. 2022, 41, 378–384. [Google Scholar] [CrossRef]
  29. Wang, T.; Cao, X.; Jiao, L. Ni2P/NiMoP heterostructure as a bifunctional electrocatalyst for energy-saving hydrogen production. eScience 2021, 1, 69–74. [Google Scholar] [CrossRef]
  30. Cao, Z.; Zhou, T.; Ma, X.; Shen, Y.; Deng, Q.; Zhang, W.; Zhao, Y. Hydrogen production from urea sewage on NiFe-based porous electrocatalysts. ACS Sustain. Chem. Eng. 2020, 8, 11007–11015. [Google Scholar] [CrossRef]
  31. Li, Z.; Wang, W.; Qian, Q.; Zhu, Y.; Feng, Y.; Zhang, Y.; Zhang, H.; Cheng, M.; Zhang, G. Magic hybrid structure as multifunctional electrocatalyst surpassing benchmark Pt/C enables practical hydrazine fuel cell integrated with energy-saving H2 production. eScience 2022, 2, 416–427. [Google Scholar] [CrossRef]
  32. Wu, T.; Sun, M.-Z.; Huang, B.-L. Non-noble metal-based bifunctional electrocatalysts for hydrogen production. Rare Met. 2022, 41, 2169–2183. [Google Scholar] [CrossRef]
  33. Hussain, S.; Rabani, I.; Vikraman, D.; Feroze, A.; Ali, M.; Seo, Y.-S.; Song, W.; An, K.-S.; Kim, H.-S.; Chun, S.-H.; et al. MoS2@ X2C (X = Mo or W) hybrids for enhanced supercapacitor and hydrogen evolution performances. Chem. Eng. J. 2021, 421, 127843. [Google Scholar] [CrossRef]
  34. Guo, X.; Wan, X.; Liu, Q.; Li, Y.; Li, W.; Shui, J. Phosphated IrMo bimetallic cluster for efficient hydrogen evolution reaction. eScience 2022, 2, 304–310. [Google Scholar] [CrossRef]
  35. Zhou, F.; Zhou, Y.; Liu, G.-G.; Wang, C.-T.; Wang, J. Recent advances in nanostructured electrocatalysts for hydrogen evolution reaction. Rare Met. 2021, 40, 3375–3405. [Google Scholar] [CrossRef]
  36. Ali, A.; Shen, P.K. Recent progress in graphene-based nanostructured electrocatalysts for overall water splitting. Electrochem. Energy Rev. 2020, 3, 370–394. [Google Scholar] [CrossRef]
  37. Jiao, P.; Ye, D.; Zhu, C.; Wu, S.; Qin, C.; An, C.; Hu, N.; Deng, Q. Non-precious transition metal single-atom catalysts for the oxygen reduction reaction: Progress and prospects. Nanoscale 2022, 14, 14322–14340. [Google Scholar] [CrossRef]
  38. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  39. Sato, S.; White, J. Photodecomposition of water over Pt/TiO2 cata lysts. Chem. Phys. Lett. 1980, 72, 83–86. [Google Scholar] [CrossRef]
  40. Purabgola, A.; Mayilswamy, N.; Kandasubramanian, B. Graphene based TiO2 composites for photocatalysis and environmental remediation: Synthesis and progress. Environ. Sci. Pollut. Res. 2022, 29, 32305–32325. [Google Scholar] [CrossRef]
  41. Momeni, M.M.; Ghayeb, Y.; Ghonchegi, Z. Visible light activity of sulfur-doped TiO2 nanostructure photoelectrodes prepared by single-step electrochemical anodizing process. J. Solid State Electrochem. 2015, 19, 1359–1366. [Google Scholar] [CrossRef]
  42. Hammer, B.; Wendt, S.; Besenbacher, F. Water adsorption on TiO2. Top. Catal. 2010, 53, 423–430. [Google Scholar] [CrossRef]
  43. Dohnálek, Z.; Lyubinetsky, I.; Rousseau, R. Thermally-driven processes on rutile TiO2 (1 1 0)-(1 × 1): A direct view at the atomic scale. Prog. Surf. Sci. 2010, 85, 161–205. [Google Scholar] [CrossRef]
  44. Du, X.; Ji, H.; Xu, Y.; Du, S.; Feng, Z.; Dong, B.; Wang, R.; Zhang, F. Covalent organic framework without cocatalyst loading for efficient photocatalytic sacrificial hydrogen production from water. Nat. Commun. 2025, 16, 3024. [Google Scholar] [CrossRef]
  45. Fatima, J.; Tahir, M.; Rehman, A.; Sagir, M.; Rafique, M.; Assiri, M.A.; Imran, M.; Alzaid, M. Structural, optical, electronic, elastic properties and population inversion of novel 2d carbides and nitrides mxene: A dft study. Mater. Sci. Eng. B 2023, 289, 116230. [Google Scholar] [CrossRef]
  46. He, Q.; Tian, D.; Jiang, H.; Cao, D.; Wei, S.; Liu, D.; Song, P.; Lin, Y.; Song, L. Achieving efficient alkaline hydrogen evolution reaction over a Ni5P4 catalyst incorporating single-atomic Ru sites. Adv. Mater. 2020, 32, 1906972. [Google Scholar] [CrossRef]
  47. Li, H.; Tao, S.; Wan, S.; Qiu, G.; Long, Q.; Yu, J.; Cao, S. S-scheme heterojunction of ZnCdS nanospheres and dibenzothiophene modified graphite carbon nitride for enhanced H2 production. Chin. J. Catal. 2023, 46, 167–176. [Google Scholar] [CrossRef]
  48. Yao, L.; Gu, Q.; Yu, X. Three-dimensional MOFs@ MXene aerogel composite derived MXene threaded hollow carbon confined CoS nanoparticles toward advanced alkali-ion batteries. ACS Nano 2021, 15, 3228–3240. [Google Scholar] [CrossRef]
  49. Yang, Z.; Jiang, L.; Wang, J.; Liu, F.; He, J.; Liu, A.; Lv, S.; You, R.; Yan, X.; Sun, P.; et al. Flexible resistive NO2 gas sensor of three-dimensional crumpled MXene Ti3C2Tx/ZnO spheres for room temperature application. Sens. Actuators B Chem. 2021, 326, 128828. [Google Scholar] [CrossRef]
  50. Xiong, Q.-Q.; Muhmood, T.; Zhao, C.-X.; Xu, J.-S.; Yang, X.-F. Synergistic etching and intercalation enables ultrathin Ti3C2T x and Nb2CT x MXene nanosheets. Rare Met. 2023, 42, 1175–1185. [Google Scholar] [CrossRef]
  51. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.-E.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A. Electric field effect in atomically thin carbon films. Science 2004, 306, 666–669. [Google Scholar] [CrossRef] [PubMed]
  52. Zhu, J.; Ha, E.; Zhao, G.; Zhou, Y.; Huang, D.; Yue, G.; Hu, L.; Sun, N.; Wang, Y.; Lee, L.Y.S.; et al. Recent advance in MXenes: A promising 2D material for catalysis, sensor and chemical adsorption. Coord. Chem. Rev. 2017, 352, 306–327. [Google Scholar] [CrossRef]
  53. Zhan, X.; Si, C.; Zhou, J.; Sun, Z. MXene and MXene-based composites: Synthesis, properties and environment-related applications. Nanoscale Horiz. 2020, 5, 235–258. [Google Scholar] [CrossRef]
  54. Ihsanullah, I. Potential of MXenes in water desalination: Current status and perspectives. Nano-Micro Lett. 2020, 12, 72. [Google Scholar] [CrossRef]
  55. Ihsanullah, I. MXenes (two-dimensional metal carbides) as emerging nanomaterials for water purification: Progress, challenges and prospects. Chem. Eng. J. 2020, 388, 124340. [Google Scholar] [CrossRef]
  56. Anasori, B.; Lukatskaya, M.R.; Gogotsi, Y. 2D metal carbides and nitrides (MXenes) for energy storage. In MXenes; Jenny Stanford Publishing: New Delhi, India, 2023; pp. 677–722. [Google Scholar]
  57. Zhang, Y.-J.; Lan, J.-H.; Wang, L.; Wu, Q.-Y.; Wang, C.-Z.; Bo, T.; Chai, Z.-F.; Shi, W.-Q. Adsorption of uranyl species on hydroxylated titanium carbide nanosheet: A first-principles study. J. Hazard. Mater. 2016, 308, 402–410. [Google Scholar] [CrossRef]
  58. Li, L.; Cheng, Q. MXene based nanocomposite films. In Exploration; Wiley Online Library: Hoboken, NJ, USA, 2022. [Google Scholar]
  59. Lagunas, F.; Kamysbayev, V.; Rui, X.; Filatov, A.; Hu, H.; Klie, R.; Talapin, D. Atomic-Resolution Imaging and Spectroscopy of Functionalized MXene Nanosheets. Microsc. Microanal. 2020, 26 (Suppl. S2), 2328–2330. [Google Scholar] [CrossRef]
  60. Hussain, I.; Arifeen, W.U.; Khan, S.A.; Aftab, S.; Javed, M.S.; Hussain, S.; Ahmad, M.; Chen, X.; Zhao, J.; Rosaiah, P.; et al. M4X3 MXenes: Application in energy storage devices. Nano-Micro Lett. 2024, 16, 215. [Google Scholar] [CrossRef]
  61. Jahan, N.; Hussain, S.; Rahman, H.U.; Manzoor, I.; Pandey, S.; Habib, K.; Ali, S.K.; Pandita, R.; Upadhyay, C. Structural, morphological and elemental analysis of selectively etched and exfoliated Ti3AlC2 MAX phase. J. Multidiscip. Appl. Nat. Sci. 2021, 1, 13–17. [Google Scholar] [CrossRef]
  62. Gogotsi, Y.; Huang, Q. MXenes: Two-dimensional building blocks for future materials and devices. ACS Publ. 2021, 15, 5775–5780. [Google Scholar] [CrossRef]
  63. Bai, S.; Yang, M.; Jiang, J.; He, X.; Zou, J.; Xiong, Z.; Liao, G.; Liu, S. Recent advances of MXenes as electrocatalysts for hydrogen evolution reaction. Npj 2D Mater. Appl. 2021, 5, 78. [Google Scholar] [CrossRef]
  64. Huang, W.; Wang, J.; Lai, W.; Guo, M. MXene Surface Architectonics: Bridging Molecular Design to Multifunctional Applications. Molecules 2025, 30, 1929. [Google Scholar] [CrossRef] [PubMed]
  65. Sukidpaneenid, S.; Chawengkijwanich, C.; Pokhum, C.; Isobe, T.; Opaprakasit, P.; Sreearunothai, P. Multi-function adsorbent-photocatalyst MXene-TiO2 composites for removal of enrofloxacin antibiotic from water. J. Environ. Sci. 2023, 124, 414–428. [Google Scholar] [CrossRef] [PubMed]
  66. Shahzad, A.; Rasool, K.; Nawaz, M.; Miran, W.; Jang, J.; Moztahida, M.; Mahmoud, K.A.; Lee, D.S. Heterostructural TiO2/Ti3C2Tx (MXene) for photocatalytic degradation of antiepileptic drug carbamazepine. Chem. Eng. J. 2018, 349, 748–755. [Google Scholar] [CrossRef]
  67. Shuck, C.E.; Ventura-Martinez, K.; Goad, A.; Uzun, S.; Shekhirev, M.; Gogotsi, Y. Safe synthesis of MAX and MXene: Guidelines to reduce risk during synthesis. ACS Chem. Health Saf. 2021, 28, 326–338. [Google Scholar] [CrossRef]
  68. Hantanasirisakul, K.; Alhabeb, M.; Lipatov, A.; Maleski, K.; Anasori, B.; Salles, P.; Ieosakulrat, C.; Pakawatpanurut, P.; Sinitskii, A.; May, S.J.; et al. Effects of synthesis and processing on optoelectronic properties of titanium carbonitride MXene. Chem. Mater. 2019, 31, 2941–2951. [Google Scholar] [CrossRef]
  69. Peng, T.; Wang, S.; Xu, Z.; Tang, T.; Zhao, Y. Multifunctional MXene/aramid nanofiber composite films for efficient electromagnetic interference shielding and repeatable early fire detection. ACS Omega 2022, 7, 29161–29170. [Google Scholar] [CrossRef]
  70. Sahoo, S.; Wickramathilaka, K.Y.; Njeri, E.; Silva, D.; Suib, S.L. A review on transition metal oxides in catalysis. Front. Chem. 2024, 12, 1374878. [Google Scholar] [CrossRef]
  71. Mozafari, M.; Soroush, M. Surface functionalization of MXenes. Mater. Adv. 2021, 2, 7277–7307. [Google Scholar] [CrossRef]
  72. Lin, H.; Chen, Y.; Shi, J. Insights into 2D MXenes for versatile biomedical applications: Current advances and challenges ahead. Adv. Sci. 2018, 5, 1800518. [Google Scholar] [CrossRef]
  73. Biswas, S.; Alegaonkar, P.S. MXene: Evolutions in chemical synthesis and recent advances in applications. Surfaces 2022, 5, 1–34. [Google Scholar]
  74. Yang, R.; Chen, X.; Ke, W.; Wu, X. Recent research progress in the structure, fabrication, and application of MXene-based heterostructures. Nanomaterials 2022, 12, 1907. [Google Scholar] [CrossRef]
  75. Shariq, M.; Alshehri, K.; Bouzgarrou, S.M.; Ali, S.K.; Alqurashi, Y.; Hassan, K.; Azooz, R. Progress in development of MXene-based nanocomposites for supercapacitor application-A review. FlatChem 2024, 44, 100609. [Google Scholar] [CrossRef]
  76. Aghayar, Z.; Malaki, M.; Zhang, Y. MXene-based ink design for printed applications. Nanomaterials 2022, 12, 4346. [Google Scholar] [CrossRef] [PubMed]
  77. Li, P.; Zhao, X.; Ding, Y.; Chen, L.; Wang, X.; Xie, H. Application studies on MXene-based flexible composites. Front. Therm. Eng. 2024, 4, 1440165. [Google Scholar] [CrossRef]
  78. Wei, Y.; Zhang, P.; Soomro, R.A.; Zhu, Q.; Xu, B. Advances in the synthesis of 2D MXenes. Adv. Mater. 2021, 33, 2103148. [Google Scholar] [CrossRef]
  79. Arole, K.; Blivin, J.W.; Saha, S.; Holta, D.E.; Zhao, X.; Sarmah, A.; Cao, H.; Radovic, M.; Lutkenhaus, J.L.; Green, M.J. Water-dispersible Ti3C2Tz MXene nanosheets by molten salt etching. iScience 2021, 24, 103403. [Google Scholar] [CrossRef]
  80. Wang, J.; He, J.; Kan, D.; Chen, K.; Song, M.; Huo, W. MXene film prepared by vacuum-assisted filtration: Properties and applications. Crystals 2022, 12, 1034. [Google Scholar] [CrossRef]
  81. Hirsch, R.; Ternes, T.; Haberer, K.; Kratz, K.-L. Occurrence of antibiotics in the aquatic environment. Sci. Total Environ. 1999, 225, 109–118. [Google Scholar]
  82. Ji, Y.; Ferronato, C.; Salvador, A.; Yang, X.; Chovelon, J.-M. Degradation of ciprofloxacin and sulfamethoxazole by ferrous-activated persulfate: Implications for remediation of groundwater contaminated by antibiotics. Sci. Total Environ. 2014, 472, 800–808. [Google Scholar]
  83. Wolfson, J.S.; Hooper, D.C. Fluoroquinolone antimicrobial agents. Clin. Microbiol. Rev. 1989, 2, 378–424. [Google Scholar] [CrossRef] [PubMed]
  84. Paul, T.; Dodd, M.C.; Strathmann, T.J. Photolytic and photocatalytic decomposition of aqueous ciprofloxacin: Transformation products and residual antibacterial activity. Water Res. 2010, 44, 3121–3132. [Google Scholar] [CrossRef] [PubMed]
  85. Jiang, C.; Ji, Y.; Shi, Y.; Chen, J.; Cai, T. Sulfate radical-based oxidation of fluoroquinolone antibiotics: Kinetics, mechanisms and effects of natural water matrices. Water Res. 2016, 106, 507–517. [Google Scholar] [CrossRef] [PubMed]
  86. Golet, E.M.; Xifra, I.; Siegrist, H.; Alder, A.C.; Giger, W. Environmental exposure assessment of fluoroquinolone antibacterial agents from sewage to soil. Environ. Sci. Technol. 2003, 37, 3243–3249. [Google Scholar] [CrossRef] [PubMed]
  87. Watkinson, A.; Murby, E.; Costanzo, S. Removal of antibiotics in conventional and advanced wastewater treatment: Implications for environmental discharge and wastewater recycling. Water Res. 2007, 41, 4164–4176. [Google Scholar] [CrossRef]
  88. Kong, D.; Liang, B.; Yun, H.; Cheng, H.; Ma, J.; Cui, M.; Wang, A.; Ren, N. Cathodic degradation of antibiotics: Characterization and pathway analysis. Water Res. 2015, 72, 281–292. [Google Scholar] [CrossRef]
  89. Liu, C.; Nanaboina, V.; Korshin, G.V.; Jiang, W. Spectroscopic study of degradation products of ciprofloxacin, norfloxacin and lomefloxacin formed in ozonated wastewater. Water Res. 2012, 46, 5235–5246. [Google Scholar] [CrossRef]
  90. Alexandrino, D.A.; Mucha, A.P.; Almeida, C.M.R.; Gao, W.; Jia, Z.; Carvalho, M.F. Biodegradation of the veterinary antibiotics enrofloxacin and ceftiofur and associated microbial community dynamics. Sci. Total Environ. 2017, 581, 359–368. [Google Scholar] [CrossRef]
  91. Barhoumi, N.; Labiadh, L.; Oturan, M.A.; Oturan, N.; Gadri, A.; Ammar, S.; Brillas, E. Electrochemical mineralization of the antibiotic levofloxacin by electro-Fenton-pyrite process. Chemosphere 2015, 141, 250–257. [Google Scholar] [CrossRef]
  92. Sayed, M.; Ismail, M.; Khan, S.; Tabassum, S.; Khan, H.M. Degradation of ciprofloxacin in water by advanced oxidation process: Kinetics study, influencing parameters and degradation pathways. Environ. Technol. 2016, 37, 590–602. [Google Scholar] [CrossRef]
  93. Chen, P.; Wang, F.; Chen, Z.-F.; Zhang, Q.; Su, Y.; Shen, L.; Yao, K.; Liu, Y.; Cai, Z.; Lv, W.; et al. Study on the photocatalytic mechanism and detoxicity of gemfibrozil by a sunlight-driven TiO2/carbon dots photocatalyst: The significant roles of reactive oxygen species. Appl. Catal. B Environ. 2017, 204, 250–259. [Google Scholar] [CrossRef]
  94. Durán-Álvarez, J.C.; Avella, E.; Ramírez-Zamora, R.M.; Zanella, R. Photocatalytic degradation of ciprofloxacin using mono-(Au, Ag and Cu) and bi-(Au–Ag and Au–Cu) metallic nanoparticles supported on TiO2 under UV-C and simulated sunlight. Catal. Today 2016, 266, 175–187. [Google Scholar] [CrossRef]
  95. Cai, X.; He, J.; Chen, L.; Chen, K.; Li, Y.; Zhang, K.; Jin, Z.; Liu, J.; Wang, C.; Wang, X.; et al. A 2D-g-C3N4 nanosheet as an eco-friendly adsorbent for various environmental pollutants in water. Chemosphere 2017, 171, 192–201. [Google Scholar] [CrossRef] [PubMed]
  96. Kim, S.D.; Cho, J.; Kim, I.S.; Vanderford, B.J.; Snyder, S.A. Occurrence and removal of pharmaceuticals and endocrine disruptors in South Korean surface, drinking, and waste waters. Water Res. 2007, 41, 1013–1021. [Google Scholar] [CrossRef] [PubMed]
  97. Botero-Coy, A.M.; Martínez-Pachón, D.; Boix, C.; Rincón, R.J.; Castillo, N.; Arias-Marín, L.; Manrique-Losada, L.; Torres-Palma, R.; Moncayo-Lasso, A.; Hernandez, F. An investigation into the occurrence and removal of pharmaceuticals in Colombian wastewater. Sci. Total Environ. 2018, 642, 842–853. [Google Scholar] [CrossRef]
  98. Mestre, A.S.; Carvalho, A.P. Photocatalytic degradation of pharmaceuticals carbamazepine, diclofenac, and sulfamethoxazole by semiconductor and carbon materials: A review. Molecules 2019, 24, 3702. [Google Scholar] [CrossRef]
  99. Yi, X.; Yuan, J.; Tang, H.; Du, Y.; Hassan, B.; Yin, K.; Chen, Y.; Liu, X. Embedding few-layer Ti3C2Tx into alkalized g-C3N4 nanosheets for efficient photocatalytic degradation. J. Colloid Interface Sci. 2020, 571, 297–306. [Google Scholar] [CrossRef]
  100. 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]
  101. 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]
  102. Cao, Y.; Fang, Y.; Lei, X.; Tan, B.; Hu, X.; Liu, B.; Chen, Q. Fabrication of novel CuFe2O4/MXene hierarchical heterostructures for enhanced photocatalytic degradation of sulfonamides under visible light. J. Hazard. Mater. 2020, 387, 122021. [Google Scholar] [CrossRef]
  103. Cui, C.; Guo, R.; Xiao, H.; Ren, E.; Song, Q.; Xiang, C.; Lai, X.; Lan, J.; Jiang, S. Bi2WO6/Nb2CTx MXene hybrid nanosheets with enhanced visible-light-driven photocatalytic activity for organic pollutants degradation. Appl. Surf. Sci. 2020, 505, 144595. [Google Scholar] [CrossRef]
  104. Kuang, P.; Low, J.; Cheng, B.; Yu, J.; Fan, J. MXene-based photocatalysts. J. Mater. Sci. Technol. 2020, 56, 18–44. [Google Scholar] [CrossRef]
  105. Barsoum, M.W.; El-Raghy, T. The MAX phases: Unique new carbide and nitride materials: Ternary ceramics turn out to be surprisingly soft and machinable, yet also heat-tolerant, strong and lightweight. Am. Sci. 2001, 89, 334–343. [Google Scholar] [CrossRef]
  106. Alhabeb, M.; Maleski, K.; Anasori, B.; Lelyukh, P.; Clark, L.; Sin, S.; Gogotsi, Y. Guidelines for synthesis and processing of two-dimensional titanium carbide (Ti3C2T x MXene). Chem. Mater. 2017, 29, 7633–7644. [Google Scholar] [CrossRef]
  107. Nawaz, M.; Miran, W.; Jang, J.; Lee, D.S. One-step hydrothermal synthesis of porous 3D reduced graphene oxide/TiO2 aerogel for carbamazepine photodegradation in aqueous solution. Appl. Catal. B Environ. 2017, 203, 85–95. [Google Scholar] [CrossRef]
  108. Xu, J.; Li, L.; Guo, C.; Zhang, Y.; Meng, W. Photocatalytic degradation of carbamazepine by tailored BiPO4: Efficiency, intermediates and pathway. Appl. Catal. B Environ. 2013, 130, 285–292. [Google Scholar] [CrossRef]
  109. Wang, F.; Feng, Y.; Chen, P.; Wang, Y.; Su, Y.; Zhang, Q.; Zeng, Y.; Xie, Z.; Liu, H.; Liu, Y.; et al. Photocatalytic degradation of fluoroquinolone antibiotics using ordered mesoporous g-C3N4 under simulated sunlight irradiation: Kinetics, mechanism, and antibacterial activity elimination. Appl. Catal. B Environ. 2018, 227, 114–122. [Google Scholar] [CrossRef]
  110. Ahmed, S.; Rasul, M.; Brown, R.; Hashib, M. Influence of parameters on the heterogeneous photocatalytic degradation of pesticides and phenolic contaminants in wastewater: A short review. J. Environ. Manag. 2011, 92, 311–330. [Google Scholar] [CrossRef]
  111. Xu, L.; Wang, G.; Ma, F.; Zhao, Y.; Lu, N.; Guo, Y.; Yang, X. Photocatalytic degradation of an aqueous sulfamethoxazole over the metallic silver and Keggin unit codoped titania nanocomposites. Appl. Surf. Sci. 2012, 258, 7039–7046. [Google Scholar] [CrossRef]
  112. Xu, D.; Ma, Y.; Wang, J.; Chen, W.; Tang, Y.; Li, X.; Li, L. Interfacial engineering of 2D/2D MXene heterostructures: Face-to-face contact for augmented photodegradation of amoxicillin. Chem. Eng. J. 2021, 426, 131246. [Google Scholar] [CrossRef]
  113. 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]
  114. Fan, Y.; Yuan, Z.; Zou, G.; Zhang, Q.; Liu, B.; Peng, Q. Two-dimensional MXene/A-TiO2 composite with unprecedented catalytic activation for sodium alanate. Catal. Today 2018, 318, 167–174. [Google Scholar] [CrossRef]
  115. Jun, B.-M.; Jang, M.; Park, C.M.; Han, J.; Yoon, Y. Selective adsorption of Cs+ by MXene (Ti3C2Tx) from model low-level radioactive wastewater. Nucl. Eng. Technol. 2020, 52, 1201–1207. [Google Scholar] [CrossRef]
  116. Ezelarab, H.A.; Abbas, S.H.; Hassan, H.A.; Abuo-Rahma, G.E.D.A. Recent updates of fluoroquinolones as antibacterial agents. Arch. Pharm. 2018, 351, 1800141. [Google Scholar] [CrossRef] [PubMed]
  117. Janecko, N.; Pokludova, L.; Blahova, J.; Svobodova, Z.; Literak, I. Implications of fluoroquinolone contamination for the aquatic environment—A review. Environ. Toxicol. Chem. 2016, 35, 2647–2656. [Google Scholar] [CrossRef] [PubMed]
  118. Pretali, L.; Fasani, E.; Sturini, M. Current advances on the photocatalytic degradation of fluoroquinolones: Photoreaction mechanism and environmental application. Photochem. Photobiol. Sci. 2022, 21, 899–912. [Google Scholar] [CrossRef]
  119. Sayed, M.; Shah, L.A.; Khan, J.A.; Shah, N.S.; Nisar, J.; Khan, H.M.; Zhang, P.; Khan, A.R. Efficient photocatalytic degradation of norfloxacin in aqueous media by hydrothermally synthesized immobilized TiO2/Ti films with exposed {001} facets. J. Phys. Chem. A 2016, 120, 9916–9931. [Google Scholar] [CrossRef]
  120. Tang, J.; Wang, R.; Liu, M.; Zhang, Z.; Song, Y.; Xue, S.; Zhao, Z.; Dionysiou, D.D. Construction of novel Z-scheme Ag/FeTiO3/Ag/BiFeO3 photocatalyst with enhanced visible-light-driven photocatalytic performance for degradation of norfloxacin. Chem. Eng. J. 2018, 351, 1056–1066. [Google Scholar] [CrossRef]
  121. Lv, X.; Yan, D.Y.; Lam, F.L.-Y.; Ng, Y.H.; Yin, S.; An, A.K. Solvothermal synthesis of copper-doped BiOBr microflowers with enhanced adsorption and visible-light driven photocatalytic degradation of norfloxacin. Chem. Eng. J. 2020, 401, 126012. [Google Scholar] [CrossRef]
  122. Tang, L.; Wang, J.; Zeng, G.; Liu, Y.; Deng, Y.; Zhou, Y.; Tang, J.; Wang, J.; Guo, Z. Enhanced photocatalytic degradation of norfloxacin in aqueous Bi2WO6 dispersions containing nonionic surfactant under visible light irradiation. J. Hazard. Mater. 2016, 306, 295–304. [Google Scholar] [CrossRef]
  123. Zhang, H.; Zheng, Y.; Zhou, H.; Zhu, S.; Yang, J. Nanocellulose-intercalated MXene NF membrane with enhanced swelling resistance for highly efficient antibiotics separation. Sep. Purif. Technol. 2023, 305, 122425. [Google Scholar] [CrossRef]
  124. Li, Z.K.; Wei, Y.; Gao, X.; Ding, L.; Lu, Z.; Deng, J.; Yang, X.; Caro, J.; Wang, H. Antibiotics separation with MXene membranes based on regularly stacked high-aspect-ratio nanosheets. Angew. Chem. Int. Ed. 2020, 59, 9751–9756. [Google Scholar] [CrossRef] [PubMed]
  125. Lu, S.; Meng, G.; Wang, C.; Chen, H. Photocatalytic inactivation of airborne bacteria in a polyurethane foam reactor loaded with a hybrid of MXene and anatase TiO2 exposing {0 0 1} facets. Chem. Eng. J. 2021, 404, 126526. [Google Scholar] [CrossRef] [PubMed]
  126. Zeng, G.; He, Z.; Wan, T.; Wang, T.; Yang, Z.; Liu, Y.; Lin, Q.; Wang, Y.; Sengupta, A.; Pu, S. A self-cleaning photocatalytic composite membrane based on g-C3N4@ MXene nanosheets for the removal of dyes and antibiotics from wastewater. Sep. Purif. Technol. 2022, 292, 121037. [Google Scholar] [CrossRef]
  127. Li, C.; Sun, Z.; Zhang, W.; Yu, C.; Zheng, S. Highly efficient g-C3N4/TiO2/kaolinite composite with novel three-dimensional structure and enhanced visible light responding ability towards ciprofloxacin and S. aureus. Appl. Catal. B Environ. 2018, 220, 272–282. [Google Scholar] [CrossRef]
  128. 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]
  129. Chuaicham, C.; Pawar, R.R.; Karthikeyan, S.; Ohtani, B.; Sasaki, K. Fabrication and characterization of ternary sepiolite/g-C3N4/Pd composites for improvement of photocatalytic degradation of ciprofloxacin under visible light irradiation. J. Colloid Interface Sci. 2020, 577, 397–405. [Google Scholar] [CrossRef]
  130. Machín, A.; Fontánez, K.; Duconge, J.; Cotto, M.C.; Petrescu, F.I.; Morant, C.; Márquez, F. Photocatalytic degradation of fluoroquinolone antibiotics in solution by Au@ ZnO-rGO-gC3N4 composites. Catalysts 2022, 12, 166. [Google Scholar] [CrossRef]
  131. Madhushree, R.; Kr, S.D. Structural investigation of Cr2CTx/NiFe2O4 MXene composite as a bifunctional electrocatalyst for water splitting. Surf. Interfaces 2024, 52, 104849. [Google Scholar]
  132. Sajid, I.H.; Iqbal, M.Z.; Rizwan, S. Recent advances in the role of MXene based hybrid architectures as electrocatalysts for water splitting. RSC Adv. 2024, 14, 6823–6847. [Google Scholar] [CrossRef]
  133. Chatenet, M.; Pollet, B.G.; Dekel, D.R.; Dionigi, F.; Deseure, J.; Millet, P.; Braatz, R.D.; Bazant, M.Z.; Eikerling, M.; Staffell, I.; et al. Water electrolysis: From textbook knowledge to the latest scientific strategies and industrial developments. Chem. Soc. Rev. 2022, 51, 4583–4762. [Google Scholar] [CrossRef]
  134. Lasia, A. Mechanism and kinetics of the hydrogen evolution reaction. Int. J. Hydrogen Energy 2019, 44, 19484–19518. [Google Scholar] [CrossRef]
  135. Wang, J.; Xu, F.; Jin, H.; Chen, Y.; Wang, Y. Non-noble metal-based carbon composites in hydrogen evolution reaction: Fundamentals to applications. Adv. Mater. 2017, 29, 1605838. [Google Scholar] [CrossRef] [PubMed]
  136. Lv, Z.; Liu, D.; Tian, W.; Dang, J. Designed synthesis of WC-based nanocomposites as low-cost, efficient and stable electrocatalysts for the hydrogen evolution reaction. CrystEngComm 2020, 22, 4580–4590. [Google Scholar] [CrossRef]
  137. Lakshmi, K.S.; Vedhanarayanan, B.; Lin, T.-W. Electrocatalytic hydrogen and oxygen evolution reactions: Role of two-dimensional layered materials and their composites. Electrochim. Acta 2023, 447, 142119. [Google Scholar] [CrossRef]
  138. Wei, Y.; Soomro, R.A.; Xie, X.; Xu, B. Design of efficient electrocatalysts for hydrogen evolution reaction based on 2D MXenes. J. Energy Chem. 2021, 55, 244–255. [Google Scholar] [CrossRef]
  139. Zhai, Y.; Ren, X.; Yan, J.; Liu, S. High density and unit activity integrated in amorphous catalysts for electrochemical water splitting. Small Struct. 2021, 2, 2000096. [Google Scholar] [CrossRef]
  140. Yuan, W.; Cheng, L.; An, Y.; Wu, H.; Yao, N.; Fan, X.; Guo, X. MXene nanofibers as highly active catalysts for hydrogen evolution reaction. ACS Sustain. Chem. Eng. 2018, 6, 8976–8982. [Google Scholar] [CrossRef]
  141. Huang, J.-J.; Liu, X.-Q.; Meng, F.-F.; He, L.-Q.; Wang, J.-X.; Wu, J.-C.; Lu, X.-H.; Tong, Y.-X.; Fang, P.-P. A facile method to produce MoSe2/MXene hybrid nanoflowers with enhanced electrocatalytic activity for hydrogen evolution. J. Electroanal. Chem. 2020, 856, 113727. [Google Scholar] [CrossRef]
  142. Reghunath, B.S.; KR, S.D.; Rajasekaran, S.; Saravanakumar, B.; William, J.J.; Pinheiro, D. Hierarchical BiFeO3/Cr2CTx MXene composite as a multifunctional catalyst for hydrogen evolution reaction and as an electrode material for energy storage devices. Electrochim. Acta 2023, 461, 142685. [Google Scholar] [CrossRef]
  143. Seh, Z.W.; Fredrickson, K.D.; Anasori, B.; Kibsgaard, J.; Strickler, A.L.; Lukatskaya, M.R.; Gogotsi, Y.; Jaramillo, T.F.; Vojvodic, A. Two-dimensional molybdenum carbide (MXene) as an efficient electrocatalyst for hydrogen evolution. ACS Energy Lett. 2016, 1, 589–594. [Google Scholar] [CrossRef]
  144. Intikhab, S.; Natu, V.; Li, J.; Li, Y.; Tao, Q.; Rosen, J.; Barsoum, M.W.; Snyder, J. Stoichiometry and surface structure dependence of hydrogen evolution reaction activity and stability of MoxC MXenes. J. Catal. 2019, 371, 325–332. [Google Scholar] [CrossRef]
  145. Thalji, M.R.; Mahmoudi, F.; Bachas, L.G.; Park, C. MXene-based electrocatalysts for water splitting: Material design surface modulation, and catalytic performance. Int. J. Mol. Sci. 2025, 26, 8019. [Google Scholar] [CrossRef]
  146. Raj, K.; Jhala, R.; Sujai, S.; Pattanayak, B.; Singh, J.; Kumar, P.; Kaladhar, M.; Singh, R.; Bisht, Y.; Gautam, A. A comprehensive review on the MXene-based catalysts for hydrogen generation via water splitting process. J. Mol. 2025, 1349, 143534. [Google Scholar]
  147. Talas, S.A.; Kolubah, P.D.; Khairova, R.; Alqahtani, M.; El-Hout, S.; Alissa, F.M.; El-Demellawi, J.K.; Castaño, P.; Mohamed, H.O. MXene-based electrocatalysis for CO2 reduction: Advances, challenges, and perspectives. Mater. Horiz. 2025, 1–35. [Google Scholar] [CrossRef]
  148. Cheng, Y.-W.; Dai, J.-H.; Zhang, Y.-M.; Song, Y. Transition metal modification and carbon vacancy promoted Cr2CO2 (MXenes): A new opportunity for a highly active catalyst for the hydrogen evolution reaction. J. Mater. Chem. A 2018, 6, 20956–20965. [Google Scholar] [CrossRef]
  149. Djire, A.; Zhang, H.; Liu, J.; Miller, E.M.; Neale, N.R. Electrocatalytic and optoelectronic characteristics of the two-dimensional titanium nitride Ti4N3Tx MXene. ACS Appl. Mater. Interfaces 2019, 11, 11812–11823. [Google Scholar] [CrossRef]
  150. Ma, S.; Xu, Z.; Jia, Z.; Chen, L.; Zhu, H.; Chen, Y.; Guo, X.; Du, M. Facile fabrication of carbon fiber skeleton structure of MoS2 supported on 2D MXene composite with highly efficient and stable hydrogen evolution reaction. Compos. Sci. Technol. 2022, 222, 109380. [Google Scholar] [CrossRef]
  151. Li, S.; Tuo, P.; Xie, J.; Zhang, X.; Xu, J.; Bao, J.; Pan, B.; Xie, Y. Ultrathin MXene nanosheets with rich fluorine termination groups realizing efficient electrocatalytic hydrogen evolution. Nano Energy 2018, 47, 512–518. [Google Scholar] [CrossRef]
  152. Zhang, P.; Wang, R.; Xiao, T.; Chang, Z.; Fang, Z.; Zhu, Z.; Xu, C.; Wang, L.; Cheng, J. The High-Performance Bifunctional Catalyst Pd/Ti3C2Tx–Carbon Nanotube for Oxygen Reduction Reaction and Hydrogen Evolution Reaction in Alkaline Medium. Energy Technol. 2020, 8, 2000306. [Google Scholar] [CrossRef]
  153. Wang, X.; Ding, J.; Song, W.; Yang, X.; Zhang, T.; Huang, Z.; Wang, H.; Han, X.; Hu, W. Cation vacancy clusters in Ti3C2Tx MXene induce ultra-strong interaction with noble metal clusters for efficient electrocatalytic hydrogen evolution. Adv. Energy Mater. 2023, 13, 2300148. [Google Scholar] [CrossRef]
  154. Wu, Y.; Wei, W.; Yu, R.; Xia, L.; Hong, X.; Zhu, J.; Li, J.; Lv, L.; Chen, W.; Zhao, Y.; et al. Anchoring sub-nanometer Pt clusters on crumpled paper-like MXene enables high hydrogen evolution mass activity. Adv. Funct. Mater. 2022, 32, 2110910. [Google Scholar] [CrossRef]
  155. Jiang, J.; Li, F.; Bai, S.; Wang, Y.; Xiang, K.; Wang, H.; Zou, J.; Hsu, J.-P. Carbonitride MXene Ti3CN (OH) x@ MoS2 hybrids as efficient electrocatalyst for enhanced hydrogen evolution. Nano Res. 2023, 16, 4656–4663. [Google Scholar] [CrossRef]
  156. Yan, L.; Song, D.; Liang, J.; Li, X.; Li, H.; Liu, Q. Fabrication of highly efficient Rh-doped cobalt–nickel-layered double hydroxide/MXene-based electrocatalyst with rich oxygen vacancies for hydrogen evolution. J. Colloid Interface Sci. 2023, 640, 338–347. [Google Scholar] [CrossRef]
  157. Zhang, L.; Zhao, X.; Chu, Z.; Wang, Q.; Cao, Y.; Li, J.; Lei, W.; Cao, J.; Si, W. Construction of Co-decorated 3D nitrogen doped-carbon nanotube/Ti3C2Tx-MXene as efficient hydrogen evolution electrocatalyst. Int. J. Hydrogen Energy 2023, 48, 15053–15064. [Google Scholar] [CrossRef]
  158. Hussain, S.; Vikraman, D.; Sheikh, Z.A.; Mehran, M.T.; Shahzad, F.; Batoo, K.M.; Kim, H.-S.; Kim, D.-K.; Ali, M.; Jung, J. WS2-embedded MXene/GO hybrid nanosheets as electrodes for asymmetric supercapacitors and hydrogen evolution reactions. Chem. Eng. J. 2023, 452, 139523. [Google Scholar] [CrossRef]
  159. Luo, R.; Li, R.; Jiang, C.; Qi, R.; Liu, M.; Luo, C.; Lin, H.; Huang, R.; Peng, H. Facile synthesis of cobalt modified 2D titanium carbide with enhanced hydrogen evolution performance in alkaline media. Int. J. Hydrogen Energy 2021, 46, 32536–32545. [Google Scholar] [CrossRef]
  160. Doyle, R.L.; Godwin, I.J.; Brandon, M.P.; Lyons, M.E. Redox and electrochemical water splitting catalytic properties of hydrated metal oxide modified electrodes. Phys. Chem. Chem. Phys. 2013, 15, 13737–13783. [Google Scholar] [CrossRef]
  161. Xu, H.; Yuan, J.; He, G.; Chen, H. Current and future trends for spinel-type electrocatalysts in electrocatalytic oxygen evolution reaction. Coord. Chem. Rev. 2023, 475, 214869. [Google Scholar] [CrossRef]
  162. Xie, Y.; Yu, H.; Deng, L.; Amin, R.; Yu, D.; Fetohi, A.E.; Maximov, M.Y.; Li, L.; El-Khatib, K.; Peng, S. Anchoring stable FeS2 nanoparticles on MXene nanosheets via interface engineering for efficient water splitting. Inorg. Chem. Front. 2022, 9, 662–669. [Google Scholar] [CrossRef]
  163. Han, S.; Chen, Y.; Hao, Y.; Xie, Y.; Xie, D.; Chen, Y.; Xiong, Y.; He, Z.; Hu, F.; Li, L.; et al. Multi-dimensional hierarchical CoS2@ MXene as trifunctional electrocatalysts for zinc-air batteries and overall water splitting. Development 2020, 17, 18. [Google Scholar] [CrossRef]
  164. Li, Y.; Du, Q.-X.; Cui, J.; Chen, X.; Yang, H.-W.; Qian, H. Identifying the intrinsic active site in bimetallic Co3S4/Ni3S2 feathers on MXene nanosheets as a heterostructure for efficient oxygen evolution reaction. Inorg. Chem. Front. 2023, 10, 184–191. [Google Scholar] [CrossRef]
  165. Zou, H.; He, B.; Kuang, P.; Yu, J.; Fan, K. Metal–organic framework-derived nickel–cobalt sulfide on ultrathin mxene nanosheets for electrocatalytic oxygen evolution. ACS Appl. Mater. Interfaces 2018, 10, 22311–22319. [Google Scholar] [CrossRef] [PubMed]
  166. Zhu, X.; Zhu, T.; Chen, Q.; Peng, W.; Li, Y.; Zhang, F.; Fan, X. FeP-CoP Nanocubes In Situ Grown on Ti3C2T x MXene as Efficient Electrocatalysts for the Oxygen Evolution Reaction. Ind. Eng. Chem. Res. 2022, 61, 10837–10845. [Google Scholar] [CrossRef]
  167. Li, N.; Han, J.; Yao, K.; Han, M.; Wang, Z.; Liu, Y.; Liu, L.; Liang, H. Synergistic phosphorized NiFeCo and MXene interaction inspired the formation of high-valence metal sites for efficient oxygen evolution. J. Mater. Sci. Technol. 2022, 106, 90–97. [Google Scholar] [CrossRef]
  168. Le, T.A.; Tran, N.Q.; Hong, Y.; Kim, M.; Lee, H. Porosity-engineering of MXene as a support material for a highly efficient electrocatalyst toward overall water splitting. ChemSusChem 2020, 13, 945–955. [Google Scholar] [CrossRef]
  169. Zhang, B.; Shan, J.; Wang, X.; Hu, Y.; Li, Y. Ru/Rh cation doping and oxygen-vacancy engineering of FeOOH nanoarrays@ Ti3C2Tx MXene heterojunction for highly efficient and stable electrocatalytic oxygen evolution. Small 2022, 18, 2200173. [Google Scholar] [CrossRef]
  170. Yan, F.; Ding, J.; Hu, L.; Li, S.; Zhang, S.; Wang, M.; He, L.; Du, M. Solution Plasma-Assisted Multivariate Metal Nanoalloys Encapsulated with Carbon Dots for Efficient Oxygen Evolution Reaction. ChemCatChem 2023, 15, e202300115. [Google Scholar] [CrossRef]
  171. Zhang, Y. Application of MXene-Based Composites for Hydrogen Production by Water Electrolysis. Ph.D. Thesis, Université de Lille, Lille, France, 2024. [Google Scholar]
  172. Li, S.; Fan, J.; Xiao, G.; Gao, S.; Cui, K.; Chao, Z. Multifunctional Co3O4/Ti3C2Tx MXene nanocomposites for integrated all solid-state asymmetric supercapacitors and energy-saving electrochemical systems of H2 production by urea and alcohols electrolysis. Int. J. Hydrogen Energy 2022, 47, 22663–22679. [Google Scholar] [CrossRef]
  173. Chen, Y.; Yao, H.; Kong, F.; Tian, H.; Meng, G.; Wang, S.; Mao, X.; Cui, X.; Hou, X.; Shi, J. V2C MXene synergistically coupling FeNi LDH nanosheets for boosting oxygen evolution reaction. Appl. Catal. B Environ. 2021, 297, 120474. [Google Scholar] [CrossRef]
  174. Hao, C.; Wu, Y.; An, Y.; Cui, B.; Lin, J.; Li, X.; Wang, D.; Jiang, M.; Cheng, Z.; Hu, S. Interface-coupling of CoFe-LDH on MXene as high-performance oxygen evolution catalyst. Mater. Today Energy 2019, 12, 453–462. [Google Scholar] [CrossRef]
  175. Li, X.; Zhang, Z.; Xiang, Q.; Chen, R.; Wu, D.; Li, G.; Wang, L. A three-dimensional flower-like NiCo-layered double hydroxide grown on nickel foam with an MXene coating for enhanced oxygen evolution reaction electrocatalysis. RSC Adv. 2021, 11, 12392–12397. [Google Scholar] [CrossRef]
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

Irfan, S.; Khan, S.B.; Lardhi, S.; AlFaify, S. Recent Advances in MXene-Based Composites for Their Efficiency in the Degradation of Antibiotics and Water Splitting. Molecules 2025, 30, 3712. https://doi.org/10.3390/molecules30183712

AMA Style

Irfan S, Khan SB, Lardhi S, AlFaify S. Recent Advances in MXene-Based Composites for Their Efficiency in the Degradation of Antibiotics and Water Splitting. Molecules. 2025; 30(18):3712. https://doi.org/10.3390/molecules30183712

Chicago/Turabian Style

Irfan, Syed, Sadaf Bashir Khan, Sheikha Lardhi, and S. AlFaify. 2025. "Recent Advances in MXene-Based Composites for Their Efficiency in the Degradation of Antibiotics and Water Splitting" Molecules 30, no. 18: 3712. https://doi.org/10.3390/molecules30183712

APA Style

Irfan, S., Khan, S. B., Lardhi, S., & AlFaify, S. (2025). Recent Advances in MXene-Based Composites for Their Efficiency in the Degradation of Antibiotics and Water Splitting. Molecules, 30(18), 3712. https://doi.org/10.3390/molecules30183712

Article Metrics

Back to TopTop