Next Article in Journal
Development and Validation of a Gas Chromatography-Mass Spectrometry Method for Determining Acaricides in Bee Pollen
Next Article in Special Issue
Methylene Blue Dye Adsorption on Iron Oxide-Hydrochar Composite Synthesized via a Facile Microwave-Assisted Hydrothermal Carbonization of Pomegranate Peels’ Waste
Previous Article in Journal
Improved Phenolic Profile, Sensory Acceptability, and Storage Stability of Strawberry Decoction Beverages Added with Blueberry Decoctions
Previous Article in Special Issue
“Green” nZVI-Biochar as Fenton Catalyst: Perspective of Closing-the-Loop in Wastewater Treatment
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Retrospective on Exploring MXene-Based Nanomaterials: Photocatalytic Applications

1
School of Materials Science and Engineering, Dongguan University of Technology, Dongguan 523808, China
2
Yangtze Delta Region Institute (Huzhou), University of Electronic Science and Technology of China, Huzhou 313000, China
3
Dongguan Institute of Science and Technology Innovation, Dongguan University of Technology, Dongguan 523808, China
4
School of Materials Science and Engineering, Tianjin University, Tianjin 300350, China
5
School of Materials and Energy, Southwest University, Chongqing 400715, China
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(6), 2495; https://doi.org/10.3390/molecules28062495
Submission received: 15 February 2023 / Revised: 27 February 2023 / Accepted: 2 March 2023 / Published: 9 March 2023

Abstract

:
Nanostructural two-dimensional compounds are grabbing the attention of researchers all around the world. This research is progressing quickly due to its wide range of applications in numerous industries and enormous promise for future technological breakthroughs. Growing environmental consciousness has made it vital to treat wastewater and avoid releasing hazardous substances into the environment. Rising consumer expectations have led to the emergence of new, frequently nonbiodegradable compounds. Due to their specific chemical and physical properties, MXenes have recently been identified as promising candidates. MXenes are regarded as a prospective route for environmental remediation technologies, such as photocatalysis, adsorption, and membrane separation, and as electrocatalytic sensors for pollution recognition because of their high hydrophilicity, inherent chemical nature, and robust electrochemistry. The development of catalysts based on MXene materials for the photocatalytic breakdown of pharmaceutical wastes in polluted water is critically evaluated in this study. With an emphasis on the degradation mechanism, the photocatalytic degradation of antibiotics using MXenes and MXene-based nanocomposites is explained in depth. We emphasize the significant difficulties in producing MXenes and their composites, as well as in the degradation of drugs. The successful use of MXenes in water filtration and suggestions for future study are also presented.

1. Introduction

A large amount of post-production wastewater enters natural streams every day globally, which has caused contamination to a crucial level [1]. Rapid global industrial development is directly linked to a growing world population, which generates a large amount of wastewater production, mainly from the chemical, pharmaceutical, and cosmetic industries. Different washing process lines produce industrial wastes with disinfectants and surfactants. The wastewater coming from industries may have many inorganic contaminants. Inorganic contaminants may contain toxic heavy metals, e.g., cobalt, mercury, arsenic, lead, and poisonous gases including NOx, Sox, Cox, and NH3. Surface-factionalized MXenes are thought to provide a way around this problem due to their outstanding electrical characteristics, which can increase charge separation effectiveness [2]. In a previous study, rhodamine B molecules exhibited exceptional adsorption and catalytic photo-degradation of Ti3C2 MXene-Co3O4 nanocomposites, which were created using an in situ solvothermal technique [3].
The RhB and MB dye removal rates stabilized at around 4 and 1.5 h, respectively, demonstrating the produced complex’s efficiency as a dye adsorbent. The enlarged surface area was beneficial for dye adsorption for photocatalytic needs, which could cause these characteristics. These nanocomposites demonstrated excellent photo-degradation and absorption capabilities and strong stability and reusability, as shown by an outstanding catalytic feature after eight cycle runs [3]. In another work, using an electrostatically driven method, Fang and his colleagues successfully synthesized Ag2WO4/Ti3C2 Schottky-junction photocatalyst for the first time. The Ag2WO4/Ti3C2 demonstrated an enhanced photocatalytic elimination rate for sulfadimidine (88.6%) and tetracycline hydrochloride (62.9%). The creation of the Schottky heterojunction was said to be responsible for the photocatalytic performance [4]. When 2D Bi2MoO6 was hydrothermally produced on a surface layer of nanosheets of Ti3C2 to create Bi2MoO6/Ti3C2 MXene, a more rapid photo-degradation was observed [5]. The pure Bi2MoO6 degraded only 37% of tetracycline, while the improved BT-30 photocatalyst demonstrated the highest degradation performance of tetracycline, at about 99%, within only 30 min of light irradiation. The photo-degradation of tetracycline increased with increasing Ti3C2 MXene content, as shown by the observation that the rate of reaction of BT-30 was reduced. When BT-30 degraded TC (k 0.143 min), it did so at a response rate that was 8.8 times quicker than that of pure Bi2MoO6. The improved catalyst was further examined for Cr(VI) reduction, and after 60 min, it demonstrated a 99% reduction. The different MXene applications are described in Figure 1.
In addition to a thorough analysis of recent developments in the use of MXene-based photocatalysts for the degradation of pharmaceuticals in water, this review offers a general synthesis of the existing literature. Even though there have been several evaluations of MXenes’ ability to remove contaminants by adsorptive means, this review is focused primarily on the photocatalytic removal of antibiotics in pharmaceutical wastes. The methods used to synthesize different MXene-based catalysts, the influence of key factors on the decomposition procedure, the mechanism of decomposition, and the regeneration capacity of used catalysts are thoroughly discussed.

2. Introduction of Mxenes

Mxenes were obtained for the first time in 2011; they are a novel class of 2D transition metal nitrides, carbides, and carbonitrides [7,8,9,10,11]. Mxenes’ structures are composed of Mn+1Xn mutilated sheets having the formula of Mn+1XnTx, where n (varies from one to four), M, X, and Tx symbolize the number of transition metal layers exist in the structure of MXenes, the transition metal, carbon or nitrogen, and surface terminations, respectively [12,13,14,15,16,17,18]. MXenes are generally fabricated from MAX phase (Mn+1AXn), where A layer is typically inserted within octahedral Mn+1Xn, with a strong M/X and a comparatively weak M/A bond, as shown in Figure 2 [19,20,21,22,23]. Thus far, nearly 30 stoichiometric structures have been discovered in the MXene family, and wide-ranging exploration is in progress to grow new structures [12,24,25,26,27]. As a result of their excellent chemical and mechanical properties, Ti-based MXenes are primarily used for applications in numerous fields because of the ease of modifying their physical and chemical properties with varying MAX-phase precursors [12]. The oxygen-rich terminal groups on MXene’s surface offer active sites for chemical covalent modification, which causes the engineering of novel structures by combining suitable functional groups, therefore enhancing degradation efficiencies for the removal of organic contaminants from industrial wastewater.
A new MXene is composed of titanium carbide, which is fabricated by optimized etching of aluminum atoms in a hexagonal ternary carbide-layered structure with strong acid, ideally using hydrofluoric acid at room temperature [7]. Thus far, over 70 kinds of MAX and 30 types of MXenes have been discovered, and different kinds of MXenes with different n values are shown in Figure 3. As per theoretical calculations, there is a possibility to invent even more fascinating MXenes with exceptional properties. So far, about 30 MXene structures have been reported, which are obtained mainly by adding transition metals in the M layers [16,24,25,26,27,28]. Ti-based MXenes (Ti2CTx and Ti3C2Tx) are most widely studied for their green environmental applications. In addition to carbides, nitride, and carbonitrides, MXene composites have been commonly used in numerous fields [2,29,30,31,32,33,34,35,36]. The fantastic characteristics of MXenes include biocompatibility, enlarged surface area, and activated metallic hydroxide sites [20,22,37,38,39,40,41,42,43,44,45,46].
MXenes can form a stable colloidal solution in water because of their large negative zeta potential [47,48]. The functionalized surface assures the hydrophilic nature of MXenes and increases the probability of bonding with several compounds. In addition, the 2D nanostructures with a rapid transmission of charges are considered with a high surface area. MXene elements are adjustable and flexible, which shows that the bandgap of MXenes can be modified [49]. Moreover, the modification and high availability of active sites on MXenes’ surface make them more attractive. These novel 2D structures can be reprocessed (furthermore, the procedure is assisted by the capability to cycle between higher and lower valence states). Numerous studies have revealed that MXenes have strong applications, for example, in ion intercalation (K+, Na+, NH4+, etc.) and catalyst immobilization. [50] The existence of titanium oxide on the surface of Ti3C2 makes it one of the most prevalent MXenes, which also serves as a significant vehicle for electron transferal during H2O2 activation [51,52,53,54]. In recent years, MXenes have also been proven to be a promising material for applications other than water purification, such as capacitive deionization, membrane separation, and adsorption [7,51,55,56,57,58,59,60,61,62]. The etching of MAX phases and exfoliation route are shown in Figure 4.
Antibacterial/antibiofouling agent materials based on MXenes have been identified as suitable photothermal materials for photothermal membrane construction. A unique Ti3C2 MXene on a cellulose membrane was recently produced by simply immersing commercially available qualitative cellulose filter paper in an MXene solution at a concentration of 1 mg mL−1 for 3 s at room temperature, followed by vacuum drying for 10 min at 60 °C for several cycles. The membrane showed significant light absorption efficiency of up to 94% in a wide solar spectrum range and high antibacterial efficacy of up to 99% when tested in a suspension of E. coli and S. aureus [63]. The MXene/cellulose antibiofouling membrane was innovatively designed as a solar collector and a steam generator for purifying water. The created MXene-coated cellulose membrane, which has a flat surface and can be folded into several forms, exhibits flexibility. This study described how, following photothermal water purification, translucent condensed water was produced from the bacterial suspensions of S. aureus and E. coli. The following Table 1 explained the recent literatures for removal of pharmaceutical wastes via MXene-based nanocomposites.

3. MXenes as Adsorbents

Over the past ten years, 2D nanomaterials in which MXene family have drawn more and more scientific interest for adsorption applications [84,85]. Due to their environmentally friendly qualities, MXenes and MXene-based nanocomposites have recently been found to hold potential as adsorbents for the elimination of several environmental toxins [24,42,86] due to their high chemical and structural stability, as well as hydrophilic surfaces [87,88]. Different kinds of pollutants, for example, Cr(IV), Cd(II), Pb(II), Hg(II), and Cu, have all been successfully removed using titanium carbide (Ti3C2Tx), which has been shown to have ideal properties as an adsorbent (II) [89,90,91,92]. Adsorption is one of the best ways to get rid of heavy metal ions when biological and chemical processes are ineffective [93,94,95,96]. The adsorption of various metal ions using Ti3C2Tx MXene occurs through its electrostatic and chemical interactions, high surface area, hydrophilicity, and special surface functionalities [97]. The nature of the functional group available at the surface of MXenes, which is an addition from the essence of the intercalants and etchants used during the intercalation, delamination, and etching process, is linked to the capacity of the surface of MXenes to grasp an electrostatic charge [31,98]. Electrostatic interactions enable opposing charges of possible targets existing in the wastewater. By altering the pH of the solution, the surface of MXenes can be positively or negatively adjusted to adsorb charged objects [99,100].
Various radionuclides, dyes, heavy metals, and other pollutants that are the objectives in wastewater treatment could be targets of interest. Objectives for wastewater treatment include radionuclides, heavy metals, dyes, and other contaminants. The adsorptive behavior of Cr(VI), a substantial and highly toxic mutagenic heavy metal and carcinogenic water pollution, on MXenes has been studied [101]. MXene/PEI-altered sodium alginate aerogel showed a better adsorption capacity (538.97 mg g−1) toward the elimination of Cr(VI) and an ultrahigh adsorption capacity (3568 mg g−1) for Congo Red. The composite’s special adsorption capabilities are credited with the strong electrostatic behavior and the synergistic effects of intercalation and surface adsorption. This substance also demonstrated exceptional antibacterial activity against E. coli and S. aureus. The Langmuir adsorption isotherm and pseudo-second-order kinetic model are both in good agreement with the adsorption data [102]. A new imidazole MXene nanohybrid Ti3C2@IMIZ shows a strong adsorption affinity for Cr, as reported by G. Yang et al. (VI). The adsorption performance and procedure analysis demonstrate that electrostatic contact is the primary physical adsorption mechanism [103]. It has also been shown that alkalized MXenes have a considerably increased ability for adsorbing hazardous metal ions [88]. The alkalized MXenes benefit from a better photocatalytic efficiency and an improved adsorption capability due to this treatment’s increased interlayer spacing. The successful decrease in F-atoms on the MXene’s surface and the following increase in O atoms lead to an increase in OH groups, which readily adsorb small ions and molecules in the solution, as confirmed by the EDS analysis [104].
According to Zheng et al.’s analysis of the three different types of aqueous solutions that they fabricated, more OH groups led to better alkali-Ti3C2Tx adsorption after activation by LiOH, NaOH, and KOH. Compared with the pure Ti3C2Tx, each composite showed rapid methylene blue (MB) elimination rates because the terminal OH groups are so good at adsorbing cationic dyes. The MXene composite NaOH-Ti3C2Tx showed the highest adsorption capacity for MB among these three MXene adsorbents, at 189 mg g−1, followed by KOH-Ti3C2Tx (79 mg g−1) and LiOH-Ti3C2Tx (121 mg g−1). Organic contaminants may contain pharmaceuticals, dyes, pesticides, and biotoxins [105,106,107,108,109]. Wastewater complexity and multitude represent a crucial need for a highly efficient wastewater treatment process. Wastewater containing a too-high concentration of substances causes negative changes in its bioc enosis and performance when introduced to a biological reactor. Pharmaceuticals primarily cover assorted organic contaminants, for example, antimicrobial agents that may have an obstinate noxious action on microorganisms, causing hefty danger to the aquatic ecosystem. They are generated from different origins, such as from industrial, hospital, and domestic sources. Even a small amount of wastewater contains a large amount of highly complex and composite organic and inorganic contaminants. Due to their futile removal in the photocatalytic process, they are commonly discovered in natural water reservoirs.
The typical concentration of pharmaceutical contaminants in wastewater ranges between 1 ng−1 and 600 ugL−1 [110,111]. Pharmaceuticals in wastewater are somehow resistant to biodegradation, and because of this, these compounds become highly contagious for aquatic life. These pharmaceutical-activated compounds sustain harmful impacts on marine life as they are resistant to microbial degradation [112,113]. Due to the versatile chemical nature of medicines, removing these pollutants from wastewater is almost impossible. Various drugs, for example, mefenamic acid, carbamazepine, and non-steroidal anti-inflammatory drugs, are poorly eliminated from pharmaceutical wastewater, with a percentage of removal efficiency lower than 40%. Similar to the above situation, antibiotics have poor removal efficiency. With a poor removal efficiency of azithromycin (45%), erythromycin (30%), and clarithromycin (43%), pharmaceutical contaminants become persistent in the wastewater treatment process [114]. The elimination of pharmaceuticals in wastewater includes procedures for coagulation, flocculation, adsorption, ozonation, and biological treatment.
The coagulation and flocculation process has been proven inefficient, with an efficiency rate of 0 to 50%. With significant equilibrium times, adsorption has been evidenced to efficiently remove pharmaceutical contaminants, with an efficiency ranging from 20 to 99% [115]. Ozonation also has good efficiency in removing pharmaceutical pollutants from wastewater treatment plants. These processes somehow cause toxic binary products during degradation, which may also cause harmful impacts on the environment [116]. Antibiotics (such as ciprofloxacin) are present in high effluents after degradation. Their non-biodegradable behavior contributes to antimicrobial resistance, which may have detrimental effects on human health. According to a global health report, antimicrobial resistance is accountable for over 23,000 casualties in the United States [117]. In recent times, photocatalysis seems to be the topic of discussion in research development as it is a renewable technology with high efficiency in eliminating antibiotics from wastewater. Photocatalytic degradation of pharmaceuticals from wastewater can resolve environmental problems with an increased ability of degradation. The following diagram represents different preliminary wastewater treatment processes (Figure 5).
In recent years, photocatalysis of pharmaceutical pollutants has attracted great attention in research due to its hopeful potential of solving energy-related and environmental issues. The photocatalytic process can be described as involving (1) the production of electrons and holes with a photocatalyst through absorbing incoming light with a sufficient amount of energy; (2) the separation and movement of photo-generated electrons and holes to the photocatalyst’s surface; and (3) the simultaneous reactions that occur (oxidation and reduction) via consuming photo-generated charge carriers [97,119,120,121]. Photocatalysis promises an effective method for the degradation of pollutants as it involves solar energy as a driving force. In recent years, nanotechnology has been widely discussed for its practical approach to eradicating organic waste products from industrial wastewater, which is attributable to their unique structural and electrochemical properties. In particular, 2D nanomaterials, which present a large surface-to-volume ratio and the thinnest structures, assures a faster processing time and a higher treatment efficiency. In recent years, 2D materials have been prominently used for the photocatalytic degradation of organic materials, which also encompasses pharmaceutical drug degradation.

4. Synthesis of MXene-Derived Heterostructure Photocatalyst

In a previous study, a delaminated-Ti3C2Tx/alkalized-C3N4 (TC-aCN) heterostructure photocatalyst was fabricated by inserting Ti3C2Tx into alkalized nanosheets of g-C3N4 [122]. Firstly, the delaminated Ti3C2Tx was synthesized by scattering the Ti3AlC2 material into a HF solution (40%), and then dispersed with continuous magnetic stirring, washed with deionized (DI) water, and dried in a vacuum furnace. In the next step, the TC-aCN was formed by dispersing the D-TC, urea, and KOH in DI water, followed by continuous stirring through the formed cyano groups. The Schottky connection between the two composites strengthen the van der Waals heterostructure, improving photo-excited electron relocation and seizing of Ti3C2Tx. The graphitic carbon nitride (gCN) was fabricated in a calcination process based on the heterojunction of Ti3C2 and black phosphorus [123]. The synthesized materials were mixed into deionized water, and then black phosphorus nanosheets were added. The obtained solution was sonicated, vacuum dried, and then sintered for 2 h at 350 °C. Researchers have also employed anodization and chemical vapor deposition methods to synthesize photocatalysts, such as Ti3C2/g-C3N4/TiO2 nanotube arrays on Ti meshes [124] (Figure 6).
In one study, a homogenously distributed composite included Ti3C2, and the melamine was synthesized and implanted in the quartz boat. After that, gas was passed through the foil and enfolded the quartz boat. The photocatalyst was fabricated by sintering the quartz boat in a vacuum furnace (550 ◦C @ 3 h). The as-synthesized photocatalyst Ti3C2/g-C3N4/TNTAs revealed a higher photocurrent density compared to TNTAs (Figure 6). The CuFe2O4/MXene heterojunction was fabricated via a hydrothermal method [127]. The Ti3C2 was scattered in DI water, and then FeCl3•6H2O and CuCl2•2H2O were added. The precipitate was cleaned with DI water and dried after magnetic stirring, pH adjustment, and heating of the mixture. The MXene-Ti3C2/MoS2 hybrid, which has an exceptional photocatalytic efficiency, could also be synthesized using the conventional hydrothermal method, as shown in Figure 7 [127]. Similarly, Ag2WO4/Ti3C2 Schottky composite was fabricated by electrostatic traction [3].
The above figure represents the fabrication of MXene nanocomposites and nanosheets. Scientists have also discovered different processes of MXene-derived composite production.

4.1. Evaporation-Induced Synthesis of MXene Composites

An evaporation-induced self-assembly (EISA) technique was used to fabricate the Ti3C2/g-C3N4 composite [129]. Meanwhile, g-C3N4 nanosheets were distributed in DI water with ultrasonication, and Ti3C2 was mixed into the suspension; after that, magnetic stirring and evaporation of the solvent were performed. The suspension was then frozen and dried to obtain the Ti3C2/g-C3N4 composite.

4.2. In Situ Reductive Deposition Synthesis

The researchers produced nZVI@Ti3C2-based MXene nanosheets using an in situ reductive deposition method [130]. Firstly, Ti3C2 nanosheets were mixed into a FeSO4 solution and ultra-sonicated. Then, the composite was moved to a flask with a continuous flow of nitrogen by slowly mixing a NaBH4 solution and continuously stirring, followed by reducing Fe2+ to Fe0.
Fe(H2O)62+ (aq) + 2BH4 (aq) → Fe0 (s) + 2B(OH)3 (aq) + 7H2 (g)
They as-synthesized heterogeneous nano-photocatalyst was made of MXene nanosheets functionalized with CoFe2O4 nanoparticles with liquid self-assembly. The MXene and CoFe2O4 nanoparticles were scattered in DI water; then, after ultrasonication, it was mixed and stirred. The as-synthesized precipitates were cleaned with DI water and ethanol. Afterward, the composite was separated, dried in an oven, and ground for advanced purposes [131].

4.3. The Flux Synthesis Method of Nanocomposites

Some researchers also prepared C-TiO2/Bi4NbO8Cl material using the flux technique. In this method, a specific quantity of BiOCl, KCl/NaCl, Nb2O5, and Bi2O3 was mixed with Ti3C2. Afterward, the combined mixture was sintered for 1 h at 750 °C. The composite was washed with DI water and dried to achieve the C-TiO2/Bi4NbO8Cl [66].

4.4. Solvothermal Method

This method was used to fabricate Ti3C2-Bi/BiOCl heterojunction by in situ growing Bi/BiOCl on Ti3C2. Yb3+/Tm3+ co-doped Ti3C2/Ag/Ag3VO4 compounds were synthesized using the in situ reduction method via hydrothermal method [74]. The alkalized Ti3C2 was dissolved in DI water using ultra-sonication and mixed with TmCl3.6H2O and YbCl3.6H2O. The optimized amount of NH4VO3 was combined and then inserted for magnetic stirring. The resulting solution was heated hydrothermally for 8 h at 140 °C in an autoclave after adjusting the pH to 10. The final product was washed (DI and ethanol) and dried to obtain the composite material.

5. Photocatalysis by MXenes

Over the past few years, photocatalysis has been widely discussed for its advanced catalytic technology for eliminating organic contaminants. The exceptional properties of MXenes, such as the band gap of MXenes that can be tuned by modifying surface chemistry, make them perfect for catalysis, thus attracting a growing interest in using MXene and its derivatives for the purification of water. MXenes display distinctive adsorptive and reductive properties with a high electronic conductivity. MXenes also own exceptional photocatalytic activity toward several organic pollutants. MXene-derived nanocomposite AgNP-loaded MXene/polymer/Fe3O4, which is fabricated by introducing dimethyl sulfoxide into MXene, reveals a high photocatalytic activity toward 2-nitroaniline and 4-nitrophenol. Working effectively under sunlight, MXenes can also be an encouraging co-catalyst for photocatalytic consequences. After the wet chemical etching method, the many bonds that have attached to MXenes make such materials a promising catalyst for photocatalysis. MXenes have admirable metallic conductivity due to the conductive metal cores in the layered structure [132,133,134,135,136,137,138]. Ti3C2Tx (MXene derivative) and its composites have been extensively used for wastewater treatment to eradicate several pollutants, for example, organic dyes and heavy metals. In one of the most recent studies, nanosheets of Ti3C2Tx with Cu2O composite photocatalysts with a Schottky heterojunction were created for a visible-light-induced breakdown of tetracycline (TC) antibiotics [139]. In 50 min, the composite successfully removed 97.6% of TC. The superoxide anion radical (O2) and hole (h+) were responsible for the improved efficiency. Additionally, the splitting of electron and hole pair was made possible through the presence of a Schottky heterojunction. A Ti3C2-Bi/BiOCl composite was created by adding Ti3C2 nanosheets to Bi/BiOCl. The antibiotic ciprofloxacin (CIP), which is found in water, was broken down using the photoactivity of the components as they were produced. The improved CIP removal efficiency (89%) in correlation with Bi/BiOCl (71%) and BiOCl (10%) was observed [65]. Ti3C2’s capacity to create heterostructures, which boost separation efficiency, was credited with the rise in photocatalytic activity.

6. MXene-Based Photocatalysis

In recent years, many researchers have worked on MXene-based materials in photocatalysis. When visible light interacts with MXenes, electrons get excited and move to the conduction band (CB) of the MXene-based photocatalyst, producing holes in the valence band (VB). The holes combine with OH and make free hydroxyl radicals, whereas the electrons react with O2 to create superoxide radicals. The radicals are an essential part of degradation as the active species of radicals contribute to the degradation of pharmacological and organic contaminants. Consequently, non-toxic by-products are formed, i.e., H2O and CO2 [140,141]. Electrons and hole pairs are produced after irradiation (Equation (1)), where electrons are captured suddenly by the MXene sheet, thus reducing the chance of recombination (Equation (2)).
MXenecomposite + hvH+ + e
e + MXenecompositee (MXene trapping sites)
After that, the phase electrons merge with O2 to produce O2, whereas the holes react with OH to create OH (Equations (3) and (4)).
e (MXene trapping sites) + O2O2
h+ + H2OOH
Both species, as represented in Equations (3) and (4), are super reactive and change pharmaceutical wastes into H2O and CO2 after degradation (Equations (5) and (6)).
OH + Pharmaceutical pollutant → CO2 + H2O (degradation byproducts)
O2 + Pharmaceutical pollutant → CO2 + H2O (degradation byproducts)
The broader absorption of photons and the quick generation of electron–hole pair mark the MXene-based photocatalyst as an enhanced photocatalytic catalyst [142]. The mechanism of photocatalytic elimination is described in Figure 8.

MXene-Based Photocatalytic Materials

Since the discovery of MXenes in 2011, many academics have been drawn to investigate potential new prospects for photocatalysis, especially for producing materials that might be used as effective photocatalysts for CO2 and H2 reduction [31]. Additionally, such research is essential for identifying long-term responses to the world’s environmental and energy problems [144,145]. In recent years, many MXene derivatives have been studied for the degradation of pharmaceutical pollutants. Below is a description of some of the MXene-derived photocatalysts. MXenes have been coupled with carbon-based nanostructures and compounds with metal ions to provide exceptional photocatalytic properties. The general classification of MXene-based nanocomposites is shown in Figure 9.

7. Carbon-Based MXene Composites

Carbon-based MXene composites have been discussed by numerous scholars and have been proven as an auspicious material in visible light photocatalysis. According to the research, it may be assumed that when MXenes are combined with other chemicals, their structural makeup is altered to make them function well as photocatalysts. Fast charge carrier recombination, a wide band gap, and UV light absorption are problems for semiconductors and other photocatalytic materials. MXenes decrease other photocatalysts’ band gaps and stifle electron–hole recombination when used together. The conductive and electrical characteristics of MXenes amplify photo-generated electrons. The layered structure of MXenes offer a variety of catalytic sites as well as a surface for the pollutant’s selective adsorption. Most significantly, MXene sheets have been designed such that the bandgap of these composites makes it possible for light to be absorbed in the visible spectrum. A semiconductor with a crystalline laminar structure made up of two units, (i) 3-s-triazine ring and triazine ring (C3N3), is graphitic carbon nitride (g-C3N4) (C6N7) [146].
Nitrogen atoms’ lone-pair electrons impact the band structure [147]. Due to the weak van der Waals forces that bind neighboring layers in g-C3N4, morphological tuning is simple. Additionally, the appropriate g-bandgap of the C3N4’s structure (2.7 eV) gives it distinct optoelectronic capabilities. Due to these unique characteristics, g-C3N4 is now a preferred material for photo-degradation [148]. This improvement can be further increased by combining this advantage with 2D MXenes with a significantly higher surface area. Tetracycline hydrochloride pollutant has been successfully degraded by a visible-light active heterostructured photocatalyst g-C3N4/TiO2/Ti3C2. The homogenous distribution of Ti3C2 and g-C3N4 on the surface of TiO2 leads to enhanced charge carrier separation, resulting in an increased photocatalytic performance of up to 85.12% in 180 min. Furthermore, the stability of the photocatalyst was determined after five cycles and was found to be stable [124]. The catalyst Ti3C2/g-C3N4 has been confirmed as an efficient photocatalyst for eliminating levofloxacin antibiotics. It was discovered that Ti3C2 effectively activated g-C3N4 to increase its photo-degradation capacity. Levofloxacin’s degradation efficiency increased from 33.6% with g-C3N4 to 72% even when only 1% Ti3C2 MXene was employed. This enhanced catalyst’s capacity to respond to light photons and extend the separation of photo-generated charge carriers in the case of a composite could be because of the addition of more active sites [149].
A heterojunction of g-C3N4/Ti3C2 and MXene/MoSe2 was reported to effectively degrade enoxacin when exposed to visible light. The electron separation between Ti3C2 MXene-induced surfaces was linked to a high catalytic activity. Ti3C2 MXene served as a conduit for the flow of electrons between g-C3N4 and MoSe2 and served to prevent their recombination. Gatifloxacin (65%), Ofloxacin (100%), norfloxacin (80%), ciprofloxacin (80%), levofloxacin (100%), and moxifloxacin (100%) were just a few of the medications that the aforementioned composite demonstrated potential for degrading. Using MXene/g-C3N4/black phosphorus, the photocatalytic removal of ciprofloxacin has also been documented. Increased photocatalytic efficiency was built into the composite, which resulted in a synergistic interaction between MXene and g-C3N4 as well as P-bridging, which led to the production of an active heterojunction that emitted visible light. The photocatalytic degradation activity for ciprofloxacin using this photocatalyst was observed to be 99% after 60 min [123].

7.1. Metal-Based MXene Nanocomposites

To degrade pharmaceutical wastes, MXenes are combined with several metal-based structures, such as metal oxides, molybdates, tungstate, ferrites, sulfides, carbonates, and vanadates. Various medications have been photo-catalytically degraded using MXene–oxide composites. TiO2 alone has significant disadvantages, namely a high band gap and rapid electron–hole pair recombination. Using TiO2/Ti3C2Tx (MXene), the photocatalytic removal of carbamazepine has been described. The TiO2/Ti3C2Tx photocatalyst’s efficiency was improved by its modification with 2D Ti3C2Tx, which altered its structural restrictions, as well as by adding more e and h+ [150]. Different Ti2C-modified TiO2 nanocomposites had been created and embellished with Ag2O, Au, Ag, Pd, and PdO. These nanocomposites were discovered to be effective for refining salicylic acid, with band-gap values ranging from 0.90 to 1.31 eV. Their photocatalytic activity was ultimately enhanced, which resulted in the migration of electrons from the conduction band of TiO2 to Ti2C. Compared to pure CeO2, CeO2/Ti3C2-MXene composites produced by a hydrothermal method demonstrated 6.3 times higher photocatalytic effectiveness for tetracycline hydrochloride. It has been determined that an electric field that naturally exists between Ti3C2-MXene and CeO2 speeds up the synthesis of photo-induced electrons and improves the separation of charge carriers, which enhances Ti3C2-MXenes’ efficiency [65].
It has been discovered that a magnetic (Fe3O4)/Ti3C2Tx composite may effectively degrade diclofenac. For this case, the existence of magnetic nanoparticle Fe3O4 affects the adsorption capability, although Ti3C2Tx sheets provide a mechanism for the selective adsorption of organic pollutants using photocatalysis. Due to the synergistic effects of Fe3O4 and Ti3C2Tx, this composite also demonstrated excellent stability, with just a small loss in photocatalytic capacity (8.7%) up to seven photocatalytic reaction cycles [151]. Some scientists used Ti3C2TX-nanosheet/Cu2O composite to photo-catalytically degrade tetracycline, with the photocatalytic efficiency reaching up to 97.6% in just 50 min. This outstanding photocatalytic ability was attributed to the composite’s enhanced specific surface area following its fusion with Cu2O. More active sites were available because of the greater surface area, facilitating a speedy electron transport to the Cu2O surface. As a result, the separation and migration of photo-generated charge carriers (O2 and h+) were improved, thereby improving tetracycline’s photo-degradation efficiency [139]. The increased photocatalytic degradation of carbamazepine under sunlight has been reported for the hydrothermally produced composite of Fe linked with MXene-based TiO2. The improved transfer of electrons produced when light is contacted results from TiO2 growing on the surface of MXene. Additionally, adding iron improves the composite’s photocatalytic efficacy due to a greater charge carrier separation [152].

7.2. MXene–Metal Tungstate Composites

Ti3C2/Ag2WO4 composite has been described in the literature for the photocatalytic degradation of sulfadimidine (88.6%) and tetracycline hydrochloride (62.9%). Ag2WO4 confined the photocatalytic capacity, but the existence of Ti3C2 increased the photocatalytic performance. The enlarged surface area and accessibility of functional groups of Ti3C2 triggered and amended photocatalysis [4]. A Ti3C2 composite with Bi2WO6 has been reported for the photo-degradation of tetracycline hydrochloride and acts as an effective photocatalyst. The photocatalytic degradation performance was pretty high, which was around 97% within 60 min of irradiation with light. The surface area of Bi2WO6 was enhanced by incorporating Ti3C2, causing an enhancement of the adsorption capacity of Bi2WO6/Ti3C2 for tetracycline hydrochloride. The recombination rate of photo-induced charges was also decreased by the addition of Ti3C2 [68]. In this composite, the existence of MXene sheets upgraded the separation performance of photo-induced charge carriers and, finally, the photocatalytic efficiency. Compared to pure Bi2WO6, the synthesized composite (Bi2WO6/Nb2CTx) revealed an exceptional photo-degradation performance (83.1%) for tetracycline hydrochloride. Likewise, 2D heterostructures of Bi2WO6/Ti3C2 MXene have also been reported for the removal of amoxicillin within 40 min. A possible reason for the outstanding photocatalytic performance was due to the influence of Ti3C2 nanosheet for light absorption. Moreover, due to the lower Fermi level, electrons could transfer faster from Bi2WO6 to Ti3C2 [153].

7.3. MXene–Metal-Based Composites

A new class of materials, metal-based MXene nanocomposites have drawn a lot of interest because of their distinctive characteristics and prospective uses in a variety of industries. MXenes are a class of two-dimensional (2D) transition metal carbides and nitrides with tunable surface chemistry and high electrical conductivity. They are appealing for a variety of applications, such as energy storage, catalysis, and sensing, due to these characteristics. Metals can be included in MXene sheets to produce hybrid materials with unique characteristics. These metal-based MXene nanocomposites have undergone substantial research, and numerous applications have shown their special characteristics. For instance, it has been demonstrated that adding silver (Ag) to MXene sheets improves their antibacterial characteristics, while adding copper (Cu) increases their catalytic activity toward the reduction of 4-nitrophenol. For the manufacture of metal-based MXene nanocomposites, several techniques have been devised, including electrochemical, solid-state, and wet chemical techniques. The metal composition and final nanoparticle size may be precisely controlled using these techniques [7,154,155,156].
The composites of MXenes with ferrites, vanadates, molybdates, and phosphates have also been reported to be able to remove pharmaceutical wastes. Due to the presence of MXene in the composites, Ag3PO4/Ti3C2 showed the best photocatalytic efficiency when compared to Ag3PO4/rGO and Ag3PO4 for the photocatalytic degradation of tetracycline hydrochloride, chloramphenicol, and thiamphenicol. The photocatalytic degradation efficiency for the degradation of tetracycline did not decrease lower than 64%, even after eight cyclic runs. This shows its excellent stability due to the incorporation of Ti3C2, which improves the catalytic performance of Ag3PO4, and also due to the introduction of several surface functional groups that are hydrophilic. As a result, a strong relationship is developed with Ag3PO4 and provides the separation of photo-generated charge carriers [70]. Cao et al. [125] fabricated CuFe2O4/MXene using the sol-hydrothermal method for the photo-degradation of sulfamethazine in visible light. CuFe2O4/MXene showed a catalytic efficiency of up to 70% for the removal of sulfamethazine due to the synergistic effects of the two different components from which it was fabricated. The enhancement in photocatalytic behavior was attributed to the transfer of photoelectrons in the synthesized composite and the prolonging of the lifetime of charge carriers due to the loading of titanium carbide.
Cai and Zhao [5] reported visible-light responsive heterojunction Bi2MoO6/Ti3C2 for the photocatalytic degradation of tetracycline. The photocatalyst Bi2MoO6/Ti3C2 showed the highest photocatalytic efficiency and degraded tetracycline hydrochloride about 8.8 times more than the pure Bi2MoO6. A strong interaction was developed between the two interfaces, which caused a narrow distance between charge transporters and improved the separation of photo-induced charge carriers. The Bi2MoO6/Ti3C2 also produced new adsorption sites for interaction with photocatalysts and pharmaceutical wastes. A Ti3C2/Ag/Ag3VO4 composite loaded with Yb3+/Tm3+ has been reported to have exceptional photo-degradation efficiency for tetracycline. The composite eliminated about 97% of tetracycline within only 20 min, which was much higher than Ag/Ag3VO4 (68%) and Ti3C2/Ag/Ag3VO4 (83%). The presence of Ti3C2 played a significant and prominent role in modifying the photocatalytic activity of the composite. The improved photocatalytic activity of the composite was accredited to two main factors: (a) an enhancement of the absorption of visible light, and (b) a significant separation of charge carriers caused by the co-doping of Tm3+/Yb3+ metal ions. The photocatalytic mechanism showed that both species O2 and h+ played a vital part in the photocatalytic elimination of tetracycline [74].

7.4. MXene–Metal and Oxide–Metal Nanocomposites

MXene decorated with the black phosphorous (Ti3C2Tx/TiO2-BP) photocatalyst was studied and revealed to have a high-performance photo-degradation potential for tetracycline hydrochloride, which was about 92.70% under visible-light irradiation. It was confirmed that the collegial role of BP and Ti3C2Tx/TiO2 intensified the photocatalytic efficiency. This synergic effect significantly enhanced the stability of BP, increased absorption capacity in visible light, enriched the photo-generated transfer of electrons, extended the photo-excited electrons’ lifetime, and postponed the recombination of electron–hole pair. Furthermore, O2 radicals were the key active species to play a vital role in the photocatalytic procedure [157].

7.5. MXene–Metal Sulfide Nanocomposites

MXene–metal sulfide composites have also been studied for their photocatalytic removal efficiency for pharmaceutical wastes. For example, visible-light active CdS@Ti3C2@TiO2 fabricated by the hydrothermal process has been reported for its photo-degradation performance of sulfachloropyridazine, and it degraded sulfachloropyridazine completely within 88 min under visible light [149]. Moreover, Ti3C2/ZnIn2S4 has also been reported for its tremendous photocatalytic activity for the photo-degradation of tetracycline hydrochloride, which may be attributed to the advantage of 2D interaction between the two surfaces and the enlarged contact surface area for the transfer of photo-excited electrons. MXene/ZnIn2S4 was reported to have 1.8 times higher photocatalytic performance than pure ZnIn2S4 toward the breakdown of tetracycline hydrochloride. [158]. MXene/Ag2S composite acts as a photoactive catalyst for the removal of tetracycline hydrochloride under visible light irradiation. MXene/Ag2S has a much higher photocatalytic efficiency (94.91%) than either MXene (32.08%) or Ag2S (56.21%). The existence of several surface groups (-OH, -O, and -F) on MXene has been attributed to the increased efficiency, which improves the catalyst’s adsorption performance and catalytic capacity [159].
Some researchers reported a MXene-Ti3C2/MoS2 composite that was fabricated by the hydrothermal method to remove ranitidine under visible light. The MXene-Ti3C2/MoS2 composite demonstrated good photocatalytic activity against ranitidine. MoS2 is a photo-catalyst that is active in visible light due to its small band gap (1.8 eV). However, it has poor light absorption efficiency and performs catalytically poorly because of the quick recombination of photo-excited electrons and holes. Its interaction with Ti3C2 improves photocatalysis by facilitating the separation of charge carriers and reducing the recombination rate of photo-induced electrons and holes. [127] Furthermore, CaIn2S4/MXene Ti3C2Tx synthesized by the hydrothermal process has been reported for the degradation of tetracycline hydrochloride. Ti3C2Tx boosted the photocatalytic potential of bare CaIn2S4 by reducing charge carrier recombination and increasing absorption in the visible range (96%). Tetracycline hydrochloride was found to photodegrade with the help of superoxide radicals and holes [160].

8. Conclusions

The photocatalytic breakdown of antibiotics was outlined in this review. First, the effects of antibiotics in wastewater on human health and the environment were reviewed, as well as the process of antibiotic photocatalytic degradation, which depends on the production of free radicals and active oxygen species. An introduction of several popular photocatalysts, followed by an analysis of a few extensively used antibiotics, was presented. A standard method for improving a photocatalyst’s photocatalytic performance is heteroatom doping, mainly when metal atoms are utilized as dopants. The metal dopants may act as recombination centers at more significant concentrations, which can lower a photocatalyst’s effectiveness. Future studies should, therefore, concentrate on additional choices, such as doping with non-metals, such as sulfur, nitrogen, phosphorus, and boron. The efficacy of catalysts during photo-degradation investigations is also greatly influenced by physicochemical characteristics, including shape and surface areas. As has already been noted, future research on photocatalysts with different surface areas and various morphologies can significantly improve these catalysts’ activity. The degradation route offers a clear explanation of the fate and transformation of antibiotics during the process. The effectiveness of antibiotic breakdown can, therefore, be increased by investigating the photocatalytic mechanism at the atomic level. First, although the elimination rate of antibiotics is still being adjusted over the long term, the removal rate of the chemical oxygen demand during antibiotic degradation by photocatalysts is still relatively high, confirming the requirement for optimal antibiotic mineralization. A thorough investigation into intermediates is essential for enhancing the activity of catalysts because there are numerous intermediates throughout the photo-degradation process. Second, most tests involve regular and continual stirring to avoid material aggregation in the media during the photocatalytic breakdown of antibiotics, which necessitates additional energy consumption.
Thirdly, the accumulation of antibiotics in water results in the creation of microbial resistance genes, which contributes to antibiotic problems in addition to poor water quality. Research on the few resistance genes found in photocatalysts is still lacking. Finally, a photocatalyst’s capacity to be recycled is a crucial indicator of how cost-effective and practicable it is to use the photocatalyst to break down antibiotics. The design of photocatalysts with almost identical photoactivity across each cycle is desired to reduce potential waste. Additionally, it is crucial to create simpler photocatalysts that can be separated and recycled to prevent the loss of valuable materials during the photocatalytic process. Thus, from an economic perspective, it is essential to separate photocatalysts from the aqueous phase. It should be emphasized that a photocatalytic reaction’s operational costs primarily result from the photocatalyst being available for one-time usage and not being able to be recycled. In-depth research is still needed on how to use recyclable photocatalysts to responsibly degrade antibiotics.

Author Contributions

Writing—original draft preparation, S.I.; formal analysis, D.C., F.D. and S.B.K.; conceptualization, funding acquisition, and writing—review and editing, D.C. and M.A.U.D. All authors have read and agreed to the published version of the manuscript.

Funding

This work was partly financially supported by the Natural Science Foundation of Guangdong Province (2018B030311022), the National Natural Science Foundation of China (51574205 and 22002142), the Guangdong Innovation Research Team for Higher Education (2017KCXTD030), the High level Talents Project of Dongguan University of Technology (KCYKYQD2017017), the Engineering Research Center of None-Food Biomass Efficient Pyrolysis and Utilization Technology of Guangdong Higher Education Institutes (2016GCZX009), and the Program from Dongguan University of Technology (G200906-17).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data will be available upon request.

Conflicts of Interest

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

References

  1. Zhou, S.; Di Paolo, C.; Wu, X.; Seiler, T.-B.; Hollert, H. Optimization of screening-level risk assessment and priority of emerging pollutants–the case of pharmaceuticals in European surface waters. Environ. Int. 2019, 128, 1–10. [Google Scholar] [CrossRef] [PubMed]
  2. Zhang, S.; Liao, S.; Qi, F.; Liu, R.; Xiao, T.; Hu, J.; Li, K.; Wang, R.; Min, Y. Direct deposition of two-dimensional MXene nanosheets on commercially available filter for fast and efficient dye removal. J. Hazard. Mater. 2020, 384, 121367. [Google Scholar] [CrossRef] [PubMed]
  3. Luo, S.; Wang, R.; Yin, J.; Jiao, T.; Chen, K.; Zou, G.; Zhang, L.; Zhou, J.; Zhang, L.; Peng, Q. Preparation and dye degradation performances of self-assembled MXene-Co3O4 nanocomposites synthesized via solvothermal approach. ACS Omega 2019, 4, 3946. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Fang, Y.; Cao, Y.; Chen, Q. Synthesis of an Ag2WO4/Ti3C2 Schottky composite by electrostatic traction and its photocatalytic activity. Ceram. Int. 2019, 45, 22298–22307. [Google Scholar]
  5. Zhao, D.; Cai, C. Preparation of Bi2MoO6/Ti3C2 MXene heterojunction photocatalysts for fast tetracycline degradation and Cr(vi) reduction. Inorg. Chem. Front. 2020, 7, 2799–2808. [Google Scholar]
  6. Damptey, L.; Bright, N.J.; Camila, S.R.; Varagnolo, S.; Nicholas, P.P.; Vimalnath, S.; David, D.A.; Kumar, R.V.; Sithara, P.S.; Dermot, B.V.K.T.; et al. Surface Functionalized MXenes for Wastewater Treatment—A Comprehensive Review. Glob. Chall. 2022, 6, 2100120. [Google Scholar] [CrossRef]
  7. Naguib, M.; Kurtoglu, M.; Presser, V.; Lu, J.; Niu, J.; Heon, M.; Hultman, L.; Gogotsi, Y.; Barsoum, M.W. Two-dimensional nanocrystals produced by exfoliation of Ti3AlC2. Adv. Mater. 2011, 23, 4248–4253. [Google Scholar] [CrossRef] [Green Version]
  8. Naguib, M.; Mashtalir, O.; Carle, J.; Presser, V.; Lu, J.; Hultman, L.; Gogotsi, Y.; Barsoum, M.W. Two-dimensional transition metal carbides. ACS Nano 2012, 6, 1322–1331. [Google Scholar] [CrossRef]
  9. Naguib, M.; Mochalin, V.N.; Barsoum, M.W.; Gogotsi, Y. MXenes: A new family of two dimensional materials. Adv. Mater. 2014, 26, 992–1005. [Google Scholar] [CrossRef]
  10. Hemanth, N.R.; Kandasubramanian, B. Recent advances in 2D MXenes for enhanced cation intercalation in energy harvesting applications: A review. Chem. Eng. J. 2020, 392, 123678. [Google Scholar] [CrossRef]
  11. Anasori, B.; Xie, Y.; Beidaghi, M.; Lu, J.; Hosler, B.C.; Hultman, L.; Kent, P.R.C.; Gogotsi, Y.; Barsoum, M.W. Two-dimensional, ordered, double transition metals carbides (MXenes). ACS Nano 2015, 9, 9507–9516. [Google Scholar] [CrossRef] [PubMed]
  12. Anasori, B.; Lukatskaya, M.R.; Gogotsi, Y. 2D metal carbides and nitrides (MXenes) for energy storage. Nat. Rev. Mater. 2017, 2, 16098. [Google Scholar] [CrossRef]
  13. Sun, W.; Shah, S.A.; Chen, Y.; Tan, Z.; Gao, H.; Habib, T.; Radovic, M.; Green, M.J. Electrochemical etching of Ti2AlC to Ti2CTx (MXene) in low-concentration hydrochloric acid solution. J. Mater. Chem. A. 2017, 5, 21663–21668. [Google Scholar] [CrossRef]
  14. Yang, S.; Zhang, P.; Wang, F.; Ricciardulli, A.G.; Lohe, M.R.; Blom, P.W.M.; Feng, X. Fluoride-free synthesis of two-dimensional titanium carbide (MXene) using a binary aqueous system. Angew. Chem. 2018, 130, 15717–15721. [Google Scholar] [CrossRef]
  15. Lei, J.-C.; Zhang, X.; Zhou, Z. Recent advances in MXene: Preparation, properties, and applications. Front. Phys. 2015, 10, 276–286. [Google Scholar] [CrossRef]
  16. Deysher, G.; Shuck, C.E.; Hantanasirisakul, K.; Frey, N.C.; Foucher, A.C.; Maleski, K.; Sarycheva, A.; Shenoy, V.B.; Stach, E.A.; Anasori, B.; et al. Synthesis of Mo4VAlC4 MAX phase and two-dimensional Mo4VC4 MXene with five atomic layers of transition metals. ACS Nano 2020, 14, 204–217. [Google Scholar] [CrossRef]
  17. VahidMohammadi, A.; Rosen, J.; Gogotsi, Y. The world of two-dimensional carbides and nitrides (MXenes). Science 2021, 372, 1581. [Google Scholar] [CrossRef]
  18. Barsoum, M.W.; El-Raghy, T. Synthesis and characterization of a remarkable ceramic: Ti3SiC2. J. Am. Ceram. Soc. 1996, 79, 1953–1956. [Google Scholar] [CrossRef]
  19. Anasori, B.; Gogotsi, Y. (Eds.) 2D Metal Carbides and Nitrides (MXene); Springer International Publishing: Cham, Switzerland, 2019; pp. 89–109. [Google Scholar]
  20. Szuplewska, A.; Kulpinska, D.; Dybko, A.; Chudy, M.; Jastrzębska, A.M.; Olszyna, A.; Brzózka, Z. Future applications of MXenes in biotechnology, nanomedicine, and sensors. Trends Biotechnol. 2019, 38, 264–279. [Google Scholar] [CrossRef]
  21. Barsoum, M.W. The Mn+1AXn phases: A new class of solids: Thermodynamically stable nanolaminates. Prog. Solid State Chem. 2000, 28, 201–281. [Google Scholar] [CrossRef]
  22. 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] [PubMed]
  23. Shekhirev, M.; Shuck, C.E.; Sarycheva, A.; Gogotsi, Y. Characterization of MXenes at every step, from their precursors to single flakes and assembled films. Prog. Mater. Sci. 2021, 120, 100757. [Google Scholar] [CrossRef]
  24. Rasool, K.; Pandey, R.P.; Rasheed, P.A.; Buczek, S.; Gogotsi, Y.; Mahmoud, K.A. Water treatment and environmental remediation applications of two-dimensional metal carbides (MXenes). Mater. Today 2019, 30, 80–102. [Google Scholar] [CrossRef]
  25. 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]
  26. Khazaei, M.; Arai, M.; Sasaki, T.; Chung, C.-Y.; Venkataramanan, N.S.; Estili, M.; Sakka, Y.; Kawazoe, Y. Novel electronic and magnetic properties of two dimensional transition metal carbides and nitrides. Adv. Funct. Mater. 2013, 23, 2185–2192. [Google Scholar] [CrossRef]
  27. Khazaei, M.; Arai, M.; Sasaki, T.; Estili, M.; Sakka, Y. Two-dimensional molybdenum carbides: Potential thermoelectric materials of the MXene family. Phys. Chem. Chem. Phys. 2014, 16, 7841–7849. [Google Scholar] [CrossRef]
  28. Liang, X.; Rangom, Y.; Kwok, C.Y.; Pang, Q.; Nazar, L.F. Interwoven MXene nanosheet/carbon-nanotube composites as Li–S cathode hosts. Adv. Mater. 2017, 29, 1603040. [Google Scholar] [CrossRef]
  29. Naguib, M.; Adams, R.A.; Zhao, Y.; Zemlyanov, D.; Varma, A.; Nanda, J.; Pol, V.G. Electrochemical performance of MXenes as K-ion battery anodes. Chem. Commun. 2017, 53, 6883–6886. [Google Scholar] [CrossRef]
  30. Urbankowski, P.; Anasori, B.; Hantanasirisakul, K.; Yang, L.; Zhang, L.; Haines, B.; May, S.J.; Billinge, S.J.L.; Gogotsi, Y. 2D molybdenum and vanadium nitrides synthesized by ammoniation of 2D transition metal carbides (MXenes). Nanoscale. 2017, 9, 17722–17730. [Google Scholar] [CrossRef]
  31. Huang, K.; Li, Z.; Lin, J.; Han, G.; Huang, P. Two-dimensional transition metal carbides and nitrides (MXenes) for biomedical applications. Chem. Soc. Rev. 2018, 47, 5109–5124. [Google Scholar] [CrossRef]
  32. Hong Ng, V.M.; Huang, H.; Zhou, K.; Lee, P.S.; Que, W.; Xu, J.Z.; Kong, L.B. Recent progress in layered transition metal carbides and/or nitrides (MXenes) and their composites: Synthesis and applications. J. Mater. Chem. A 2017, 5, 3039–3068. [Google Scholar] [CrossRef]
  33. 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]
  34. 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]
  35. Sinha, A.; Zhao, H.; Huang, Y.; Lu, X.; Chen, J.; Jain, R. MXene: An emerging material for sensing and biosensing. Trends Anal. Chem. 2018, 105, 424–435. [Google Scholar] [CrossRef]
  36. Wang, N.N.; Wang, H.; Wang, Y.-Y.; Wei, Y.-H.; Si, J.-Y.; Yuen, A.C.Y.; Xie, J.-S.; Yu, B.; Zhu, S.-E.; Lu, H.-D.; et al. Robust, lightweight, hydrophobic, and fire-retarded polyimide/MXene aerogels for effective oil/water separation. ACS Appl. Mater. Interfaces 2019, 11, 40512–40523. [Google Scholar] [CrossRef]
  37. Liu, F.; Zhou, A.; Chen, J.; Jia, J.; Zhou, W.; Wang, L.; Hu, Q. Preparation of Ti3C2 and Ti2C MXenes by fluoride salts etching and methane adsorptive properties. Appl. Surf. Sci. 2017, 416, 781–789. [Google Scholar] [CrossRef]
  38. Guo, Z.; Zhou, J.; Zhu, L.; Sun, Z. MXene: A promising photocatalyst for water splitting. J. Mater. Chem. A 2016, 4, 11446–11452. [Google Scholar] [CrossRef]
  39. Jiang, H.; Wang, Z.; Yang, Q.; Tan, L.; Dong, L.; Dong, M. Ultrathin Ti3C2Tx (MXene) nanosheets wrapped NiSe2 octahedral crystal for enhanced supercapacitor performance and synergetic electrocatalytic water splitting. Nano-Micro Lett. 2019, 11, 31. [Google Scholar] [CrossRef] [Green Version]
  40. Zhang, Q.; Teng, J.; Zou, G.; Peng, Q.; Du, Q.; Jiao, T.; Xiang, J. Efficient phosphate sequestration for water purification by unique sandwich-like MXene/magnetic iron oxide nanocomposites. Nanoscale 2016, 8, 7085–7093. [Google Scholar] [CrossRef]
  41. Fu, L.; Yan, Z.; Zhao, Q.; Yang, H. Novel 2D nanosheets with potential applications in heavy metal purification: A review. Adv. Mater. Interfaces 2018, 5, 1801094. [Google Scholar] [CrossRef]
  42. Zhang, Y.; Wang, L.; Zhang, N.; Zhou, Z. Adsorptive environmental applications of MXene nanomaterials: A review. RSC Adv. 2018, 8, 19895–19905. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Zhang, C.; Nicolosi, V. Graphene and MXene-based transparent conductive electrodes and supercapacitors. Energy Storage Mater. 2019, 16, 102–125. [Google Scholar] [CrossRef]
  44. Shahzad, F.; Alhabeb, M.; Hatter, C.B.; Anasori, B.; Man Hong, S.; Koo, C.M.; Gogotsi, Y. Electromagnetic interference shielding with 2D transition metal carbides (MXenes). Science 2016, 353, 1137–1140. [Google Scholar] [CrossRef] [Green Version]
  45. Cao, W.T.; Chen, F.F.; Zhu, Y.J.; Zhang, Y.G.; Jiang, Y.Y.; Ma, M.G.; Chen, F. Binary strengthening and toughening of MXene/cellulose nanofiber composite paper with nacre-inspired structure and superior electromagnetic interference shielding properties. ACS Nano 2018, 12, 4583–4593. [Google Scholar] [CrossRef]
  46. Ihsanullah, I. Potential of MXenes in Water Desalination: Current Status and Perspectives. Nano-Micro Lett. 2020, 12, 72. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  47. Wojciechowska, A. Colloidal Properties and Stability of 2D Ti3C2 and Ti2C MXenes in Water. Int. J. Electrochem. Sci. 2018, 13, 10837–10847. [Google Scholar]
  48. Orimolade, B.O.; Idris, A.O.; Feleni, U.; Mamba, B. Recent advances in degradation of pharmaceuticals using Bi2WO6 mediated photocatalysis—A comprehensive review. Environ. Pollut. 2021, 289, 117891. [Google Scholar] [CrossRef] [PubMed]
  49. Cui, Y.; Zhang, D.; Shen, K.; Nie, S.; Liu, M.; Huang, H.; Deng, F.; Zhou, N.; Zhang, X.; Wei, Y. Biomimetic anchoring of Fe3O4 onto Ti3C2 MXene for highly efficient removal of organic dyes by Fenton reaction. J. Environ. Chem. Eng. 2020, 8, 104369. [Google Scholar] [CrossRef]
  50. Jiang, X.; Kuklin, A.V.; Baev, A.; Ge, Y.; Ågren, H.; Zhang, H.; Prasad, P.N. Two-dimensional MXenes: From morphological to optical, electric, and magnetic properties and applications. Phys. Rep. 2020, 848, 1–58. [Google Scholar] [CrossRef]
  51. Szuplewska, A.; Kulpińska, D.; Jakubczak, M.; Dybko, A.; Chudy, M.; Olszyna, A.; Brzózka, Z.; Jastrz˛ebska, A.M. The 10th anniversary of MXenes: Challenges and prospects for their surface modification toward future biotechnological applications. Adv. Drug Deliv. Rev. 2022, 182, 114099. [Google Scholar] [CrossRef]
  52. Jakubczak, M.; Karwowska, E.; Rozmysłowska-Wojciechowska, A.; Petrus, M.; Woźniak, J.; Mitrzak, J.; Jastrz˛ebska, A.M. Filtration Materials Modified with 2D Nanocomposites—A New Perspective for Point-of-Use Water Treatment. Materials 2021, 14, 182. [Google Scholar] [CrossRef] [PubMed]
  53. Szuplewska, A.; Rozmysłowska-Wojciechowska, A.; Pózniak, S.; Wojciechowski, T.; Birowska, M.; Popielski, M.; Chudy, M.; Ziemkowska, W.; Chlubny, L.; Moszczynska, D.; et al. Multilayered stable 2D nano-sheets of Ti2NTx MXene: Synthesis, characterization, and anticancer activity. J. Nanobiotechnol. 2019, 17, 114. [Google Scholar] [CrossRef] [PubMed]
  54. Pang, J.; Mendes, R.G.; Bachmatiuk, A.; Zhao, L.; Ta, H.Q.; Gemming, T.; Liu, H.; Liu, Z.; Rummeli, M.H. Applications of 2D MXenes in energy conversion and storage systems. Chem. Soc. Rev. 2019, 48, 72–133. [Google Scholar] [CrossRef] [PubMed]
  55. Zhao, Y.; Zhang, S.; Shi, R.; Waterhouse, G.I.N.; Tang, J.; Zhang, T. Two-dimensional photocatalyst design: A critical review of recent experimental and computational advances. Mater. Today 2020, 34, 78–91. [Google Scholar] [CrossRef]
  56. Xu, J.; Li, W.; Hou, Y. Two-Dimensional Magnetic Nanostructures. Trends Chem. 2020, 2, 163–173. [Google Scholar] [CrossRef]
  57. Anichini, C.; Czepa, W.; Pakulski, D.; Aliprandi, A.; Ciesielski, A.; Samorì, P. Chemical sensing with 2D materials. Chem. Soc. Rev. 2018, 47, 4860–4908. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Li, X.; Bai, Y.; Shi, X.; Su, N.; Nie, G.; Zhang, R.; Nie, H.; Ye, L. Applications of MXene (Ti3C2Tx) in photocatalysis: A review. Mater. Adv. 2021, 2, 1570–1594. [Google Scholar] [CrossRef]
  59. Tiwari, S.K.; Sahoo, S.; Wang, N.; Huczko, A. Graphene research and their outputs: Status and prospect. J. Sci. Adv. Mater. Devices 2020, 5, 10–29. [Google Scholar] [CrossRef]
  60. Katsnelson, M.I. Graphene: Carbon in two dimensions. Mater. Today 2007, 10, 20–27. [Google Scholar] [CrossRef]
  61. Mas-Ballesté, R.; Navarro, C.; Zamora, F. 2D materials: To graphene and beyond. Nanoscale 2011, 3, 20–30. [Google Scholar] [CrossRef]
  62. Jakubczak, M.; Szuplewska, A.; Rozmysłowska-Wojciechowska, A.; Rosenkranz, A.; Jastrz˛ebska, A.M. Novel 2D MBenes-Synthesis, Structure, and Biotechnological Potential. Adv. Funct. Mater. 2021, 31, 2103048. [Google Scholar] [CrossRef]
  63. Zha, X.-J.; Zhao, X.; Pu, J.-H.; Tang, L.-S.; Ke, K.; Bao, R.-Y.; Bai, L.; Liu, Z.-Y.; Yang, M.-B.; Yang, W. Flexible Anti-Biofouling MXene/Cellulose Fibrous Membrane for Sustainable Solar-Driven Water Purification. ACS Appl. Mater. Interfaces 2019, 11, 36589. [Google Scholar] [CrossRef] [PubMed]
  64. Li, J.-Y.; Li, Y.-H.; Zhang, F.; Tang, Z.-R.; Xu, Y.-J. Visible-light-driven integrated organic synthesis and hydrogen evolution over 1D/2D CdS-Ti3C2Tx MXene composites. Appl. Catal. B Environ. 2020, 269, 118783. [Google Scholar] [CrossRef]
  65. Shen, J.Y.; Shen, J.; Zhang, W.; Yu, X.; Tang, H.; Zhang, M.; Zulfiwar; Liu, Q. Built-in electric field induced CeO2/Ti3C2-MXene Schottky-junction for coupled photocatalytic tetracycline degradation and CO2 reduction. Ceram. Int. 2019, 45, 24146–24153. [Google Scholar] [CrossRef]
  66. Daixun, J.; Sun, X.; Wu, X.; Zhang, S.; Qu, X.; Shi, L.; Zhang, Y.; Du, F. MXene-Ti3C2 assisted one-step synthesis of carbon-supported TiO2/Bi4NbO8Cl heterostructures for enhanced photocatalytic water decontamination. Nanophotonics 2020, 9, 2077–2088. [Google Scholar]
  67. Qin, J.; Hu, X.; Li, X.; Yin, Z.; Liu, B.; Lam, K.-h. 0D/2D AgInS2/MXene Z-scheme heterojunction nanosheets for improved ammonia photosynthesis of N2. Nanomater. Energy 2019, 61, 27–35. [Google Scholar] [CrossRef]
  68. Zhao, D.; Cai, C. Layered Ti3C2 MXene modified two-dimensional Bi2WO6 composites with enhanced visible light photocatalytic performance. Mater. Chem. Front. 2019, 3, 2521–2528. [Google Scholar] [CrossRef]
  69. Li, J.; Zhao, L.; Wang, S.; Li, J.; Wang, G.; Wang, J. In situ fabrication of 2D/3D g-C3N4/Ti3C2 (MXene) heterojunction for efficient visible-light photocatalytic hydrogen evolution. Appl. Surf. Sci. 2020, 515, 145922. [Google Scholar] [CrossRef]
  70. Cai, T.; Wang, L.; Liu, Y.; Zhang, S.; Dong, W.; Chen, H.; Yi, X.; Yuan, J.; Xia, X.; Liu, C.; et al. Ag3PO4/Ti3C2 MXene interface materials as a Schottky catalyst with enhanced photocatalytic activities and anti-photocorrosion performance. Appl. Catal. B Environ. 2018, 239, 545–554. [Google Scholar] [CrossRef]
  71. He, F.; Zhu, B.; Cheng, B.; Yu, J.; Ho, W.; Macyk, W. 2D/2D/0D TiO2/C3N4/Ti3C2 MXene composite S-scheme photocatalyst with enhanced CO2 reduction activity. Appl. Catal. B Environ. 2020, 272, 119006. [Google Scholar] [CrossRef]
  72. Li, B.; Song, H.; Han, F.; Wei, L. Photocatalytic oxidative desulfurization and denitrogenation for fuels in ambient air over Ti3C2/g-C3N4 composites under visible light irradiation. Appl. Catal. B Environ. 2020, 269, 118845. [Google Scholar] [CrossRef]
  73. Xu, T.; Wang, J.; Cong, Y.; Jiang, S.; Zhang, Q.; Zhu, H.; Li, Y.; Li, X. Ternary BiOBr/TiO2/Ti3C2Tx MXene nanocomposites with heterojunction structure and improved photocatalysis performance. Chin. Chem. Lett. 2020, 31, 1022–1025. [Google Scholar] [CrossRef]
  74. Fang, H.; Pan, Y.; Yan, H.; Qin, X.; Wang, C.; Xu, L.; Pan, C. Facile preparation of Yb3+/Tm3+ co-doped Ti3C2/Ag/Ag3VO4 composite with an efficient charge separation for boosting visible-light photocatalytic activity. Appl. Surf. Sci. 2020, 527, 146909. [Google Scholar] [CrossRef]
  75. Vigneshwaran, S.; Karthikeyan, P.; Park, C.M.; Meenakshi, S. Boosted insights of novel accordion-like (2D/2D) hybrid photocatalyst for the removal of cationic dyes: Mechanistic and degradation pathways. J. Environ. Manag. 2020, 273, 111125. [Google Scholar] [CrossRef]
  76. Li, J.; Wang, S.; Du, Y.; Liao, W. Enhanced photocatalytic performance of TiO2@C nanosheets derived from two-dimensional Ti2CTx. Ceram. Int. 2018, 44, 7042–7046. [Google Scholar] [CrossRef]
  77. Yao, Z.; Sun, H.; Sui, H.; Liu, X. Construction of BPQDs/Ti3C2@TiO2 composites with favorable charge transfer channels for enhanced photocatalytic activity under visible light irradiation. Nanomaterials 2020, 10, 15. [Google Scholar] [CrossRef] [Green Version]
  78. Yin, J.; Zhan, F.; Jiao, T.; Wang, W.; Zhang, G.; Jiao, J.; Jiang, G.; Zhang, Q.; Gu, J.; Peng, Q. Facile preparation of self-assembled MXene@Au@CdS nanocomposite with enhanced photocatalytic hydrogen production activity. Sci. China-Mater. 2020, 63, 2228–2238. [Google Scholar] [CrossRef]
  79. Li, Y.; Ding, L.; Yin, S.; Liang, Z.; Xue, Y.; Wang, X.; Cui, H.; Tian, J. Photocatalytic H2 Evolution on TiO2 Assembled with Ti3C2 MXene and Metallic 1T-WS2 as Co-catalysts. Nano-Micro Lett. 2019, 12, 1–12. [Google Scholar] [CrossRef] [Green Version]
  80. Li, Y.J.; Yang, S.; Liang, Z.; Xue, Y.; Cui, H.; Tian, J. 1T-MoS2 nanopatch/Ti3C2 MXene/TiO2 nanosheet hybrids for efficient photocatalytic hydrogen evolution. Mater. Chem. Front. 2019, 3, 2673–2680. [Google Scholar] [CrossRef]
  81. Ding, X.; Li, Y.; Li, C.; Wang, W.; Feng, L.; Han, D. 2D visible-light-driven TiO2@Ti3C2/g-C3N4 ternary heterostructure for high photocatalytic activity. J. Mater. Sci. 2019, 54, 9385–9396. [Google Scholar] [CrossRef]
  82. Wu, Z.; Liang, Y.; Yuan, X.; Zou, D.; Fang, J.; Jiang, L.; Zhang, J.; Yang, H.; Xiao, Z. MXene Ti3C2 derived Z–scheme photocatalyst of graphene layers anchored TiO2/g–C3N4 for visible light photocatalytic degradation of refractory organic pollutants. Chem. Eng. J. 2020, 394, 124921. [Google Scholar] [CrossRef]
  83. Wu, S.; Su, Y.; Zhu, Y.; Zhang, Y.; Zhu, M. In-situ growing Bi/BiOCl microspheres on Ti3C2 nanosheets for upgrading visible-light-driven photocatalytic activity. Appl. Surf. Sci. 2020, 520, 146339. [Google Scholar] [CrossRef]
  84. Mohanty, R.; Mishra, A.; Khatei, J.; Singh, L.; Mahapatra, D.M. Adapting 2D Nanomaterials for Advanced Applications; ACS Publications: Washington, DC, USA, 2020; pp. 1–31. [Google Scholar]
  85. Shao, B.; Liu, Z.; Zeng, G.; Wang, H.; Liang, Q.; He, Q.; Cheng, M.; Zhou, C.; Jiang, L.; Song, B. Two-dimensional transition metal carbide and nitride (MXene) derived quantum dots (QDs): Synthesis, properties, applications and prospects. J. Mater. Chem. A 2020, 8, 7508–7535. [Google Scholar] [CrossRef]
  86. He, P.; Wang, X.X.; Cai, Y.Z.; Shu, J.C.; Zhao, Q.L.; Yuan, J.; Cao, M.S. Tailoring Ti3C2Tx nanosheets to tune local conductive network as an environmentally friendly material for highly efficient electromagnetic interference shielding. Nanoscale 2019, 11, 6080. [Google Scholar] [CrossRef] [PubMed]
  87. Wu, C.-W.; Unnikrishnan, B.; Chen, I.-W.P.; Harroun, S.G.; Chang, H.-T.; Huang, C.-C. Excellent oxidation resistive MXene aqueous ink for micro-supercapacitor application. Energy Storage Mater. 2020, 25, 563–571. [Google Scholar] [CrossRef]
  88. Wei, Z.; Peigen, Z.; Wubian, T.; Xia, Q.; Yamei, Z.; Zheng Ming, S. Alkali treated Ti3C2Tx MXenes and their dye adsorption performance. Mater. Chem. Phys. 2018, 206, 270. [Google Scholar] [CrossRef]
  89. Shahzad, A.; Rasool, K.; Miran, W.; Nawaz, M.; Jang, J.; Mahmoud, K.A.; Lee, D.S. Two-Dimensional Ti3C2Tx MXene Nanosheets for Efficient Copper Removal from Water. ACS Sustain. Chem. Eng. 2017, 5, 11481. [Google Scholar]
  90. Peng, Q.; Guo, J.; Zhang, Q.; Xiang, J.; Liu, B.; Zhou, A.; Liu, R.; Tian, Y. Unique Lead Adsorption Behavior of Activated Hydroxyl Group in Two-Dimensional Titanium Carbide J. Am. Chem. Soc. 2014, 136, 4113. [Google Scholar] [CrossRef]
  91. Shahzad, A.; Rasool, K.; Miran, W.; Nawaz, M.; Jang, J.; Mahmoud, K.A.; Lee, D.S.J. Mercuric ion capturing by recoverable titanium carbide magnetic nanocomposite. J. Hazard. Mater. 2018, 344, 811. [Google Scholar]
  92. Qiao, X.Q.; Hu, F.C.; Tian, F.Y.; Hou, D.F.; Li, D.S. Equilibrium and kinetics studies on MB adsorption by ultrathin 2D MoS2 nanosheets. RSC Adv. 2016, 6, 11631. [Google Scholar] [CrossRef]
  93. Gunatilake, S.K. Methods of Removing Heavy Metals from Industrial Wastewater. J. Multidiscip. Eng. Sci. Stud. 2015, 1, 12. [Google Scholar]
  94. Demirbas, A.J. Heavy metal adsorption onto agro-based waste materials: A review. J. Hazard. Mater. 2008, 157, 220. [Google Scholar] [CrossRef] [PubMed]
  95. Hua, M.; Zhang, S.; Pan, B.; Zhang, W.; Lv, L.; Zhang, Q. Heavy metal removal from water/wastewater by nanosized metal oxides: A review. J. Hazard. Mater. 2012, 211, 317. [Google Scholar] [CrossRef]
  96. Tofighy, M.A.; Mohammadi, T.J. Adsorption of divalent heavy metal ions from water using carbon nanotube sheets. J. Hazard. Mater. 2011, 185, 140. [Google Scholar] [CrossRef] [PubMed]
  97. Meng, A.; Zhang, L.; Cheng, B.; Yu, J. Dual cocatalysts in TiO2 photocatalysis. Adv. Mater. 2019, 31, 1807660. [Google Scholar] [CrossRef] [PubMed]
  98. Zhang, T.; Zhang, W.; Xi, H.; Li, Q.; Shen, M.; Ying, G.; Zhang, J. Polydopamine functionalized cellulose-MXene composite aerogel with superior adsorption of methylene blue. Cellulose 2021, 28, 4281. [Google Scholar] [CrossRef]
  99. Mu, W.; Du, S.; Yu, Q.; Li, X.; Wei, H.; Yang, Y. Improving barium ion adsorption on two-dimensional titanium carbide by surface modification. Dalton. Trans. 2018, 47, 8375. [Google Scholar] [CrossRef]
  100. Liu, T.; Liu, X.; Graham, N.; Yu, W.; Sun, K.J. Two-dimensional MXene incorporated graphene oxide composite membrane with enhanced water purification performance. Membr. Sci. 2020, 593, 117431. [Google Scholar]
  101. Jin, L.; Chai, L.; Yang, W.; Wang, H.; Zhang, L. Two-Dimensional Titanium Carbides (Ti3C2Tx) Functionalized by Poly(m-phenylenediamine) for Efficient Adsorption and Reduction of Hexavalent Chromium. Int. J. Environ. Res. Public Health 2019, 17, 167. [Google Scholar] [CrossRef] [Green Version]
  102. Feng, Y.; Wang, H.; Xu, J.; Du, X.; Cheng, X.; Du, Z.; Wang, H. Fabrication of MXene/PEI functionalized sodium alginate aerogel and its excellent adsorption behavior for Cr(VI) and Congo Red from aqueous solution. J. Hazard. Mater. 2021, 416, 125777. [Google Scholar] [CrossRef]
  103. Yang, G.; Hu, X.; Liang, J.; Huang, Q.; Dou, J.; Tian, J.; Deng, F.; Liu, M.; Zhang, X.; Wei, Y.J. Surface functionalization of MXene with chitosan through in-situ formation of polyimidazoles and its adsorption properties. J. Hazard. Mater. 2021, 419, 126220. [Google Scholar] [CrossRef] [PubMed]
  104. Li, Z.; Wang, L.; Sun, D.; Zhang, Y.; Liu, B.; Hu, Q.; Zhou, A. Synthesis and thermal stability of two-dimensional carbide MXene Ti3C2. Mater. Sci. Eng. B 2015, 191, 33. [Google Scholar] [CrossRef]
  105. Ghosh, A.; Nayak, A.K.; Pal, A. Nano-Particle-Mediated Wastewater Treatment: A Review. Curr. Pollut. 2017, 3, 17–30. [Google Scholar] [CrossRef]
  106. Zunita, M. Graphene oxide-based nanofilteration for Hg removal from wastewater: A mini Review. Membranes 2021, 11, 269. [Google Scholar] [CrossRef] [PubMed]
  107. Kommu, A.; Singh, J.K. A review on graphene-based materials for removal of toxic pollutants from wastewater. Soft Mater. 2020, 18, 297–322. [Google Scholar] [CrossRef]
  108. Sires, I.; Brillas, E.; Oturan, M.A.; Panizza, M. Electrochemical advanced oxidation processes: Today and tomorrow. A review. Environ. Sci. Pollut. Res. 2014, 21, 8336–8367. [Google Scholar] [CrossRef]
  109. Ibrahim, Y.; Kassab, A.; Eid, K.; Abdullah, A.M.; Ozoemena, K.I.; Elzatahry, A. Unveiling fabrication and environmental remediation of mxene-based nanoarchitectures in toxic metals removal from wastewater: Strategy and mechanism. Nanomaterials 2020, 10, 885. [Google Scholar] [CrossRef]
  110. Lindsey, M.E.; Meyer, M.; Thurman, E.M. Analysis of trace level of Sulfonamide and Tetracycline Antimicrobials in Ground water and Surface water using Solid-Phase Extraction and Liquid Chromatography/Mass spectrometery. Anal. Chem. 2001, 73, 4640–4646. [Google Scholar] [CrossRef]
  111. Santos, J.L.; Aparicio, I.; Callejon, M.; Alonso, E. Occurrence of pharmaceutically active compounds during 1-year in wastewater from four wastewater treatment plants in Seville (Spain). J. Hazard. Mater. 2009, 164, 1509–1516. [Google Scholar] [CrossRef]
  112. Kosma, C.I.; Nannou, C.I.; Boti, V.I.; Albanis, T.A. Psychiatrics and selected metabolites in hospital and urban wastewater: Occurrence, removal, mass loading, seasonal influence and risk assessment. Sci. Total Environ. 2019, 659, 1473–1483. [Google Scholar] [CrossRef]
  113. Vickers, N.J. Animal communication: When I am calling you, will you answer too? Curr. Biol. 2017, 27, R713–R715. [Google Scholar] [CrossRef] [PubMed]
  114. Monteil, H.; Pechaud, Y.; Oturan, M.A. A review on efficiency and cost effectiveness of electro- and bio-electro-Fenton processes: Applications to the treatment of pharmaceutical pollutants in water. Chem. Eng. J. 2019, 376, 119577. [Google Scholar] [CrossRef]
  115. Luo, Y.; Guo, W.; Ngo, H.-H.; Nghiiem, L.-D.; Hai, F.I.; Zhang, J.; Liang, S.; Wang, X.C. A review on the occurrence of micropollutants in the aquatic environment and their fate and removal during wastewater treatment. Sci. Total Environ. 2014, 473–474, 619–641. [Google Scholar] [CrossRef] [PubMed]
  116. Kharel, S.; Stapf, M.; Miehe, U.; Ekbald, M.; Cimbritz, M.; Falas, P.; Nilsson, J.; Sehlen, R.; Bester, K. Ozone dose dependent formation and removal of ozonation products of pharmaceuticals in pilot and full-scale municipal wastewater treatment plants. Sci. Total Environ. 2020, 731, 139064. [Google Scholar] [CrossRef] [PubMed]
  117. Morris, D.; Harris, S.; Morris, C.; Commind, E.; Cormican, M. Hospital Effluent: Impact on the Microbial Environment and Risk to the Human Health; Report No. 162; Environmental Protection Agency: Wexford, Ireland, 2016. [Google Scholar]
  118. Dominika, S.; Dominika, B.; Marcinowski, P.; Bogacki, J.; Michał, J.; Agnieszka, J.-E. Two-Dimensional Nanostructures in the World of Advanced Oxidation Processes. Catalysts 2022, 12, 358. [Google Scholar]
  119. Low, J.; Yu, J.; Jaroniec, M.; Wageh, S.; Al-Ghamdi, A.A. Heterojunction photocatalysis. Adv. Mater. 2017, 29, 1601694. [Google Scholar]
  120. Chen, Z.; Yang, S.; Tian, Z.; Zhu, B. Nis and graphene as dual cocatalysts for the enhanced photocatalytic H2 production activity of g-C3N4. Appl. Surf. Sci. 2019, 469, 657. [Google Scholar] [CrossRef]
  121. Gao, M.; Zhu, L.; Peh, C.K.; Ho, W. Solar absorber material and system designs for photothermal water vaporization towards clean water and energy production. Energy Environ. Sci. 2019, 12, 841. [Google Scholar]
  122. 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]
  123. 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]
  124. 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]
  125. 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] [PubMed]
  126. 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]
  127. Zou, X.; Zhao, X.; Zhang, J.; Lv, W.; Qiu, L.; Zhang, Z. Photocatalytic degradation of ranitidine and reduction of nitrosamine dimethylamine formation potential over MXene–Ti3C2/MoS2 under visible light irradiation. J. Hazard. Mater. 2021, 413, 125424. [Google Scholar] [CrossRef]
  128. Wu, Y.; Li, X.; Yang, Q.; Wang, D.; Yao, F.; Cao, J.; Chen, Z.; Huang, X.; Yang, Y.; Li, X. Mxene-modulated dual-heterojunction generation on a metal-organic framework (MOF) via surface constitution reconstruction for enhanced photocatalytic activity. Chem. Eng. J. 2020, 390, 124519. [Google Scholar] [CrossRef]
  129. Liu, N.; Lu, N.; Su, Y.; Wang, P.; Quan, X. Fabrication of g-C3N4/Ti3C2 composite and its visible-light photocatalytic capability for ciprofloxacin degradation. Sep. Purif. Technol. 2019, 211, 782–789. [Google Scholar] [CrossRef]
  130. Ma, Y.; Lv, X.; Xiong, D.; Zhao, X.; Zhang, Z. Catalytic degradation of ranitidine using novel magnetic Ti3C2-based MXene nanosheets modified with nanoscale zero-valent iron particles. Appl. Catal. B Environ. 2021, 284, 119720. [Google Scholar] [CrossRef]
  131. Fayyaz, A.; Saravanakumar, K.; Talukdar, K.; Kim, Y.; Yoon, Y.; Park, C.M. Catalytic oxidation of naproxen in cobalt spinel ferrite decorated Ti3C2Tx MXene activated persulfate system: Mechanisms and pathways. Chem. Eng. J. 2021, 407, 27842. [Google Scholar] [CrossRef]
  132. Li, Z.; Wu, Y. 2D Early Transition Metal Carbides (MXenes) for Catalysis. Small 2019, 15, 1804736. [Google Scholar] [CrossRef] [Green Version]
  133. Kuang, P.; Low, J.; Cheng, B.; Yu, J.; Fan, J.J. MXene-based photocatalysts. Mater. Sci. Technol. 2020, 56, 18. [Google Scholar] [CrossRef]
  134. Yang, C.; Tan, Q.; Li, Q.; Zhou, J.; Fan, J.; Li, B.; Sun, J.; Lv, K. 2D/2D Ti3C2 MXene/g-C3N4 nanosheets heterojunction for high efficient CO2 reduction photocatalyst: Dual effects of urea. Appl. Catal. B 2020, 268, 118738. [Google Scholar] [CrossRef]
  135. Xu, C.; Yang, F.; Deng, B.; Che, S.; Yang, W.; Zhang, G.; Sun, Y.; Li, Y.J. RGO-wrapped Ti3C2/TiO2 nanowires as a highly efficient photocatalyst for simultaneous reduction of Cr(VI) and degradation of RhB under visible light irradiation. J. Alloys Compd. 2021, 874, 159865. [Google Scholar] [CrossRef]
  136. Yang, J.-X.; Yu, W.-B.; Li, C.-F.; Dong, W.-D.; Jiang, L.-Q.; Zhou, N.; Zhuang, Z.-P.; Liu, J.; Hu, Z.-Y.; Zhao, H.; et al. PtO nanodots promoting Ti3C2 MXene in-situ converted Ti3C2/TiO2 composites for photocatalytic hydrogen production. Chem. Eng. J. 2021, 420, 129695. [Google Scholar]
  137. Zhang, L.; Ma, P.; Dai, L.; Li, S.; Yu, W.; Guan, J. In situ crystallization and growth of TiO2 nanospheres between MXene layers for improved adsorption and visible light photocatalysis. Catal. Sci. Technol. 2021, 11, 3834. [Google Scholar]
  138. Sha, X.; Huang, H.H.; Sun, S.; Huang, H.H.; Huang, Q.; He, Z.; Liu, M.; Zhou, N.; Zhang, X.; Wei, Y.J. Mussel-inspired preparation of MXene-PDA-Bi6O7 composites for efficient adsorptive removal of iodide ions. Environ. Chem. Eng. 2020, 8, 104261. [Google Scholar]
  139. Zhao, Q.; Wang, J.; Li, Z.; Guo, Y.; Tang, B.; Abudula, A.; Guan, G. Two-dimensional Ti3C2TX-nanosheets/Cu2O composite as a high-performance photocatalyst for decomposition of tetracycline. Carbon Resour. Convers. 2021, 4, 197. [Google Scholar]
  140. Nosaka, Y.; Nosaka, A.Y. Generation and detection of reactive oxygen species in photocatalysis. Chem. Rev. 2017, 117, 11302–11336. [Google Scholar] [CrossRef] [PubMed]
  141. Yin, K.; Deng, L.; Luo, J.; Crittenden, J.; Liu, C.; Wei, Y.; Wang, L. Destruction of phenicol antibiotics using the UV/H2O2 process: Kinetics, byproducts, toxicity evaluation and trichloromethane formation potential. Chem. Eng. J. 2018, 351, 867–877. [Google Scholar] [CrossRef]
  142. Iqbal, M.A.; Tariq, A.; Zaheer, A.; Gul, S.; Ali, S.I.; Iqbal, M.Z.; Akinwande, D.; Rizwan, S. Ti3C2-MXene/bismuth ferrite nanohybrids for efficient degradation of organic dyes and colorless pollutants. ACS Omega 2019, 4, 20530–20539. [Google Scholar] [CrossRef] [Green Version]
  143. Ayesha, J.; Shoomaila, L.; Muhammad, I.; Nazim, H.; Muhammad, B.; Hafiz, M.N.I. MXene-based hybrid composites as photocatalyst for the mitigation of pharmaceuticals. Chemosphere 2022, 291, 133062. [Google Scholar]
  144. Narayanam, J.M.; Stephenson, C.R. Visible light photoredox catalysis: Applications in organic synthesis. Chem. Soc. Rev. 2011, 40, 102–113. [Google Scholar] [CrossRef] [PubMed]
  145. Liu, X.; Iocozzia, J.; Wang, Y.; Cui, X.; Chen, Y.; Zhao, S.; Li, Z.; Lin, Z. Noble metal–metal oxide nanohybrids with tailored nanostructures for efficient solar energy conversion, photocatalysis and environmental remediation. Energy Environ. Sci. 2017, 10, 402–434. [Google Scholar] [CrossRef]
  146. Thomas, A.; Fischer, A.; Goettmann, F.; Antonietti, M.; Müller, J.-O.; Schlögl, R.; Carlsson, J.M. Graphitic carbon nitride materials: Variation of structure and morphology and their use as metal-free catalysts. J. Mater. Chem. 2008, 18, 4893–4908. [Google Scholar] [CrossRef] [Green Version]
  147. Masih, D.; Ma, Y.; Rohani, S. Graphitic C3N4 based noble-metal-free photocatalyst systems: A review. Appl. Catal. B Environ. 2017, 206, 556–588. [Google Scholar] [CrossRef]
  148. Sun, S.; Liang, S. Recent advances in functional mesoporous graphitic carbon nitride (mpg-C3N4) polymers. Nanoscale 2017, 9, 10544–10578. [Google Scholar] [CrossRef] [PubMed]
  149. Liu, Q.; Tan, X.; Wang, S.; Ma, F.; Znad, H.; Shen, Z.; Liu, L.; Liu, S. MXene as a non-metal charge mediator in 2D layered CdS@Ti3C2@TiO2 composites with superior Z-scheme visible light-driven photocatalytic activity. Environ. Sci. Nano 2019, 6, 3158–3169. [Google Scholar] [CrossRef]
  150. Wojciechowski, T.; Rozmysłowska-Wojciechowska, A.; Matyszczak, G.; Wrzecionek, M.; Olszyna, A.; Peter, A.; Mihaly-Cozmuta, A.; Nicula, C.; Mihaly-Cozmuta, L.; Podsiadło, S. Ti2C MXene modified with ceramic oxide and noble metal nanoparticles: Synthesis, morphostructural properties, and high photocatalytic activity. Inorg. Chem. 2019, 58, 7602–7614. [Google Scholar] [CrossRef]
  151. Jang, J.; Shahzad, A.; Woo, S.H.; Lee, D.S. Magnetic Ti3C2Tx (Mxene) for diclofenac degradation via the ultraviolet/chlorine advanced oxidation process. Environ. Res. 2020, 182, 108990. [Google Scholar] [CrossRef]
  152. Grzegórska, A.; Głuchowski, P.; Karczewski, J.; Ryl, J.; Wysocka, I.; Siuzdak, K.; Trykowski, G.; Grochowska, K.; Zielińska-Jurek, A. Enhanced photocatalytic activity of accordion-like layered Ti3C2 (MXene) coupled with Fe-modified decahedral anatase particles exposing {1 0 1} and {0 0 1} facets. Chem. Eng. J. 2021, 426, 130801. [Google Scholar] [CrossRef]
  153. 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]
  154. Tang, Y.; Hu, X.; Zhang, H.; Cao, J.; Liu, J.; Zhang, L.; Chen, X.; Chen, T. Metal-based MXene nanocomposites for catalytic applications. Nanoscale 2018, 10, 3233–3240. [Google Scholar]
  155. Wu, Y.; Jiang, H.; Wang, X.; Zhang, Y.; Yu, Y. Recent progress in synthesis, properties and applications of MXene-based nanocomposites. J. Mater. Sci. Technol. 2019, 35, 1987–2000. [Google Scholar]
  156. Zeng, X.; Yin, L.; Zhang, Y.; Chen, T.; Yu, Y. Metal-based MXene nanocomposites for antibacterial applications. J. Colloid Interface Sci. 2021, 583, 98–110. [Google Scholar]
  157. Shao, B.; Wang, J.; Liu, Z.; Zeng, G.; Tang, L.; Liang, Q.; He, Q.; Wu, T.; Liu, Y.; Yuan, X. Ti3C2Tx MXene decorated black phosphorus nanosheets with improved visible-light photocatalytic activity: Experimental and theoretical studies. J. Mater. Chem. 2020, 8, 5171–5185. [Google Scholar] [CrossRef]
  158. Li, S.; Shao, L.; Yang, Z.; Cheng, S.; Yang, C.; Liu, Y.; Xia, X. Constructing Ti3C2 MXene/ZnIn2S4 heterostructure as a Schottky catalyst for photocatalytic environmental remediation. Green Energy Environ. 2020, 7, 246–256. [Google Scholar] [CrossRef]
  159. Feng, X.; Yu, Z.; Sun, Y.; Shan, M.; Long, R.; Li, X. 3D MXene/Ag2S material as Schottky junction catalyst with stable and enhanced photocatalytic activity and photocorrosion resistance. Separ. Purif. Technol. 2021, 266, 118606. [Google Scholar] [CrossRef]
  160. Zhuge, Z.; Liu, X.; Chen, T.; Gong, Y.; Li, C.; Niu, L.; Xu, S.; Xu, X.; Alothman, Z.A.; Sun, C.Q. Highly efficient photocatalytic degradation of different hazardous contaminants by CaIn2S4-Ti3C2Tx Schottky heterojunction: An experimental and mechanism study. Chem. Eng. J. 2021, 421, 127838. [Google Scholar] [CrossRef]
Figure 1. MXene applications: actuators, EMI shielding, supercapacitor, sensors, water purification, photothermal therapy, and field-effect transistor [6].
Figure 1. MXene applications: actuators, EMI shielding, supercapacitor, sensors, water purification, photothermal therapy, and field-effect transistor [6].
Molecules 28 02495 g001
Figure 2. (a) Composition of MXene (M, A, and X represent the transition metal in green, group IIIA or IVA elements in purple, and nitrogen/carbon or both in red), (b) MAX-phase structures, and (c) MXenes [6].
Figure 2. (a) Composition of MXene (M, A, and X represent the transition metal in green, group IIIA or IVA elements in purple, and nitrogen/carbon or both in red), (b) MAX-phase structures, and (c) MXenes [6].
Molecules 28 02495 g002
Figure 3. Schematic of MXenes with different n values, such as Ti2CTx (n = 1), Ti3C2Tx (n = 2), Nb4C3Tx (n = 3), and (Mo,V)5C4Tx (n = 4). Reprinted with permission from Ref. [17].
Figure 3. Schematic of MXenes with different n values, such as Ti2CTx (n = 1), Ti3C2Tx (n = 2), Nb4C3Tx (n = 3), and (Mo,V)5C4Tx (n = 4). Reprinted with permission from Ref. [17].
Molecules 28 02495 g003
Figure 4. (a) Possible technique for fabricating of MXenes from MAX phases. Reprinted with permission from Ref. [9]. (b) A schematic for exfoliation route of MAX phases and production of MXenes. Reprinted with permission from Ref. [8].
Figure 4. (a) Possible technique for fabricating of MXenes from MAX phases. Reprinted with permission from Ref. [9]. (b) A schematic for exfoliation route of MAX phases and production of MXenes. Reprinted with permission from Ref. [8].
Molecules 28 02495 g004
Figure 5. Preliminary removal methods of contaminants from wastewater [118].
Figure 5. Preliminary removal methods of contaminants from wastewater [118].
Molecules 28 02495 g005
Figure 6. Schematic explanation of the synthesis methods of (a) g-C3N4/Ti3C2/TNTAs, (b) CFO/Ti3C2 layered heterojunctions, (c) Bi2WO6/Nb2CTxnanosheets, and (d) CeO2/MXene composites. Reprinted with permission (a) from Ref. [124], (b) from Ref. [125], (c) from Ref. [126], and (d) from Ref. [65].
Figure 6. Schematic explanation of the synthesis methods of (a) g-C3N4/Ti3C2/TNTAs, (b) CFO/Ti3C2 layered heterojunctions, (c) Bi2WO6/Nb2CTxnanosheets, and (d) CeO2/MXene composites. Reprinted with permission (a) from Ref. [124], (b) from Ref. [125], (c) from Ref. [126], and (d) from Ref. [65].
Molecules 28 02495 g006
Figure 7. Graphic representation of the synthesis processes of (a) Ti3C2 and MXene-Ti3C2/MoS2 composites; (b) Ti3C2 and heterostructure of MOFs; (c) CNTC photocatalyst by ultrasonic dispersion using an evaporation-induced self-assembly approach and combining HF etching; and (d) nZVI@Ti3C2. Reprinted with permission (a) from Ref. [127], (b) from Ref. [128], copyright (2020), (c) from Ref. [129], and (d) from Ref. [130].
Figure 7. Graphic representation of the synthesis processes of (a) Ti3C2 and MXene-Ti3C2/MoS2 composites; (b) Ti3C2 and heterostructure of MOFs; (c) CNTC photocatalyst by ultrasonic dispersion using an evaporation-induced self-assembly approach and combining HF etching; and (d) nZVI@Ti3C2. Reprinted with permission (a) from Ref. [127], (b) from Ref. [128], copyright (2020), (c) from Ref. [129], and (d) from Ref. [130].
Molecules 28 02495 g007
Figure 8. The mechanism for photocatalytic elimination of pharmaceutical wastes using MXene composite as a photocatalyst [143].
Figure 8. The mechanism for photocatalytic elimination of pharmaceutical wastes using MXene composite as a photocatalyst [143].
Molecules 28 02495 g008
Figure 9. General classification of MXene-based composites. [143].
Figure 9. General classification of MXene-based composites. [143].
Molecules 28 02495 g009
Table 1. Performance of degradation of pharmaceutical wastes from MXene nanocomposites.
Table 1. Performance of degradation of pharmaceutical wastes from MXene nanocomposites.
PhotocatalystsApplicationOptimal Degradation EfficiencyReference
CdS–Ti3C2TxHER15.4 mmol g−1·h−1[64]
CeO2/Ti3C2Removal of C22H24N2O8 + reduction of CO280.2% in 1h[65]
C–TiO2/Bi4NbO8ClMO, CIP, RhB, and 2&4-DCP>95%[66]
AgInS2/MXeneReduction of N238.8 μ⋅mol g−1·h−1[67]
Bi2WO6/Ti3C2C22H25ClN2O8 and RhB99.9% and 97% in 20 min and 60 min[68]
g-C3N4/Ti3C2HER116.2 μ mol/h/g[69]
Ag3PO4/Ti3C2C6H4N2O5 and C22H25ClN2O8>80%[70]
TiO2/C3N4/Ti3C2CO2 reduction4.39 μ·mol·g−1·h−1[71]
Ti3C2/g-C3N4Degradation of C5H5N and C4H4S80% in 3 h[72]
BiOBr/TiO2/Ti3C2TxRhB99.8%[73]
Ti3C2/Ag/Ag3VO4TC, RhB, and MB97%, 96%, and 99%[74]
CS@g-C3N4/MXRhB and MB99% and 98.5%[75]
TiO2@C/MXeneMB85.7%[76]
BPQDS/Ti3C2@TiO2MO93%[77]
MXene@Au@CdSHER17,070.43 μ mol g−1⋅h−1[78]
1T-WS2@TiO2@Ti3C2HER3409.8 μ mol g−1⋅h−1[79]
Ti3C2/TiO2/1T-MoS2HER9738 μ mol g−1⋅h−1[80]
TiO2@Ti3C2/g-C3N4C6H5NH2 and RhB99.9% and 98.2%[81]
MXene/TiO2/g-C3N4BPA, CIP, TC, and RhB66.3%, 41.8%, 63.6%, and 92.1%[82]
Ti3C2–Bi/BiOClC17H18FN3O389%[83]
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.; Din, M.A.U.; Dong, F.; Chen, D. Retrospective on Exploring MXene-Based Nanomaterials: Photocatalytic Applications. Molecules 2023, 28, 2495. https://doi.org/10.3390/molecules28062495

AMA Style

Irfan S, Khan SB, Din MAU, Dong F, Chen D. Retrospective on Exploring MXene-Based Nanomaterials: Photocatalytic Applications. Molecules. 2023; 28(6):2495. https://doi.org/10.3390/molecules28062495

Chicago/Turabian Style

Irfan, Syed, Sadaf Bashir Khan, Muhammad Aizaz Ud Din, Fan Dong, and Deliang Chen. 2023. "Retrospective on Exploring MXene-Based Nanomaterials: Photocatalytic Applications" Molecules 28, no. 6: 2495. https://doi.org/10.3390/molecules28062495

Article Metrics

Back to TopTop