Next Article in Journal
Superluminal Molecular and Nanomaterial Probes Based on Fast Ions or Electrons
Previous Article in Journal
Covalent Bonding of MXene/COF Heterojunction for Ultralong Cycling Li-Ion Battery Electrodes
Previous Article in Special Issue
The Simultaneous Detection of Dopamine and Uric Acid In Vivo Based on a 3D Reduced Graphene Oxide–MXene Composite Electrode
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Research Progress on Ti3C2Tx-Based Composite Materials in Antibacterial Field

1
Department of Stomatology, School of Stomatology and Ophthalmology, Hubei University of Science and Technology, Xianning 437100, China
2
Department of Computer Science and Technology, China Three Gorges University, Yichang 443002, China
3
Hubei Key Laboratory of Radiation Chemistry and Functional Materials, Non-Power Nuclear Technology Collaborative Innovation Center, Hubei University of Science and Technology, Xianning 437100, China
*
Authors to whom correspondence should be addressed.
These authors contributed equally to this work.
Molecules 2024, 29(12), 2902; https://doi.org/10.3390/molecules29122902
Submission received: 21 May 2024 / Revised: 13 June 2024 / Accepted: 14 June 2024 / Published: 18 June 2024
(This article belongs to the Special Issue The Way Forward in MXenes Materials)

Abstract

:
The integration of two-dimensional Ti3C2Tx nanosheets and other materials offers broader application options in the antibacterial field. Ti3C2Tx-based composites demonstrate synergistic physical, chemical, and photodynamic antibacterial activity. In this review, we aim to explore the potential of Ti3C2Tx-based composites in the fabrication of an antibiotic-free antibacterial agent with a focus on their systematic classification, manufacturing technology, and application potential. We investigate various components of Ti3C2Tx-based composites, such as metals, metal oxides, metal sulfides, organic frameworks, photosensitizers, etc. We also summarize the fabrication techniques used for preparing Ti3C2Tx-based composites, including solution mixing, chemical synthesis, layer-by-layer self-assembly, electrostatic assembly, and three-dimensional (3D) printing. The most recent developments in antibacterial application are also thoroughly discussed, with special attention to the medical, water treatment, food preservation, flexible textile, and industrial sectors. Ultimately, the future directions and opportunities are delineated, underscoring the focus of further research, such as elucidating microscopic mechanisms, achieving a balance between biocompatibility and antibacterial efficiency, and investigating effective, eco-friendly synthesis techniques combined with intelligent technology. A survey of the literature provides a comprehensive overview of the state-of-the-art developments in Ti3C2Tx-based composites and their potential applications in various fields. This comprehensive review covers the variety, preparation methods, and applications of Ti3C2Tx-based composites, drawing upon a total of 171 English-language references. Notably, 155 of these references are from the past five years, indicating significant recent progress and interest in this research area.

1. Introduction

Bacterial infections pose a significant global threat, resulting in millions of fatalities annually and imposing severe consequences on both individuals and societies worldwide [1,2]. Traditional antimicrobial agents, primarily derived from chemically modified natural compounds known as antibiotics, have historically served as our primary defense against bacterial infections [3]. However, despite the development of numerous classes of antibiotics, the widespread and prolonged use of these agents has led to the emergence and spread of drug-resistant strains, escalating the peril posed by bacterial infections to a critical level and posing a profound threat to global public health [4,5,6]. Examples of such threats include wound infections [7], disease dissemination [8], as well as water and soil contamination [9,10]. In response to this urgent need, there is an imperative to develop novel and effective antimicrobial materials.
In recent decades, researchers have tirelessly explored and uncovered a myriad of antimicrobial agents, spanning a wide spectrum from organic to inorganic materials, and even nano-scale materials [11]. Notably, environmentally friendly nano-scale antimicrobial materials with low toxicity and high safety standards have emerged as a pivotal frontier, offering remarkable antimicrobial properties and witnessing burgeoning utilization across various antimicrobial applications [12]. These materials furnish enduring antimicrobial activity and critically thwart the emergence of drug-resistant strains, heralding a revolutionary breakthrough in the realm of antimicrobial intervention [13]. Common nanomaterials with antimicrobial properties include graphene, carbon-based materials, transition metal dichalcogenides, black phosphorus, noble metal nanoparticles, metal–organic frameworks (MOFs), conjugated polymers, and MXenes [14,15,16]. Among these, MXenes stand out owing to their superior surface area, metallic conductivity, light absorption capabilities, photothermal conversion efficiency, reactive oxide species (ROS) generation, and lower cellular toxicity, rendering them exceptional candidates as nano-scale antimicrobial agents [17,18,19,20].
MXenes, first synthesized in 2011, represent a novel class of two-dimensional layered ternary carbides prepared by Naguib et al. [21]. Their chemical formula is Mn+1XnTx, where M refers to transition metals (Sc, Ti, V, Zr, Hf, Mo, Ta, and Nb), and X typically represents nitrogen (N) or carbon (C). “T” denotes surface terminations (e.g., -F, -Cl, -OH, and =O), and “x” (in Tx) indicates the number of surface terminations, defining three common structures (M2XTx, M3X2Tx, and M4X3Tx) [22]. In the medical field, Ti3C2Tx and Ti2CTx are the most widely utilized MXenes [23]. Ti3C2Tx, derived from hydrofluoric acid (HF) etching of the Al layers from the Ti3AlC2 MAX phase [24], exhibits lower cytotoxicity, a higher specific surface area, and remarkable optical properties. Its photodynamic and photothermal sterilization effects under white and near-infrared (NIR) light hold immense promise in the field of antimicrobials [25,26,27]. However, as a single semiconductor material, Ti3C2Tx still presents drawbacks, particularly in photocatalytic applications, due to its poor solar response and carrier separation capabilities [28].
In recent years, researchers have developed numerous Ti3C2Tx composites to enhance antimicrobial performance. Combining Ti3C2Tx with noble metal nanoparticles, metal oxides, MOFs, and other materials significantly enhances its bactericidal capability. While mainstream reviews predominantly summarize the synthesis of Ti3C2Tx and its application in biomedical research, water purification, and biosafety research [29,30,31,32,33], comprehensive reviews on the classification and fabrication of Ti3C2Tx antimicrobial composites as well as their applications in antimicrobial fields remain relatively scarce [34]. Thus, this review explores the systematic classification of Ti3C2Tx-based antimicrobial composites and their antibacterial efficiency, followed by a discussion of the main synthesis methods. Subsequently, we comprehensively review the latest advancements of Ti3C2Tx-based composites in antimicrobial applications across the medical, water treatment, flexible fabric, food preservation, and industrial sectors. Finally, we outline the challenges and prospects, emphasizing the need for further research to elucidate microscopic mechanisms of action, achieve a balance between biocompatibility and antibacterial efficiency, and explore efficient, environmental friendly synthesis methods combined with smart technologies. The primary objective of this review is to elucidate the types of materials, the preparation method used for Ti3C2Tx-based antibacterial materials, as well as their potential applications in the antibacterial field. In these domains, adjusting the shape and dimension of the final material, reducing the material cost, and enhancing industrial production efficiency will create the possibility of transformative applications (Figure 1).

2. Classification of the Ti3C2Tx Composite Antimicrobial Materials

With its outstanding physical and chemical properties, Ti3C2Tx holds immense promise to evolve into an even more superior and stable composite material. Currently, the main types of Ti3C2Tx composite antimicrobial materials encompass Ti3C2Tx/metal composites, Ti3C2Tx/metal oxide composites, Ti3C2Tx/metal sulfide composites, Ti3C2Tx/organic framework composites, Ti3C2Tx/antibiotic composites, Ti3C2Tx/antibiofilm agent composites, and Ti3C2Tx/photosensitizer composites. Combining Ti3C2Tx with various types of antimicrobial materials enhances its antimicrobial properties through avenues such as photodynamic effects, photothermal effects, and physical cutting actions (Table 1).
Table 1. Classification of Ti3C2Tx composite antimicrobial materials.
Table 1. Classification of Ti3C2Tx composite antimicrobial materials.
ClassificationConstituentCytocompatibilitySterilizationApplicationReferences
MetalAuThe bacteria with materials in the dark showed essentially normal bacterial morphologyIt effectively kills Staphylococcus aureus and prevents the development of biofilm.Wound antibacterial healing auxiliary material[35]
AuNot mentionedThe antibacterial efficiency is as high as 99.9%.Multifunctional flexible pressure sensor[36]
Metal oxideV2O5Not mentionedIt is an effective antibacterial agent against Gram-positive bacteria and Gram-negative bacteria.Waste disposal[37]
Metal sulfideCuSNot mentionedIt has good antibacterial activity against Escherichia coli and Staphylococcus aureus, with sterilization rates of 99.6% and 99.1%, respectively.Disinfecting and antibacterial materials[38]
Organic frameworkMOFsIn the cytotoxicity test, the L929 cell survival rates were all above 75% after 12 h of incubationAfter two rounds of bacterial inactivation and six months of indoor environment preservation, 99.9999% of Gram-negative Escherichia coli and Gram-positive Staphylococcus aureus can be eliminated.Antibacterial material[39]
AntibioticDOXNot mentionedIt has an obvious inhibitory effect on Gram-negative Escherichia coli and Gram-positive Bacillus subtilis within 5 h.Drug delivery[40]
AntibiofilmPVAThe relative viability changes of L929 cells after the co-incubation of the PVA hydrogel, Ag@M-PVA hydrogel, and Ag@M-H-PVA hydrogel extracts with L929 cells for 1 d and 3 d were above 85%The inhibition rate of Escherichia coli under 808 nm near infrared radiation increased from 51.78% to 97.5%.Tissue engineering repair material[41]
Carbon
matrix
GrapheneOnly pure functionalized graphene shows some noteworthy cytotoxic activity after 24 h at a high concentration of 200 μg mL−1; however, after 48 h, the cytotoxicity of functionalized graphene and 50% Ti3C2Tx—50% FG and 25% Ti3C2Tx—75% FG increase to 80%, 55%, and 60%, respectivelyFor Escherichia coli, at the dosage of 200 mg/mL, Ti3C2 functionalized graphene nanocomposites can completely kill bacteria.Biomedical application[42]
GO-PEINot mentionedGO-PEI/MXene folded microspheres have strong adsorption capacity, and the antibacterial effect can prolong the service life.Molecular imprinting technique[43]
Natural antibacterial materialChitosanNot mentionedThe sterilization effect is close to 100%.Health monitoring[44]

2.1. Single Ti3C2Tx

As a quintessential MXene material, Ti3C2Tx plays a pivotal role in various domains, particularly showcasing significant breakthroughs in antibacterial applications in recent years [45]. Its outstanding antibacterial efficacy primarily arises from two mechanisms: physical slicing and photocatalytic effects [46]. Moreover, the method of preparation exerts a substantial influence on its antibacterial performance. Compared to the conventional wet chemical etching method, which lacks mechanical force and yields Ti3C2Tx lacking sharp edges, Ti3C2Tx prepared via the one-step mechanical exfoliation method exhibits superior “nanoscale knife effects”. Its irregular sharp edges can disrupt bacterial cell walls, leading to enhanced bacteriostatic effects [47].
Two-dimensional Ti3C2Tx nanosheets, utilizing the strategy of tetramethylammonium hydroxide-driven intercalation and delamination, demonstrate excellent biocompatibility and synergistic photothermal–photodynamic antibacterial activity against Escherichia coli (E. coli) and Staphylococcus aureus (S. aureus). Compared to traditionally HF-exfoliated Ti3C2Tx nanosheets, photothermal–photodynamic Ti3C2Tx nanosheets exhibit significant antibacterial enhancement at concentrations of 0.5 mg/mL or 1.0 mg/mL, achieving nearly 100% bacterial inhibition due to improved biofilm disruption and leakage of bacterial contents under NIR illumination [48]. Furthermore, certain photothermal treatments applied to Ti3C2Tx bolster its antibacterial performance. For instance, photoactivated Ti3C2Tx demonstrates robust thermal effects and potent antibacterial activity, and it is capable of suppressing the expression of inflammatory factors and combating methicillin-resistant Staphylococcus aureus (MRSA) infections [49].

2.2. Ti3C2Tx/Metal Composites

Metal nanoparticles exhibit unique surface plasmon resonance effects, excellent conductivity, and heat resistance. When doped with semiconductor materials to form composites, they can significantly enhance photothermal efficiency. Upon binding with Ti3C2Tx, metal nanoparticles infiltrate the interlayer gaps, markedly enhancing the antibacterial activity of the composite. Common Ti3C2Tx/metal composites include Cu/Ti3C2Tx, Au/Ti3C2Tx, Ag/Ti3C2Tx, and Pt/Ti3C2Tx [50,51,52,53].
Au/Ti3C2Tx photothermal films can completely deactivate E. coli within 20 min, demonstrating remarkable photothermal antibacterial performance [54]. Ti3C2Tx-Au nanoparticles kill over 99% of bacteria with a low dose of nanoparticle suspension (50 μg/mL) under visible light irradiation for just 10 min [55]. Ti3C2Tx-Au nanobipyramids demonstrate broad-spectrum antibacterial properties for various bacterial strains (bactericidal rates all above 95%) within 8 min under 808 nm laser irradiation with a photothermal conversion efficiency of 50.41% [56]. Two-dimensional core–shell Ti3C2Tx@Au nanocomposites expand the infrared absorption range from NIR-I (650–950 nm) to NIR-II (1000–1350 nm) [57]. Incorporating Cu and Ag into the interlayer spacing of Ti3C2Tx also enhances antibacterial activity [50,58]. Based on electrostatic self-assembly between Cu2+ and MXene, Cu(II)@Ti3C2Tx photothermal composites embedded in hyaluronic acid hydrogels are formulated as antibacterial dressings, offering easy application to wounds of diverse shapes and ensuring long-term wound protection. Additionally, these readily prepared Cu(II)@Ti3C2Tx composites simultaneously function as a photothermal antibacterial barrier, ROS scavenger, and angiogenesis promoter, thereby expediting the healing process of infected wounds [59] (Figure 2). The anchoring of Ag nanoparticles on Ti3C2Tx and covalent organic framework (COF) composites results in exceptional antibacterial properties against S. aureus and Pseudomonas aeruginosa (P.aeruginosa). The synergistic catalytic effect of Ag and Ti3C2Tx significantly reduces the work function along the interface, while the built-in electric field between the layers drives rapid carrier migration, thereby effectively improving catalytic performance [60]. In addition to binding with individual metal particles, the combination of Ti3C2Tx with bimetallic nanoparticles significantly enhances antibacterial efficiency. AgPd or AgAu bimetallic nanocrystals, sandwiched between a Ti3C2Tx nanosheet and polydopamine (PDA) layer, possess a cage-like nanostructure, strong absorption in the NIR region, and a local surface plasmonic resonance effect. These characteristics are conducive to high catalytic efficiency and antibacterial performance. Moreover, owing to the presence of Ag ions and photothermal coupling antibacterial properties, composite nanosheets exhibit effective antibacterial activity against both Gram-negative (E. coli) and Gram-positive (S. aureus) bacteria [61,62].

2.3. Ti3C2Tx/Metal Oxide Composites

Besides its interaction with metal ions, Ti3C2Tx enhances its antibacterial efficacy through combining with metal oxides. The antibacterial performance of metal oxides is closely related to their ability to generate oxygen free radicals and their physical interaction with bacterial membranes. Facilitated by their porous structure and numerous channels, metal oxides have high-density, evenly distributed catalytic active centers, which promote the thorough penetration of small-molecule substrates, optimizing the transport and diffusion of products to achieve antibacterial effects [63]. Ti3C2Tx nanosheets, with their large surface area and excellent hydrophilic properties, are considered ideal carriers for metal oxides [64]. Common Ti3C2Tx/metal oxide composite materials include TiO2/Ti3C2Tx, Cu2O/Ti3C2Tx, ZnO/Ti3C2Tx, and various ferrite/Ti3C2Tx composites [65,66,67,68]. Among these, TiO2/Ti3C2Tx composites have been extensively studied [69].
Two-dimensional membranes constructed from in situ-generated nano TiO2/Ti3C2Tx layers exhibit excellent photocatalytic performance. Compared to non-photocatalytic Ti3C2Tx membranes, TiO2/Ti3C2Tx membranes under ultraviolet irradiation demonstrate a very high flux recovery ratio (80%) and over 95% resistance to E. coli [70]. The presence of nano TiO2 particles as interlayer support layers widens the interlayer channels, enhances membrane permeability, and improves self-cleaning properties and long-term membrane operational stability. Combining the antibacterial properties of copper, a multifunctional Cu2O@Ti3C2Tx/α-tricalcium phosphate scaffold with internal and external sandwiching constructed via 3D printing utilizes the excellent photothermal properties of Ti3C2Tx to control programmed temperature. This includes brief PTT to rapidly eliminate early bacteria and periodic low photothermal stimulation to promote bone tissue growth [71]. Simultaneously, the slow release of endogenous copper ions strengthens antibacterial efficiency, promotes angiogenesis, and improves the repair effect. ZnO/Ti3C2Tx heterojunctions synthesized by the hydrothermal growth of ZnO on the surface of lamellar Ti3C2Tx exhibit superior photothermal performance compared to ZnO and Ti3C2Tx alone [72]. Local thermal effects not only disrupt the bacterial membrane integrity but also promote the release of Zn2+ ions, further enhancing antibacterial performance with a sterilization rate of 100% at a 150 μg·mL(-1) concentration (superior to either ZnO or Ti3C2Tx). CuFe2O4-Ti3C2Tx antibacterial heterojunctions, constructed by in situ growth, produce high heat and exogenous ROS under NIR light irradiation [73]. Additionally, they consume glutathione in bacteria through Fenton-like reactions and generate endogenous ROS, resulting in a robust and lasting antibacterial cascade effect. Furthermore, novel magnetic Ti3C2Tx@Fe3O4/Au/PDA nanosheets exhibit excellent photothermal-magnetic decoupling antibacterial performance [74]. The antibacterial effect against E. coli and S. aureus at a concentration of 120 μg/mL approaches 100%.
TiO2/Ti3C2Tx membranes excel in photocatalytic performance and water flux recovery for water treatment applications, while the Cu2O@Ti3C2Tx/α-tricalcium phosphate scaffold leverages its multifunctional properties to promote bone tissue growth and enhance antibacterial efficiency for medical implants. ZnO/Ti3C2Tx heterojunctions demonstrate superior photothermal performance and antibacterial action through the combined effects of local thermal effects and Zn2+ ion release. CuFe2O4-Ti3C2Tx antibacterial heterojunctions, on the other hand, produce high heat and exogenous ROS under NIR irradiation, offering another effective antibacterial solution. Each material exhibits unique advantages tailored for specific applications, whether it be water treatment, medical implants, or antibacterial applications.

2.4. Ti3C2Tx/Metal Sulfide Composites

Given the escalating issue of antibiotic resistance, environmentally friendly photoelectric materials hold promise as alternatives to antibiotics. Metal sulfides, characterized by excellent biocompatibility, a low cost, and particularly noteworthy photothermal properties, have garnered special attention. They possess economic viability, chemical stability, and the ability to easily couple with other semiconductors [75]. For instance, the interface Schottky heterojunction designed by the contact potential difference between Bi2S3 and Ti3C2Tx inhibits electron backflow and boosts charge transfer and separation, thereby intensively improving the amount of ROS under 808 nm NIR radiation [76]. This enables the material to achieve a bactericidal rate of 99.86% of S. aureus and 99.92% of E. coli with the assistance of hyperthermia within 10 min. BiOI@Bi2S3/Ti3C2Tx composite materials synthesized through the in situ vulcanization of two-dimensional Ti3C2Tx with hollow flower-shaped BiOI achieve a photothermal conversion efficiency of 57.8% under 808 nm laser irradiation, thereby boosting PTT antibacterial effects [77]. Meanwhile, the close contact of heterojunction interfaces enhances the charge transfer and suppresses electron–hole recombination, improving the antibacterial performance of PDT. The combined PTT and PDT antibacterial actions result in efficiencies of 99.7% and 99.8% against P. aeruginosa and S. aureus, respectively. Ti3C2Tx/ZnIn2S4 heterojunctions achieve controllable electromagnetic properties through the dual optimization of intercalated nano interfaces of Ti3C2Tx and an S vacancy structure design, exhibiting strong antibacterial activity [78]. MoS2 demonstrates excellent photocatalytic performance and infrared photothermal effects, making it suitable for various antibacterial pathways such as photocatalytic antibacterial, enzyme-like catalytic antibacterial, physical antibacterial, and photothermal-assisted antibacterial pathways [79]. For example, Ti3C2Tx/MoS2 heterojunctions induce local heating and increase extracellular ROS levels, resulting in bacterial inactivation under NIR laser irradiation [80]. Meanwhile, Mo4+ ions readily invade ruptured bacterial cell membranes, inducing intracellular ROS and depleting intracellular glutathione. Under the pressure of “ROS hurricanes” from both internal and external sides, bacteria are hugely slaughtered. When MoS2 grows in a vertical arrangement on Ti3C2Tx, it damages the peptidoglycan mesh in the bacterial wall, leading to significant morphological changes and thus bacteriostatic effects [81]. In addition, CuS nanoparticles also can rapidly convert light energy into heat under NIR light irradiation, making them suitable for photothermal sterilization [82]. Accordion-like Ti3C2Tx@CuS structures enhance the separation efficiency of CuS electron–hole pairs, generating more ROS to kill bacteria [83]. Moreover, with the synergistic effect of sustained copper ion release, they maintain over 90% long-term sterilization ability within 30 days. Furthermore, FeS@lauramidopropyl betaine@Ti3C2Tx nanoenzymes exhibit catalytic activity similar to peroxidases, promoting the production of hydroxyl radicals by catalyzing the decomposition of H2O2 [84]. They demonstrate excellent antibacterial activity against typical Gram-negative E. coli and Gram-positive S. aureus, integrating chemical kinetics and PTT to provide an effective strategy for bacterial inhibition and wound healing.

2.5. Ti3C2Tx/Organic Frameworks Composites

Organic frameworks refer to multidimensional network structures composed of organic molecules or polymers. These materials exhibit highly ordered microstructures and offer a high degree of tunability in terms of performance. Currently, in the field of antibacterial materials, the main organic frameworks that are being combined with Ti3C2Tx are MOFs and COFs.
MOFs are a class of emerging hybrid porous materials composed of clusters of metal ions or clusters connected by organic linkers. They possess advantages such as a high surface area, a tunable structure, chemical functionality, and host–guest interactions, making them widely applicable in various fields [85]. MOFs have attracted considerable attention regarding antibacterial potential as they select diverse metal ions and organic ligands to control their outstanding physical and chemical properties. For instance, utilizing Ti3C2Tx as a conductive carrier and UiO66-NH2 (C48H30N6O32Zr6) as positively charged nano-MOFs, Ti3C2Tx-based Schottky heterojunctions exhibit inhibition rates greater than 95% against E. coli and S. aureus under visible light irradiation [86]. During the formation of the Schottky heterojunction, UiO66-NH2 nanoparticles grow in situ on the surface and inner surface of Ti3C2Tx, not only combining the antibacterial properties of Ti3C2Tx but also effectively reducing its bandgap energy, thereby enhancing its photocatalytic activity. Smart photothermochromic self-disinfecting textiles prepared through the in situ growth of porphyrin-type MOF (PCN-224, C144H112N12O64Zr15) and the electrospray coating of Ti3C2Tx colloid inactivate 99.9999% of E. coli and S. aureus within 30 min. Loading nano-silver on the basis of PCN-224/Ti3C2Tx not only synergistically enhances the antibacterial effect through photocatalytic single-state oxygen generated by PCN-224 and the heat generated by Ti3C2Tx, but also promotes the degradation of nano-silver to release silver ions, achieving rapid bacterial inactivation and persistent bacterial inhibition [87] (Figure 3).
COFs are porous crystalline materials formed by covalently connecting organic molecules. They possess advantages such as tunable bandgaps and large surface areas. However, compared to traditional photocatalytic bactericides, the antibacterial effect of COFs is not ideal, mainly due to the low separation efficiency of electron–hole pairs. An effective solution is to modify the surface of COFs to enhance the separation efficiency and migration rate of photo-generated charge carriers [88]. The in situ growth of TpPa-1-COF (C15H14N2O6) on Ti3C2Tx through Schiff base reaction improves the antibacterial rates to 98.90% and 99.62%, respectively, against S. aureus and P. aeruginosa, which are 50% and 33% higher than those of pure TpPa-1-COF [89]. Electrochemical impedance spectroscopy and photoluminescence spectra indicate that the migration rate of photo-generated carriers between layers is increased, thereby achieving efficient photocatalysis and reducing photocorrosion.

2.6. Ti3C2Tx/Antibiotic Composites

Antibiotics, primarily secondary metabolites produced by bacteria, fungi, or other microorganisms, or synthetic analogs thereof, exert their bactericidal effect by targeting specific processes in bacterial cell DNA, RNA, and protein synthesis, thereby disrupting normal bacterial metabolism and inhibiting bacterial growth [90]. However, the overuse of antibiotics can lead to the emergence of antibiotic-resistant strains [91,92].
Combining Ti3C2Tx with antibiotics not only enhances antibacterial efficiency, but also effectively reduces the use of antibiotics. Common antibiotics, such as amoxicillin, ciprofloxacin, and tobramycin, are often employed in such composite materials. For instance, electrospun nanofiber membranes produced by loading amoxicillin onto a Ti3C2Tx and polyvinyl alcohol mixture achieve a killing efficiency rate of 96.1% against S. aureus upon laser irradiation, while the effective bactericidal rate is 99.1% when combined with an NIR laser [93]. These composite nanofiber membranes not only control the release of amoxicillin to combat bacterial infections, but also utilize Ti3C2Tx to convert NIR laser into heat, thereby destroying the non-cellular components of bacteria and synergistically leading to bacterial inactivation. Similarly, the combination of ciprofloxacin and Ti3C2Tx into nanocomposite materials achieve an outstanding in vitro bactericidal effect of >99.99999% against MRSA [45]. This composite material demonstrates long-term inhibition in a mouse MRSA-induced abscess model, avoiding bacterial rebound post PTT. Ciprofloxacin-Ti3C2Tx nanocomposite materials can be prepared as sprayable hydrogels, offering excellent photothermal conversion capability and biocompatibility, ultimately enhancing antibacterial activity [94] (Figure 4). Multifunctional implants composed of tobramycin and Ti3C2Tx nanosheets exhibit robust antibacterial properties against both Gram-negative and Gram-positive bacteria [95]. Importantly, these implants demonstrated excellent cellular compatibility and osteogenic potential, promoting bone cell proliferation, diffusion, alkaline phosphatase activity, calcium matrix mineralization, and in vivo bone integration. In addition, encapsulating tick-derived antimicrobial peptides as model drugs in hydrogels with Ti3C2Tx nanoparticles resulted in exceptional antibacterial properties, excellent biocompatibility, and superior skin tissue regeneration capabilities, offering broad applications in wound healing [96].

2.7. Ti3C2Tx/Antibiofilm Composites

Antibiofilm agents refer to compounds and drugs that inhibit the formation or disrupt the structure of bacterial biofilms. Common antibiofilm agents include PDA, polyethyleneimine, polycaprolactone, and others.
PDA is renowned for its excellent photothermal properties. When PDA is combined with Ti3C2Tx, its photothermal effects and free radical scavenging capabilities are enhanced, and it can even produce Fenton-like activity, exhibiting promising antibacterial and anti-inflammatory strategies [97,98]. Coating in situ-grown Ti3C2Tx@gold nanorods with PDA results in synergistic photodynamic antibacterial activity [99]. Composite nanoparticles generate more 1O2 under 660 nm laser irradiation with the increase in the PDA content, reducing the minimum inhibitory concentration against E. coli and S. aureus to approximately 0.07 mg/mL. Additionally, the PDA coating further enhances the antioxidative and antibacterial abilities of Ti3C2Tx, providing the scaffold with excellent ROS scavenging capacity [100] and tissue adhesion properties [101], thereby promoting wound healing in diabetic infection wounds [102] and wounds infected with multidrug-resistant bacteria [103] (Figure 5). Moreover, PDA coating can passivate the sharp edges of Ti3C2Tx nanosheets, preventing damage to normal tissue cells [104].
Polyethyleneimine and Ti3C2Tx-modified sodium alginate aerogels also exhibit significantly reinforced mechanical strength, achieving a 99.99% antibacterial rate against S. aureus and E. coli within 2 h [105]. Polycaprolactone is commonly used to prepare electrospun nanofiber membranes with biocompatibility and biodegradability [106]. The presence of Ti3C2Tx helps reduce the diameter of electrospun nanofibers, significantly enhancing the hydrophilicity, electroactivity, and antibacterial activity of membranes [107]. Furthermore, this nanofiber membrane significantly promotes the adhesion, proliferation, and migration of NIH 3T3 cells under electrical stimulation, thereby accelerating wound closure, increasing granulation tissue formation, collagen deposition, and wound vascularization, making it a promising multifunctional wound dressing [108].
In summary, the combination of Ti3C2Tx with antibiofilm agents offers a versatile approach for enhancing antibacterial activity, promoting wound healing, and preventing biofilm formation.

2.8. Ti3C2Tx/Photosensitizer Composites

Photosensitizers are compounds capable of absorbing specific wavelengths of light energy and undergoing photodynamic reactions to produce ROS, which play a crucial role in PDT. Common photosensitizers include indocyanine green (ICG), ruthenium (Ru), and zinc phthalocyanine (ZnTCPP). Loading ICG onto Ti3C2Tx nanosheets combines the photothermal effect of Ti3C2Tx with the photodynamic effect of ICG, resulting in a 100% increase in the vitality loss of S. aureus under NIR light spectrum [109]. Ti3C2Tx nanocomposites based on Ru(II) complex grafting achieve almost 100% bactericidal activity against E. coli (200 μg/mL) and S. aureus (100 μg/mL) upon exposure to a xenon lamp [110] (Figure 6). ZnTCPP-modified Ti3C2Tx captures bacteria via surface electrostatic interactions [111]. The Schottky junction formed between Ti3C2Tx and ZnTCPP promotes visible light utilization, accelerates charge separation, enhances the migration rate of photogenerated charges, and ultimately improves photocatalytic activity. This visible light-responsive organic–inorganic hybrid achieves antibacterial efficiencies of 99.86% and 99.92% against S. aureus and E. coli, respectively, within 10 min. Porphyrin MOFs combined with Ti3C2Tx not only kill drug-resistant bacteria by generating abundant ROS but also promote the proliferation of stem cells by regulating the cell cycle, DNA replication, and apoptosis [112]. They promote osteogenic differentiation through key signaling pathways, including calcium, Wnt, and TGF-beta signaling pathways. A sandwich structure composed of TiO2, Ti3C2Tx and a photosensitizer (PANi (Ni-phytate)) achieves the bidirectional transfer of hot electrons under 808 nm NIR light irradiation, elevating the photothermal conversion efficiency to 43.3% and realizing a bactericidal effect of 99.9% against E. coli [113]. This study presents a novel approach to construct NIR light-induced antibacterial materials, realizing the synergistic effect of photocatalytic therapy and PTT.
With the increasing demand for antibacterial materials, Ti3C2Tx-based antibacterial materials are rapidly evolving from singular components towards composite materials. This includes using Ti3C2Tx as a carrier to load various antibacterial drugs or directly incorporating it into 3D networks, such as antibacterial gels or films. Through this approach, the advantages of different antibacterial substances can be combined to achieve a broader antibacterial spectrum and more durable antibacterial effects. In recent years, utilizing the photocatalytic properties of Ti3C2Tx for photothermal treatment has emerged as a new trend in developing antibacterial effects, especially by combining Ti3C2Tx with other photocatalytic materials to enhance antibacterial performance through increased photocatalytic activity. This evolution demonstrates diversification strategies in the development of Ti3C2Tx-based antibacterial materials, where wise material selection and combination offer new possibilities for developing efficient antibacterial materials.

3. Preparation Methods

The preparation methods for Ti3C2Tx antibacterial composite materials are diverse and can be broadly categorized into the following approaches: the solution mixing method, chemical synthesis method, layer-by-layer self-assembly method, electrostatic assembly method, and 3D printing. These methods offer various ways to combine Ti3C2Tx with other materials to enhance its antibacterial properties. The choice of preparation method significantly influences the structure and performance of Ti3C2Tx antibacterial composite materials (Table 2).
Table 2. Preparation methods of Ti3C2Tx-based antibacterial composite materials.
Table 2. Preparation methods of Ti3C2Tx-based antibacterial composite materials.
ClassificationConstituentSterilizationApplicationReference
Solution mixingCationic polymer poly-L-lysine, Ti3C2At the concentration of 200 mg L−1, the number of living cells of E. coli. decreased by two orders of magnitude.Biotechnology or nanomedicine[114]
Amidoxime group, Ti3C2, grafted polyamideExcellent anti-biological pollution performance (92.9% antibacterial rate).Extraction of uranium from seawater[115]
Ti3C2, Al2O3, Ag, Cu A total of 99.6% bacteria can be collected in the filter.Sewage disposal[116]
Ti3C2, tanninRemarkable antibacterial properties against E. coli and S. aureus.Intelligent wearable device[117]
Chemical synthesisBiOI, CeO2, Ti3C2The photocatalytic bacteriostatic efficiency of E. coli and S. aureus were 99.76% and 99.89% respectively.Double static electricity and antifouling of seawater[118]
Ti3C2, Al2O3, Ag, SiO2, PdAntibacterial properties.Ecotoxicology[119]
Ag, Ti3C2, Cu2OExcellent antibacterial activity against P. aeruginosa and S.aureus.Antifouling[120]
TiO2, W18O49Z, Ti3C2The photocatalytic sterilization rate of E. coil under illumination is 93.7% (λ ≥ 365 nm, 6 h).Teleportation[121]
Layer-by-layer self-assemblyTi3C2, dimethyl siloxaneA strong antibacterial effect on S. aureus and E. coli (≥99% adhered bacterial cells).Medical implantation[122]
Electrostatic assemblyGraphite carbonitride, Ti3C2In complex water matrix, 7 × 107 CFU/mL of S. aureus can be completely inactivated.Natural water disinfection[123]
Fe2O3, Ti3C2, glucose oxidaseSimultaneously promote Fe2+/Fe3+ to penetrate into bacteria and causes planktonic bacteria to die.Fight stubborn infection[124]
3D printing PDA, Ti3C2, CFPEEK Effectively kill bacteria after 10 min of NIR irradiation at 808 nm. The antibacterial rate reached 100%.Orthopedic implant[125]

3.1. Solution Mixing Method

In this approach, Ti3C2Tx is dispersed in a solution containing other antibacterial agents or materials. Through mixing and subsequent processing, a homogeneous composite material is obtained. This method allows for precise control over the composition and structure of the resulting composite, thereby enabling tailored antibacterial properties. With its advantages of simplicity, low cost, convenience, stability of solution properties, and controllability, the solution mixing method holds great promise for antibacterial applications. Through the solution mixing method, the most biocompatible Ti3C2Tx/metal composites can be selected by adjusting the mass ratio of the precursor [50]. Metal nanoclusters integrated with Ti3C2Tx via the solution mixing method induce bacterial membrane damage, instigating the generation of high concentrations of localized ROS. This process effectively oxidizes bacterial membrane lipids, amplifies membrane rupture, and induces severe bacterial DNA fragmentation, ultimately culminating in the demise of Gram-positive and Gram-negative bacteria through a synergistic physical and chemical antibacterial mechanism [126]. Utilizing the solution mixing method, metal oxide CaO2 nanoparticles can also be effectively loaded onto the surface of amino-functionalized Ti3C2Tx, yielding Ti3C2Tx composite nanosheets endowed with sonochemical/chemical dynamic capability. When subjected to ultrasound treatment for just one hour, these nanosheets prompt the shrinkage, rupture, or complete dissolution of E. coli and S. aureus [127]. This effect arises from the generation of acoustic sensitizer TiO2 on a Ti3C2Tx oxidated surface facilitated by CaO2. This, in turn, promotes the Fenton reaction initiated by self-supplied H2O2, thereby augmenting the production of ROS. Furthermore, the solution mixing method can hybridize two-dimensional Ti3C2 sheets into a silane film on an aluminum alloy surface. This hybridization exhibits excellent adhesion to the substrate via Si-O-Al chemical bonds, significantly enhancing the thickness and compactness of the silane film, thereby inhibiting both anode and cathode corrosions. Meanwhile, Ti3C2Tx within the film demonstrates excellent antibacterial properties, as evidenced by PI staining, revealing that 90% of E. coli were dyed red after 24 h of incubation [128]. In addition, incorporating different mass ratios of Ti3C2Tx nanosheets into epoxy resin through solution mixing and direct curing results in Ti3C2Tx/resin composite materials. These materials not only exhibit improved mechanical and abrasive properties compared to pure resin but also manifest increased antibacterial activity with Ti3C2Tx loading, potentially reducing secondary caries caused by bacterial accumulation at the margins of resin composite materials [129]. In the same way, the construction of composite membranes by solution mixing Ti3C2Tx with other two-dimensional materials enhances the permeability and selectivity of substrate membranes, further intensifying antibacterial activity against E. coli and S. aureus [130] (Figure 7).

3.2. Chemical Synthesis Method

Chemical synthesis methods involve blending Ti3C2Tx nanosheets or particles with other materials through chemical reactions to form composite materials. The abundant hydrophilic functional groups on the surface of Ti3C2Tx facilitate tight connections with various semiconductors and strong interaction with water molecules. Common chemical synthesis methods include solvent thermal synthesis and in situ synthesis. Leveraging the controllability, precision, and flexibility of this method enables composite materials with enhanced stability and catalytic performance to be synthesized. For instance, employing the simple wet chemistry method to grow Ag2S nanoparticles in situ on the surface of two-dimensional Ti3C2Tx leads to a decrease in the bandgap of the sample with an increase in the Ti3C2Tx content, which subsequently results in a redshift in light absorption and increased light absorption. With the combined action of Ti3C2Tx/Ag2S and 808 nm NIR light irradiation, the surfaces of S. aureus exhibit roughness and varying degrees of pits and damage. Thus, the synergistic antibacterial rate reaches as high as 99.99% [131]. Lattice-strain-rich Ti3C2Tx synthesized utilizing the solvent thermal method greatly enhances the sonochemical effect, thereby realizing extraordinary sonodynamic therapy. The intervention of meso-tetra(4-carboxyphenyl) porphyrin disrupts all Ti-O and most Ti-F chemical bonds on the surface of Ti3C2Tx. Recombination between these exposed Ti atoms with amino groups perturbs the order of Ti atoms and leads to Ti atom displacement, ultimately resulting in Ti3C2Tx lattice distortion. This inherent lattice strain narrows the band gap of Ti3C2Tx, facilitating the separation and transfer of electron–hole pairs, the generation of ultrasound-mediated ROS, and subsequent robust antibacterial ability against MRSA (99.77 ± 0.16%) [132] (Figure 7). Moreover, a novel Ti3C2Tx-based Schottky junction prepared via the solvent thermal method effectively reduces its band gap energy and enhances the photocatalytic activity of Ti3C2Tx, achieving inhibitory rates greater than 95% against E. coli and S. aureus under visible light irradiation [86]. Additionally, a ternary composite material prepared by covalent connection and Schif base reaction exhibits good antibacterial properties against S. aureus and P. aeruginosa. The covalent bonding between Ti3C2Tx and COFs greatly enhances stability, while the electron transfer channel formed between the ternary materials significantly improves the efficiency of carrier separation, prolonging the lifetime of photogenerated carriers. Furthermore, the slow release of Ag+ from a ternary material through electrostatic action maintains long-term antibacterial properties [60].

3.3. Layer-by-Layer Self-Assembly Method

Layer-by-layer (LBL) self-assembly stands as a technique employed to fabricate multilayer structures at both nanoscale and microscale levels. Typically, it entails the alternate deposition of Ti3C2Tx and other diverse materials onto a substrate, thereby yielding multilayer films with tailored properties. For example, multilayered crumpled Ti3C2Tx coatings with sharp-edged peaks via LBL self-assembly exhibit on-demand dynamic deformation, and they are capable of removing over 99% of adherent bacterial cells and creating a self-cleaning surface with restored functionality [122] (Figure 7). A multilayer Ti3C2Tx/chitosan/Ag coating formed via the self-assembly approach demonstrated outstanding antimicrobial properties, effectively reducing both Gram-negative bacteria (P. aeruginosa) and Gram-positive bacteria (S. aureus) by 99.9% [133]. Additionally, a poly(L-lactic acid) membrane assembled with positively charged chitosan and negatively charged Ag@Ti3C2Tx exhibited a synergistic antimicrobial effect, resulting in growth inhibition rates of 91.27% against E. coli and 96.11% against S. aureus [134]. A sandwich-structured superhydrophobic composite film created using superhydrophobic Ti3C2Tx and nanocellulose as raw materials via the LBL self-assembly method demonstrated excellent self-cleaning ability, water repellency, and durability, as well as high photothermal conversion capability and stability. This composite film successfully achieves controllable photo-driven motion and enhanced antimicrobial performance through the simultaneous integration of superhydrophobicity, anti-adhesion, and sustained photothermal sterilization properties [135].

3.4. Electrostatic Assembly Method

Electrostatic assembly techniques, notably electrospinning and electrostatic self-assembly, have emerged as pivotal methods for fabricating Ti3C2Tx antibacterial composite materials. Electrospinning technology facilitates the uniform dispersion of Ti3C2Tx nanofibers in a polymer matrix, yielding composite materials with an elevated surface area and robust mechanical properties. Immobilizing Ti3C2Tx on electrospun polycaprolactone membranes [106] or the formation of hybrid Ti3C2Tx-based biointerfaces [136,137] makes infected wound microenvironments undergo a transformative shift towards regenerative states, characterized by bacterial eradication, hemostasis, enhanced epithelialization, collagen deposition, and angiogenesis upon NIR light irradiation. Coating Ti3C2Tx/Ag [138] or self-assembling Ag@Ti3C2Tx hybrids [139] on other electrospun scaffolds also manifests significant antibacterial effects. Notably, functionalized electrospun nanofiber membranes exhibit high photothermal effects, omnidirectional light harvesting, rapid evaporation rates, and strong antibacterial properties, achieving a notable antibacterial efficiency of 99.9% against E. coli after 24 h of contact [140]. Furthermore, wound dressings incorporating delaminated Ti3C2Tx flakes within chitosan nanofibers demonstrate substantial reductions in colony forming units of both Gram-negative E. coli and Gram-positive S. aureus [141] (Figure 7). Ti3C2Tx/zeolite imidazole framework-8 (ZIF-8)/polylactic acid composite membranes exhibit superior PTT and PDT performance, with antibacterial rates of 99.9% and 99.8% against E. coli and MRSA, respectively, offering a promising avenue for combating infections by extremely drug-resistant bacteria [142]. Moreover, Ti3C2Tx-based electrospun fiber membranes promote the repair and regeneration of peripheral nerve injuries by restoring electrical signal transmission and regulating neuronal membrane function. The addition of Ti3C2Tx enhances the hydrophilicity, conductivity, and biocompatibility of membranes, thereby facilitating cell growth and exhibiting commendable antibacterial properties (as high as 80%) [143]. Piezoelectric stimulation under external ultrasonication augments the growth and proliferation of Schwann cells cultured on these electrospun scaffolds, fostering axonal elongation and myelin sheath formation and ultimately restoring motor and sensory functions in regenerated nerve rats [144].
The electrostatic self-assembly technique involves the sequential deposition of negatively charged Ti3C2Tx and positively charged functional materials, typically chitosan [145] and its derivatives [146] (Figure 7), through electrostatic interactions to form uniform and stable multilayered ultra-thin composite materials. This self-assembly strategy enhances the dispersion of Ti3C2Tx nanosheets, establishing a highly interconnected 3D network that facilitates rapid electron transport pathways. The incorporation of chitosan and Ti3C2Tx confers multifunctionality to the material, including excellent conductivity, sensitivity, mechanical strength (up to 1900%), and flexibility [147]. Except for chitosan, negatively charged Ti3C2Tx can form multifunctional nanomaterials with positively charged metal ions for photothermal antibacterial applications [56]. For example, the integration of Cu2+ and Ti3C2Tx into hyaluronic acid hydrogels yields antibacterial dressings with rapid adhesion, self-healing, and inject ability, which can conveniently be applied to wounds of various shapes, providing long-term wound protection. In these antibacterial dressings, the photothermal complexes act as antibacterial barriers, ROS scavengers, and vascular regeneration promoters to accelerate the healing process of infected wounds, ameliorate inflammation, enhance collagen deposition, foster vascular formation, and ultimately facilitate wound closure [59]. Additionally, antibacterial textiles with the adsorption of cationic Bi3+ onto Ti3C2Tx nanosheet surfaces exhibit antibacterial efficacy rates of 98.64% and 99.89% against S. aureus and E. coli, respectively, offering a novel approach for designing light-excited antibacterial textiles [148].

3.5. Three-Dimensional Printing

Three-dimensional printing, as a transformative technology for the on-demand fabrication of three-dimensional objects, offers unparalleled advantages in manufacturing complex-shaped and structured items by translating digital designs into tangible objects through layer-by-layer stacking. The integration of Ti3C2Tx with other materials using 3D printing technology has been widely applied in bone tissue engineering [149]. Incorporating Ti3C2Tx into 3D scaffolds enhances the mechanical properties, antibacterial performance, and osteogenicity, showing significant potential for repairing infected bone defects [150]. Multifunctional scaffolds with a sandwich-like structure of Cu2O@Ti3C2Tx/α-tricalcium phosphate are directly fabricated through 3D printing to achieve programmable temperature control. These scaffolds support brief PTT to swiftly eliminate early bacteria and periodic low-level photothermal stimulation to stimulate bone tissue growth, thereby minimizing damage to healthy cells and tissues. Meanwhile, endogenous copper ions are gradually released inside the scaffolds within a safe dosage range, enhancing the antibacterial effects of early PTT and promoting angiogenesis, ultimately improving the repair outcomes [71]. Personalized Ti3C2Tx composite 3D bio-printed hydrogels with rat bone marrow mesenchymal stem cells have both photothermal antibacterial and osteogenic capabilities to accelerate the healing and bone regeneration of rat mandibular defects infected with S. aureus [151]. Poly (lactic acid) tracheal stents uniformly incorporated with CuFe2O4-Ti3C2Tx heterojunction via selective laser technology deplete glutathione in bacteria by generating high heat and exogenous ROS under NIR light exposure. Furthermore, ROS are continuously produced through a Fenton-like reaction, achieving a robust and persistent antibacterial cascade effect. After only 15 min of NIR light exposure, 96.49% of S. aureus and 95.33% of P. aeruginosa are eradicated. These heterojunction tracheal stents offer potential prospects for patients with tracheal injuries with the aid of CuFe2O4-Ti3C2Tx heterojunction [73] (Figure 7). Three-dimensional printing technology has opened up new directions in material design, particularly offering significant advantages in the fabrication of items with complex shapes and structures, and it holds limitless potential for various applications.
At present, these are the only common synthesis methods for Ti3C2Tx composite antibacterial materials (Figure 7), but it is believed that with the development of science and continuous in-depth research by researchers, more synthesis methods will be found to cope with more scenarios and broaden the application prospect of Ti3C2Tx composite antibacterial materials.
The preparation methods for Ti3C2Tx-based composite antibacterial materials have evolved from simple mixing to chemical reactions, self-assembly, and finally, customization through 3D printing. In particular, 3D printing technology enables the direct printing of Ti3C2Tx with polymers into composite structures, achieving precise control over the structure and performance of antibacterial materials, providing greater flexibility and customization for their applications, and offering insights and inspiration for the design and preparation of future antibacterial materials.
Figure 7. Synthesis methods of Ti3C2Tx composite material [73,122,130,132,141,146]. (A) Solution mixing method. (B) Chemical synthesis method. (C) Layer-by-layer self-assembly. (D) Three-dimensional printing method. (E) (a) Electrostatic spinning method. (b) Electrostatic self-assembly.
Figure 7. Synthesis methods of Ti3C2Tx composite material [73,122,130,132,141,146]. (A) Solution mixing method. (B) Chemical synthesis method. (C) Layer-by-layer self-assembly. (D) Three-dimensional printing method. (E) (a) Electrostatic spinning method. (b) Electrostatic self-assembly.
Molecules 29 02902 g007

4. Application of Ti3C2Tx Composite in Antibacterial Field

With its unique physical structure and chemical characteristics, Ti3C2Tx has a broad application prospect in the antibacterial field, including in the medicine, water treatment, textile, and food industries and other aspects (Figure 8).

4.1. Medical Field

With the continuous progress of medicine, there are more and more requirements for antibacterial properties. As a new nanomaterial, Ti3C2Tx shows great potential in the field of medical antibacterial materials.

4.1.1. Wound Healing

Ti3C2Tx exhibits the ability to prevent bacterial resistance by inducing bacterial death through mechanical damage and oxidative stress. Moreover, Ti3C2Tx, which is rich in functional groups, has great potential for modification with molecules such as proteins, growth factors, and nanoparticles. These properties have attracted more and more attention from researchers, who are ready to apply them in the modification and construction of wound dressings against multidrug-resistant bacteria [155]. For example, immobilizing lysozyme on the hybrid interface of Ti3C2Tx has been shown to enable the precise control of local heat and enhance the photothermal response of loaded lysozyme biocatalysis, effectively inhibiting the proliferation of MRSA and accelerating the disinfection of mouse wounds [156]. This enhancement is mainly attributed to the photothermally enhanced lysozyme activity, along with the bacterial death caused by local mild hyperthermia and physical destruction. The excellent photothermal effect of Ti3C2Tx provides new design and application ideas for skin wound dressing. During wound healing, Ti3C2Tx increases the material temperature to 40 °C or higher, significantly inhibiting bacterial growth activity. A multifunctional polyvinylidene fluoride (PVDF) membrane loaded with chitosan–Ti3C2Tx suspension exhibits an inhibitory ability close to 100% under NIR irradiation. Animal experiment results demonstrate that the wound healing rate reaches 95% with the aid of NIR irradiation for 14 days, which is markedly higher than that of the group without NIR irradiation [145]. Moreover, the surface charges of Ti3C2Tx promote blood cell aggregation, platelet activation, and clot formation. The novel composite membrane containing graphene oxide and Ti3C2Tx significantly increases tensile strength, platelet adsorption, and the blood clotting time in addition to having a 97% antibacterial rate against both Gram-positive and Gram-negative bacteria [157]. Additionally, multifunctional hydrogels based on regenerated bacterial cellulose and Ti3C2Tx have good mechanical properties, flexibility, biodegradability, and high water absorption capacity. These hydrogels promote skin wound healing through the electrical stimulation of skin cell behavior with the help of the electrical conductivity of Ti3C2Tx. These hydrogels significantly enhance the proliferation activity of NIH3T3 cells and accelerate the wound healing process when coupled with electrical stimulation, providing an effective synergistic treatment strategy and proving that they are attractive candidates for wound dressings [93]. Furthermore, since abundant functional groups on the surface of Ti3C2Tx can react with ROS, the addition of Ti3C2Tx to skin wound dressings can improve the overall ROS scavenging ability of the dressings [158]. In summary, Ti3C2Tx-based composite antimicrobial materials, with their excellent biocompatibility, photothermal effect, and electrical conductivity, effectively reduce the risks of bacterial adhesion and infection, improving wound healing and patient recovery times.

4.1.2. Medical Sensor

Ti3C2Tx composite antibacterial materials offer significant advantages in the development of electrochemiluminescence sensors [159], flexible electronic devices [160], and wearable skin sensors [161] due to their outstanding electrochemical and antibacterial properties. Sensors constructed from Ti3C2Tx composites have the capability to monitor human movements in real time, encompassing large deformations (e.g., finger, elbow, wrist, and knee bending) and small deformations (such as mouth movements and throat sounds) [52]. These sensors are applied in real-time movement detection in amyotrophic lateral sclerosis patients, writing encryption, and human–machine information transfer [162]. Electrospun nanofiber strain sensors with a Ti3C2Tx-based conductive interlayer demonstrate remarkable performance in gesture language recognition [163]. Moreover, multifunctional, robust, flexible knitted strain sensors composed of conductive Ti3C2Tx nanosheets serve as intelligent sensing gloves for accessible communication among individuals with impaired hearing [164]. Conductive organic hydrogels based on Ti3C2Tx nanosheets exhibit the highest pressure-sensing performance (S-p1 = 782.7 kPa−1) and ultra-low-temperature tolerance (−81 °C). Zinc-activated Ti3C2Tx nanosheets endow hydrogel sensors with good antibacterial capabilities and have potential to be used for medical applications [165]. Radiation-synthesized poly(ionic liquid)/Ti3C2Tx gels display excellent antifreeze properties (−60 °C), durability (90 days at room temperature), and outstanding antibacterial activity against E. coli (99.82%) and S. aureus (99.98%). Multifunctional sensors based on these hybrid gels offer real-time and rapid response capabilities in detecting various human activities, recognizing writing strokes, and facilitating encrypted information transmission [166]. Capacitive pressure sensors prepared using carbon nanotubes, Ti3C2Tx, and polydimethylsiloxane composites feature an ultra-low detection limit, a wide detection range, high sensitivity, a fast response time, and excellent stability. These sensors accurately recognize mechanical stimuli while meeting the requirements for elasticity, antibacterial activity, wearing comfort, and long-term stability, thus providing innovative solutions for monitoring human activities and detecting diseases non-invasively [167]. Some sensors also possess superhydrophobic and anti-fouling properties, enabling them to resist interference from water droplets and flow [168], thereby making them suitable for use in humid or rainy conditions and complex environments [169].
Ti3C2Tx composite-based sensors also demonstrate impressive photothermal conversion capabilities. Under NIR irradiation (808 nm, 0.2 W/cm2) for 10 s, they reach a temperature of 55 °C, effectively achieving the complete sterilization of both S. aureus and E. coli [170]. These sensors hold potential for further development into a diagnostic and treatment system that monitors human movement while simultaneously providing thermotherapy and antibacterial treatment at the site of motion injury [171]. The introduction of noble metal Ag nanoparticles, leveraging the synergistic photothermal effects of Ti3C2Tx and Ag, enhances photothermal energy storage and antibacterial properties, making this method suitable for moderate PTT and pain relief [172]. The exceptional electrical conductivity and distinct plasma effects of metal nanoparticles have amplified sensor sensitivity by fourfold. Within the range of 0–100 kPa, the linear sensitivity measures 24.5 kPa−1. Additionally, there is a 25.3 °C electrothermal enhancement at 7.5 V and a 5.6 °C photothermal enhancement at 100 mW cm−2. Moreover, these enhancements result in an antibacterial efficiency of up to 99.9% [36]. Strain sensors assembled with Ti3C2Tx nanosheets and PDA/Ni2+ demonstrate strong antibacterial effects, rapid response to NIR radiation, and controllable surface temperatures, showing great potential for on-demand thermotherapy [152] (Figure 8). Epidermal sensors assembled from antibacterial Ti3C2Tx hydrogels can sensitively monitor human movements during rehabilitation training, detect tiny electrophysiological signals for the diagnosis of cardiovascular and muscle-related diseases, facilitate wearable human–machine interaction, and treat wound infections to promote wound healing [173]. Wireless flexible sensors made of Ti3C2Tx-based hydrogels, benefiting from the synergistic effect of antibacterial properties and electrical stimulation, provide potent angiogenic effects for diabetic wounds, promote collagen production, and accelerate wound healing, thus effectively safeguarding high-risk populations [174].

4.1.3. Medical Implant Field

Ti3C2Tx composite antibacterial materials hold broad application prospects in the field of medical implants due to their good biocompatibility. A 2019 study first explored the application of Ti3C2Tx films in bone tissue engineering and guided bone regeneration therapy. This study confirmed the high cell compatibility of Ti3C2Tx films and their capacity to foster osteogenic differentiation in vitro. After the implantation of Ti3C2Tx membranes into rat subcutaneous and cranial defect sites, they exhibited favorable biocompatibility, osteoinductivity, and bone regeneration activity in vivo [175]. Infections present a substantial risk to individuals with medical implants, potentially leading to implant failure, tissue necrosis, and even amputation. Ti3C2Tx can mitigate and manage microbial biofilms on indwelling implantable materials and devices by its exceptional antibacterial properties, effectively addressing both antibacterial and bone integration [176]. Crumpled Ti3C2Tx coatings on implants exert robust antibacterial effects against S. aureus and E. coli, effectively eliminating >= 99% of adherent bacterial cells and restoring surface functionality [122]. Various thicknesses of few-layer Ti3C2Tx coatings, fabricated through anodic electrophoretic deposition, notably enhance the initial adhesion and proliferation capabilities of osteogenic precursor cells such as MC3T3-E1. Moreover, they effectively inhibit the adhesion and cell viability of S. aureus and drug-resistant strains, with the antibacterial efficacy directly correlating with the Ti3C2Tx content [177]. Integrating Ti3C2Tx nanosheets with other materials to formulate composite coatings represents a promising strategy for antibiotic-free antibacterial treatment in orthopedic implants, effectively combating pathogenic infections and circumventing antibiotic resistance. For example, embedding plasma-coated cobalt nanowires with Ti3C2Tx can rapidly capture photogenerated electrons from Ti3C2Tx nanosheets, inhibit hot electron–hole recombination, and promote charge carrier transfer under 808 nm NIR irradiation, thereby increasing the yield of bactericidal ROS [178]. CoFe2O4/Ti3C2Tx nanosheet-modified hydrogels spin-coated onto hard tissue implant materials endow the implants with potent anti-pathogen capabilities through PTT local heat, PDT, and chemodynamic therapy ROS accumulation properties [179]. Ti3C2Tx composite hydrogel scaffolds loaded with rat bone marrow mesenchymal stem cells using 3D printing technology demonstrate a reliance on the biocompatibility and osteogenic capacity of Ti3C2Tx. This synergy of antibacterial and osteogenic activities in vivo promotes the healing and bone regeneration of rat mandibular defect infections with S. aureus [151]. Beyond bone implants, Ti3C2Tx demonstrates promising potential for neural repair in neural catheter implant applications, attributed to its outstanding conductivity. Composite fiber membranes, incorporating varying Ti3C2Tx contents via electrospinning, contribute to the repair and regeneration of peripheral nerve injuries by restoring electrical signal transmission and regulating neuronal membrane functions. The addition of Ti3C2Tx improves the hydrophilicity and conductivity of electrospun composite membranes, fostering favorable biocompatibility for cell growth while achieving an antibacterial rate of up to 80%, effectively preventing post-implantation wound inflammation [143].

4.2. Water Treatment

As a novel two-dimensional nanomaterial, Ti3C2Tx has superior performance in water treatment, particularly when prepared with Ti3C2Tx antibacterial composites. Currently, Ti3C2Tx antibacterial composite materials are primarily utilized in seawater desalination and sewage treatment.

4.2.1. Seawater Desalination

Water scarcity has become a global urgent concern, exacerbated by the relentless progression of society and economy, which intensifies the imbalance between water supply and demand. Against this backdrop, seawater desalination technology has emerged as a pivotal solution to mitigate water scarcity. Sustainable solar-driven seawater desalination technology stands out as a promising avenue to overcome freshwater resource shortages [180]. Nonetheless, solar water evaporation materials encounter serious challenges, such as scaling, low freshwater yield, and inadequate long-term performance in seawater [181]. Ti3C2Tx composites boast high photothermal conversion efficiency, substantially reducing energy consumption in the seawater desalination process, thereby heralding a technological revolution in solar thermal desalination [182]. The Ag/Ti3C2Tx/aromatic nanofiber composite aerogel photothermal evaporator, characterized by internal 3D micropores and layered structures, showcases exceptional and efficient salt resistance, desalination, antibacterial properties, and heavy metal ion wastewater purification capabilities. Its excellent thermal stability ensures resilience across a wide temperature range from −500 °C to 196 °C [183]. Inspired by natural trees, a biomimetic solar-driven interfacial evaporator has been successfully developed using vertically aligned hydrophilic sodium alginate aerogel and a layered assembly of Ti3C2Tx-interwoven carbon nanotube networks as a micro-nano photothermal absorption layer. This innovative design exhibits excellent water evaporation performance, reaching up to 2.416 kg m−2 h−1 (after subtracting dark conditions) with a solar energy utilization efficiency of up to 90.56%, which is far higher than that of most reported seawater desalination devices. Furthermore, the solar-driven evaporator boasts robust antibacterial properties and long-term operational stability [184]. An array of Janus photothermal conversion materials with asymmetric characteristics has gradually attracted the interest of researchers. Notably, the Janus photothermal porous fiber membrane, composed of a hydrophobic Ti3C2Tx/polydimethylsiloxane coating and a hydrophilic polylactic acid/TiO2 nanofluid porous fiber membrane, exhibits a stable photothermal conversion efficiency of 60% with a freshwater yield of 1 kg m−2 h−1 and an ion concentration of <=1 ppm under direct one-sun irradiation, outperforming other membranes reported in the literature. With the synergistic effect of TiO2 nanofluid and Ti3C2Tx, the Janus fiber membrane achieves a salt rejection rate of up to 99.95% and antibacterial activity of up to 100% [185]. Moreover, the waste biomass solar evaporator with a Janus structure made from porous pomelo peel and a Ti3C2Tx coating has an evaporation rate of 1.48 kg m−2 h−1 and a photothermal conversion efficiency of 92.3%. Even after 50 cycles under simulated one-sun lighting, it sustains high evaporation performance and exhibits excellent antibacterial properties [186].

4.2.2. Sewage Disposal

As the contamination of the environment by industrial pollutants and organic dyes continues to escalate, the desire for clean water resources becomes increasingly urgent. In recent years, photocatalysis has emerged as a predominant method for addressing water pollution, with Ti3C2Tx standing out due to its dual functionality in photocatalytic degradation and antibacterial action [187]. Ti3C2Tx exhibits significant advantages in wastewater treatment, effectively eliminating organic pollutants, heavy metal ions, and bacteria from contaminated water sources [105]. Micron-thick Ti3C2Tx membranes, filtered on a PVDF carrier, enhance the antibacterial activity of the PVDF membrane against Bacillus subtilis (B. subtilis) and E. coli by more than 73% and 67%, respectively. This underscores the potential of Ti3C2Tx-modified membranes not only in combating common waterborne bacteria but also as antifouling membranes in wastewater treatment processes [188]. Furthermore, the combination of Ti3C2Tx with cellulose to form Ti3C2Tx/cellulose photothermal membranes achieves nearly 94% light absorption efficiency and demonstrates antibacterial efficiency exceeding 99%, bringing promising prospects for practical application in long-term wastewater treatment [153] (Figure 8). Composite membranes formed by combining Ti3C2Tx nanosheets with silver nanowires [189] or nanoparticles [190] exhibit an inhibition rate exceeding 99% against E. coli and S. aureus, positioning them as attractive candidates for developing nanofiltration membranes for water purification and biomedical applications. Moreover, Ti3C2Tx-functionalized phages, prepared by loading Ti3C2Tx nanoparticle fragments onto well-characterized phages via electrostatic bonding, have high specificity for the recognition of host receptors, the targeting of phage and bacterial surface negative charges [191]. This approach can significantly increase the adsorption rate and stability of phages in aquatic environments, demonstrating robust antibacterial effects on bacterial cell targets and reducing artificial contamination in water samples by 99.99%. To scale up wastewater treatment efforts, researchers have designed a simple yet effective antibacterial nanorobot by loading ultra-small gold nanorods onto the surface of Ti3C2Tx nanosheets. This nanorobot enhances light absorption cross-sections, improving mobility and antibacterial capabilities under irradiation. Compared to traditional static antibacterial methods, this unique self-propelled capability offers an effective and sustainable approach for treating bacterial pollution in water on a large scale [192].

4.3. Textile Field

The pressing demand for multifunctional and high-performance wearable heaters in future human health applications is undeniable. Wearable heaters, fashioned by decorating fibers with Ti3C2Tx through a straightforward solution dipping coating technique, hold promising potential and stand as an optimal choice for forthcoming health management and safeguarding endeavors. The Ti3C2Tx-decorated fabrics boast commendable dual-driven energy conversion capabilities, seamlessly integrating photothermal and electrothermal functionalities, which achieve an impressive antibacterial efficiency surpassing 99% against Gram-negative E. coli within a mere 20 min of contact [46] (Figure 8). These fabrics also excel in electromagnetic interference shielding, demonstrating an efficiency of 42.1 dB in the x-band [193]. A triboelectric nanogenerator, utilizing fabric as the electrode, converts open-circuit voltage into an electrical signal by harnessing the energy from humans’ tiny movements, showcasing immense potential for self-powered electronic device applications [194]. Customized Janus textiles with asymmetric optical properties precisely regulate body temperature to adapt seamlessly to fluctuating climatic conditions, rendering them indispensable for optimal personal thermal management and climate resilience efforts [195]. Furthermore, self-disinfecting textiles, harnessing synergistic photodynamic and photothermal effects, herald a breakthrough in healthcare applications. Antibacterial medical textiles, achieved through the curling of exfoliated 2D Ti3C2Tx MXene sheets on polypropylene fibers, exhibit unparalleled efficacy in reducing bacterial viability by 100% through physical contact and light-induced ROS generation [196]. Intelligent photochromic self-disinfecting textiles, engineered through the in-situ growth of MOF, electro-sprayed Ti3C2Tx colloid, and screen-printed thermochromic dyes, demonstrate the capability to neutralize 99.9999% of Gram-negative (E. coli ATCC 8099) and Gram-positive (S. aureus ATCC 6538) bacteria within 30 min [87]. Mechanism insights reveal that light-driven singlet oxygen and thermal effects are pivotal in bacterial inactivation. Combining Ti3C2Tx quantum dots with potent antibacterials forms an organic–inorganic hybrid layer on fiber materials. This enables textiles to exhibit varying fluorescence recovery capabilities following exposure to different pathogen concentrations, achieving a remarkable bactericidal efficiency exceeding 99.99% against both E. coli and S. aureus within a brief time frame of 30 min [197]. Additionally, pioneering the exploration of multifunctional coatings via composite nanomaterial design holds the key to advancing the next generation of polyester textiles. Polyester textiles, assembled with MoO3-x quantum dots on Ti3C2Tx, exhibit exceptional bactericidal efficacy exceeding 99% against E. coli and S. aureus under sunlight, while concurrently mitigating smoke by 36.16% [198]. Nonwoven fabric created by combining a conductive material such as Ti3C2Tx with traditional textile PET achieves remarkable antibacterial efficiency surpassing 99.99% against E. coli under simulated sunlight exposure at a rate of 300 mW/cm2. The formation of a dense protective Ti3C2Tx layer during combustion significantly enhances fabric flame retardancy, showcasing boundless potential in crafting adaptive wearable technologies [199].

4.4. Food Filed

The globalization of markets and the surge in global population have spurred an escalating demand for food, presenting significant challenges in ensuring food safety. Amid this backdrop, averting bacterial contamination emerges as a prevalent strategy to prolong the shelf life of food items and uphold their quality. Integrating Ti3C2Tx particles into a polylactic acid polymer matrix to craft a composite film not only maintains the thermal and mechanical properties of polylactic acid but also provides the film with the inherent antibacterial prowess of Ti3C2Tx. Remarkably, a mere 0.5 wt.% of Ti3C2Tx is adequate to achieve a staggering 99.9999% bactericidal activity against Listeria monocytogenes (a six-log reduction), while for Salmonella, only 5 wt.% of Ti3C2Tx suffices to attain a 99.999% bactericidal efficacy (a five-log reduction) [200]. Furthermore, the amalgamation of Ti3C2Tx and tannic acid into chitosan networks via hydrogen bonding and electrostatic self-assembly engenders biopolymer-based composite films with substantially enhanced antibacterial and antioxidant capabilities. These films boast exceptional water vapor and oxygen barrier properties, effectively extending the shelf life of perishable produce like bananas and grapes. Consequently, these composite films emerge as ideal solutions for storing fruits and vegetables [154] (Figure 8). Moreover, Ti3C2Tx exhibits the remarkable ability to impede ice crystal formation and growth, thereby mitigating cryoinjury during cooling and thawing, which could otherwise jeopardize stem cells. This effect translates into a notable enhancement in cell viability, elevating it from 38.4% to 80.9%. Additionally, Ti3C2Tx MXene showcases synergistic antibacterial activity under laser irradiation, facilitating the sterile cryopreservation of stem cells [201].

4.5. Other Filed

In the realm of drug delivery, Ti3C2Tx composite antibacterial materials provide a versatile platform for developing drug delivery systems. They enable targeted drug delivery, diminish drug toxicity, and enhance the pharmacokinetics of drug molecules. By grafting doxorubicin (DOX) as a model drug onto functionalized Ti3C2Tx nanosheets, a nanoparticle with exceptional drug loading and encapsulation efficiency, as well as outstanding dual (ROS- and pH-responsive) properties, is achieved. Moreover, Ti3C2Tx-DOX nanosheets exhibit excellent antibacterial activity against both Gram-negative E. coli and Gram-positive B. subtilis [202]. Incorporating Ti3C2Tx into a modified polyurethane enhances its physicochemical properties and biocompatibility. This composite film controls drug release from the polymer matrix during gradual degradation and swelling, facilitating precise drug delivery to the site of the disease, thereby improving therapeutic efficacy and minimizing damage to healthy tissues [203]. Expanding into the industrial sector, the application of Ti3C2Tx materials has broadened, particularly in fabricating composites with high-performance antibacterial and antioxidant properties. An innovative strategy involves intercalating Ti3C2Tx into montmorillonite layers to form a nanocomposite with a micro-nano protective structure. This composite material exhibits remarkable antibacterial efficacy, with a suppression rate exceeding 99.9% against S. aureus. This high antibacterial performance persists for up to 28 days. Additionally, these composites demonstrate potent antioxidant activity, providing a solid theoretical foundation and practical application potential for the preparation and maintenance of valuable furniture [204]. Furthermore, the development of a copper–organic phosphate–Ti3C2Tx heterojunction has ushered in new opportunities for the industrial sector. This heterojunction not only achieves over 99.9% antibacterial efficiency, but also yields epoxy resin nanocomposites with satisfactory flame-retardant properties. Such a material, coupling high antibacterial and flame-retardant performance, creates more opportunities for applications in industries such as coatings, medical devices, and furniture. It enhances product safety, durability, and inspires new ideas and solutions for product innovation in these sectors [205].
Ti3C2Tx-based composite antibacterial materials possess broad application prospects, spanning from food preservation to seawater desalination, wastewater treatment, and medical fields such as drug delivery systems, wound dressings, medical sensors, and bone implants, and they even have significant applications in industrial sectors. In the future, Ti3C2Tx composite antibacterial materials are expected to achieve breakthroughs in more frontier fields with their excellent antibacterial properties and wide application potential.

5. Conclusions

In summary, Ti3C2Tx materials exhibit tremendous potential in the field of antibacterial applications due to their unique physical exfoliation, photodynamic, and photothermal properties. However, inherent limitations such as inadequate solar response capability and carrier separation efficiency have constrained their photocatalytic performance, leading to suboptimal antibacterial efficacy. To overcome these intrinsic constraints, researchers have significantly enhanced the antibacterial properties of Ti3C2Tx composite materials through compositional diversification and modifications in composite methodologies. Over the past decade, Ti3C2Tx composite materials have made rapid strides in antibacterial applications, not only achieving breakthroughs in material compositional diversity but also demonstrating extensive possibilities in practical applications. As technology matures and scientific research deepens, the research and application prospects of Ti3C2Tx composite antibacterial materials are expected to broaden further, promising significant contributions to human health, environmental protection, and industrial development.

6. Challenges and Prospects

Ti3C2Tx composite antibacterial materials, as a promising nanomaterial with broad application potential in the antibacterial field, have made significant breakthroughs. However, several challenges remain. With continuous technological advancements, further enhancing the performance of Ti3C2Tx composite antibacterial materials can promote their widespread application and development in fields such as medicine, environmental protection, and daily essentials as follows:
(1) Structure optimization for customized design: A deep understanding of the relationship between the microstructure and properties of materials enables precise control of the structural parameters of composite materials, such as nano-scale composite components, morphology, and crystal structure, to achieve customized designs of material properties. Utilizing computational simulation, biomimetic design, and modern material characterization techniques, such as transmission electron microscopy and atomic force microscopy, enables the multi-scale characterization, prediction, and optimization of the structures and properties of Ti3C2Tx composite antibacterial materials. This comprehensive understanding of the structure–property relationship facilitates precise control tailored to the requirements of various application scenarios.
(2) The in-depth exploration of antibacterial mechanisms: With the continuous development of biotechnology and materials science, future research will focus on understanding the release mechanisms and control strategies of antibacterial active substances in Ti3C2Tx composite materials at the molecular level, utilizing methods such as computational simulation and molecular dynamics simulation. Effective methods will be sought to precisely control the release rate, amount, and mode of antibacterial active substances. The interaction processes between antibacterial active substances and bacteria, especially drug-resistant bacteria, will be thoroughly explored to reveal the possible adaptability and evolution processes of bacteria to the materials and to study the action mechanisms of the materials on drug-resistant bacteria. Additionally, considering that Ti3C2Tx composite antibacterial materials may possess multiple antibacterial mechanisms, such as physical capture, chemical reaction, photocatalysis, etc., future research will focus on exploring the interactions and synergistic effects among these different antibacterial mechanisms. This will further enhance the antibacterial effectiveness of the materials and expand their applicability in different scenarios.
(3) The synergistic optimization of biocompatibility and antibacterial performance: Currently, Ti3C2Tx composite antibacterial materials are increasingly used in the field of human implants. However, there is limited research on their biocompatibility, especially regarding their long-term usage. In the future, it is necessary to focus on establishing and improving biocompatibility evaluation standards and to explore how surface modification and functionalization can be utilized to control parameters such as nanoscale structure and surface charge, thereby improving the interface properties between the material and the biological system to reduce immune reactions and cell toxicity. Additionally, future research could explore synergistic optimization strategies between biocompatibility and antibacterial performance. This entails modulating the material’s structure, composition, and surface properties to achieve the dual optimization of biocompatibility and antibacterial performance.
(4) The optimization of preparation processes for green and environmentally friendly outcomes: Utilizing green solvents, such as water and ethanol, as replacements for traditional organic solvents reduces environmental pollution and chemical usage. Optimizing process parameters, such as temperature, time, and pressure, can decrease energy consumption and exhaust emissions. Additionally, exploring novel preparation methods like 3D printing and electron beam irradiation techniques enables the precise control and functional modification of the structures of Ti3C2Tx composite materials, enhancing their practicality and specificity. These approaches provide more expedient and efficient solutions and market prospects for the application of materials.
In conclusion, Ti3C2Tx composite antibacterial materials face a series of challenges but also have broad application prospects. Through the continuous optimization of preparation processes, structural and performance regulation, and in-depth research into antibacterial mechanisms, Ti3C2Tx composite antibacterial materials will play a role in more fields, contributing to the development of human society.

Author Contributions

Conceptualization, B.H. and Y.L.; writing—original draft preparation, H.C., Y.W. (Yilun Wang), X.C., Y.W. (Yue Wu), Q.D., W.Z., T.W., and Q.Y.; writing—review and editing, H.C. and B.H.; visualization, Z.W.; supervision, B.H.; project administration, B.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the “Provincial Natural Science Foundation Project”, grant number 2024AFC063.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Huo, J.; Jia, Q.; Huang, H.; Zhang, J.; Li, P.; Dong, X.; Huang, W. Emerging Photothermal-Derived Multimodal Synergistic Therapy in Combating Bacterial Infections. Chem. Soc. Rev. 2021, 50, 8762–8789. [Google Scholar] [CrossRef] [PubMed]
  2. Tavakolian, M.; Jafari, S.M.; Van De Ven, T.G.M. A Review on Surface-Functionalized Cellulosic Nanostructures as Biocompatible Antibacterial Materials. Nano-Micro Lett. 2020, 12, 73. [Google Scholar] [CrossRef]
  3. Von Nussbaum, F.; Brands, M.; Hinzen, B.; Weigand, S.; Haebich, D. Antibacterial Natural Products in Medicinal Chemistry—Exodus or Revival? Angew. Chem. Int. Ed. 2006, 45, 5072–5129. [Google Scholar] [CrossRef]
  4. Rutledge-Taylor, K.; Matlow, A.; Gravel, D.; Embree, J.; Le Saux, N.; Johnston, L.; Suh, K.; Embil, J.; Henderson, E.; John, M.; et al. A Point Prevalence Survey of Health Care-Associated Infections in Canadian Pediatric Inpatients. Am. J. Infect. Control 2012, 40, 491–496. [Google Scholar] [CrossRef] [PubMed]
  5. Bafail, A.S.; Alamri, A.M.; Spivakovsky, S. Effect of Antibiotics on Implant Failure and Postoperative Infection: Question: Among Patients Receiving Dental Implants, Does the Use of Antibiotics, When Compared with a Control Group, Reduce the Frequency of Implant Failure and Postoperative Infection? Evid. Based Dent. 2014, 15, 58. [Google Scholar] [CrossRef] [PubMed]
  6. De Kraker, M.E.A.; Stewardson, A.J.; Harbarth, S. Will 10 Million People Die a Year Due to Antimicrobial Resistance by 2050? PLoS Med. 2016, 13, e1002184. [Google Scholar] [CrossRef]
  7. Li, Y.; Wang, D.; Wen, J.; Yu, P.; Liu, J.; Li, J.; Chu, H. Chemically Grafted Nanozyme Composite Cryogels to Enhance Antibacterial and Biocompatible Performance for Bioliquid Regulation and Adaptive Bacteria Trapping. ACS Nano 2021, 15, 19672–19683. [Google Scholar] [CrossRef]
  8. Gao, H.; Cui, C.; Wang, L.; Jacobs-Lorena, M.; Wang, S. Mosquito Microbiota and Implications for Disease Control. Trends Parasitol. 2020, 36, 98–111. [Google Scholar] [CrossRef] [PubMed]
  9. Goddard, F.G.B.; Chang, H.H.; Clasen, T.F.; Sarnat, J.A. Exposure Measurement Error and the Characterization of Child Exposure to Fecal Contamination in Drinking Water. Npj Clean Water 2020, 3, 19. [Google Scholar] [CrossRef]
  10. Xu, H.; Chen, Z.; Wu, X.; Zhao, L.; Wang, N.; Mao, D.; Ren, H.; Luo, Y. Antibiotic Contamination Amplifies the Impact of Foreign Antibiotic-Resistant Bacteria on Soil Bacterial Community. Sci. Total Environ. 2021, 758, 143693. [Google Scholar] [CrossRef]
  11. Russell, A.D. Mechanisms of Antimicrobial Action of Antiseptics and Disinfectants: An Increasingly Important Area of Investigation. J. Antimicrob. Chemother. 2002, 49, 597–599. [Google Scholar] [CrossRef] [PubMed]
  12. Al-Jumaili, A.; Alancherry, S.; Bazaka, K.; Jacob, M. Review on the Antimicrobial Properties of Carbon Nanostructures. Materials 2017, 10, 1066. [Google Scholar] [CrossRef]
  13. Zheng, J.; Li, J.; Zhang, L.; Chen, X.; Yu, Y.; Huang, H. Post-Graphene 2D Materials-Based Antimicrobial Agents: Focus on Fabrication Strategies and Biosafety Assessments. J. Mater. Sci. 2020, 55, 7226–7246. [Google Scholar] [CrossRef]
  14. Miao, H.; Teng, Z.; Wang, C.; Chong, H.; Wang, G. Recent Progress in Two-Dimensional Antimicrobial Nanomaterials. Chem. Eur. J. 2019, 25, 929–944. [Google Scholar] [CrossRef] [PubMed]
  15. Liang, Y.; Zhang, H.; Yuan, H.; Lu, W.; Li, Z.; Wang, L.; Gao, L.-H. Conjugated Polymer and Triphenylamine Derivative Codoped Nanoparticles for Photothermal and Photodynamic Antimicrobial Therapy. ACS Appl. Bio Mater. 2020, 3, 3494–3499. [Google Scholar] [CrossRef] [PubMed]
  16. Yang, Y.; Deng, Y.; Huang, J.; Fan, X.; Cheng, C.; Nie, C.; Ma, L.; Zhao, W.; Zhao, C. Size-Transformable Metal–Organic Framework–Derived Nanocarbons for Localized Chemo-Photothermal Bacterial Ablation and Wound Disinfection. Adv. Funct. Mater. 2019, 29, 1900143. [Google Scholar] [CrossRef]
  17. Zhou, X.; Wang, Z.; Chan, Y.K.; Yang, Y.; Jiao, Z.; Li, L.; Li, J.; Liang, K.; Deng, Y. Infection Micromilieu-Activated Nanocatalytic Membrane for Orchestrating Rapid Sterilization and Stalled Chronic Wound Regeneration. Adv. Funct. Mater. 2022, 32, 2109469. [Google Scholar] [CrossRef]
  18. Mathis, T.S.; Maleski, K.; Goad, A.; Sarycheva, A.; Anayee, M.; Foucher, A.C.; Hantanasirisakul, K.; Shuck, C.E.; Stach, E.A.; Gogotsi, Y. Modified MAX Phase Synthesis for Environmentally Stable and Highly Conductive Ti3C2 MXene. ACS Nano 2021, 15, 6420–6429. [Google Scholar] [CrossRef]
  19. Cui, L.; Wen, J.; Deng, Q.; Du, X.; Tang, T.; Li, M.; Xiao, J.; Jiang, L.; Hu, G.; Cao, X.; et al. Improving the Photocatalytic Activity of Ti3C2 MXene by Surface Modification of N Doped. Materials 2023, 16, 2836. [Google Scholar] [CrossRef]
  20. Kumar, J.A.; Prakash, P.; Krithiga, T.; Amarnath, D.J.; Premkumar, J.; Rajamohan, N.; Vasseghian, Y.; Saravanan, P.; Rajasimman, M. Methods of Synthesis, Characteristics, and Environmental Applications of MXene: A Comprehensive Review. Chemosphere 2022, 286, 131607. [Google Scholar] [CrossRef]
  21. 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] [PubMed]
  22. 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]
  23. Solangi, N.H.; Mazari, S.A.; Mubarak, N.M.; Karri, R.R.; Rajamohan, N.; Vo, D.-V.N. Recent Trends in MXene-Based Material for Biomedical Applications. Environ. Res. 2023, 222, 115337. [Google Scholar] [CrossRef] [PubMed]
  24. Murali, G.; Reddy Modigunta, J.K.; Park, Y.H.; Lee, J.-H.; Rawal, J.; Lee, S.-Y.; In, I.; Park, S.-J. A Review on MXene Synthesis, Stability, and Photocatalytic Applications. ACS Nano 2022, 16, 13370–13429. [Google Scholar] [CrossRef] [PubMed]
  25. Chen, K.; Chen, Y.; Deng, Q.; Jeong, S.-H.; Jang, T.-S.; Du, S.; Kim, H.-E.; Huang, Q.; Han, C.-M. Strong and Biocompatible Poly(Lactic Acid) Membrane Enhanced by Ti3C2Tz (MXene) Nanosheets for Guided Bone Regeneration. Mater. Lett. 2018, 229, 114–117. [Google Scholar] [CrossRef]
  26. Li, Z.; Zhang, H.; Han, J.; Chen, Y.; Lin, H.; Yang, T. Surface Nanopore Engineering of 2D MXenes for Targeted and Synergistic Multitherapies of Hepatocellular Carcinoma. Adv. Mater. 2019, 31, 1902282. [Google Scholar] [CrossRef] [PubMed]
  27. Rasool, K.; Helal, M.; Ali, A.; Ren, C.E.; Gogotsi, Y.; Mahmoud, K.A. Antibacterial Activity of Ti3C2Tx MXene. ACS Nano 2016, 10, 3674–3684. [Google Scholar] [CrossRef]
  28. Han, X.; Huang, J.; Lin, H.; Wang, Z.; Li, P.; Chen, Y. 2D Ultrathin MXene-Based Drug-Delivery Nanoplatform for Synergistic Photothermal Ablation and Chemotherapy of Cancer. Adv. Healthc. Mater. 2018, 7, e1701394. [Google Scholar] [CrossRef] [PubMed]
  29. Yu, F.; Zhang, X.; Yang, Z.; Yang, P.; Ma, J. Environmental Applications of Two-Dimensional Transition Metal Carbides and Nitrides for Water Purification: A Review. Environ. Chem. Lett. 2022, 20, 633–660. [Google Scholar] [CrossRef]
  30. Mahar, I.; Memon, F.H.; Lee, J.-W.; Kim, K.H.; Ahmed, R.; Soomro, F.; Rehman, F.; Memon, A.A.; Thebo, K.H.; Choi, K.H. Two-Dimensional Transition Metal Carbides and Nitrides (MXenes) for Water Purification and Antibacterial Applications. Membranes 2021, 11, 869. [Google Scholar] [CrossRef]
  31. Seidi, F.; Arabi Shamsabadi, A.; Dadashi Firouzjaei, M.; Elliott, M.; Saeb, M.R.; Huang, Y.; Li, C.; Xiao, H.; Anasori, B. MXenes Antibacterial Properties and Applications: A Review and Perspective. Small 2023, 19, 2206716. [Google Scholar] [CrossRef] [PubMed]
  32. Zhang, Z.; Qi, Z.; Kong, W.; Zhang, R.; Yao, C. Applications of MXene and Its Modified Materials in Skin Wound Repair. Front. Bioeng. Biotechnol. 2023, 11, 1154301. [Google Scholar] [CrossRef] [PubMed]
  33. Parajuli, D.; Murali, N.; K. C., D.; Karki, B.; Samatha, K.; Kim, A.A.; Park, M.; Pant, B. Advancements in MXene-Polymer Nanocomposites in Energy Storage and Biomedical Applications. Polymers 2022, 14, 3433. [Google Scholar] [CrossRef] [PubMed]
  34. Hao, S.; Han, H.; Yang, Z.; Chen, M.; Jiang, Y.; Lu, G.; Dong, L.; Wen, H.; Li, H.; Liu, J.; et al. Recent Advancements on Photothermal Conversion and Antibacterial Applications over MXenes-Based Materials. Nano-Micro Lett. 2022, 14, 178. [Google Scholar] [CrossRef] [PubMed]
  35. Chen, X.; Wang, J.; Endo, Y.; Liu, G.; Wang, N.; Shao, Z. Natural Photothermal-Responsive MXene-Based Antimicrobial Nanomaterials Ti3C2/Au NPs for the Treatment of Deep Staphylococcus Aureus-Infected Wound. Mater. Des. 2023, 232, 112138. [Google Scholar] [CrossRef]
  36. Yu, Q.; Pan, J.; Li, J.; Su, C.; Huang, Y.; Bi, S.; Jiang, J.; Chen, N. A General Strategy to Immobilize Metal Nanoparticles on MXene Composite Fabrics for Enhanced Sensing Performance and Endowed Multifunctionality. J. Mater. Chem. C 2022, 10, 13143–13156. [Google Scholar] [CrossRef]
  37. Tahir, T.; Chaudhary, K.; Warsi, M.F.; Saif, M.S.; Alsafari, I.A.; Shakir, I.; Agboola, P.O.; Haider, S.; Zulfiqar, S. Synthesis of Sponge like Gd3+ Doped Vanadium Oxide/2D MXene Composites for Improved Degradation of Industrial Effluents and Pathogens. Ceram. Int. 2022, 48, 1969–1980. [Google Scholar] [CrossRef]
  38. Li, Q.; Wang, W.; Feng, H.; Cao, L.; Wang, H.; Wang, D.; Chen, S. NIR-Triggered Photocatalytic and Photothermal Performance for Sterilization Based on Copper Sulfide Nanoparticles Anchored on Ti3C2T MXene. J. Colloid Interface Sci. 2021, 604, 810–822. [Google Scholar] [CrossRef] [PubMed]
  39. Nie, X.; Wu, S.; Huang, F.; Li, W.; Qiao, H.; Wang, Q.; Wei, Q. “Dew-of-Leaf” Structure Multiple Synergetic Antimicrobial Modality Hybrid: A Rapid and Long Lasting Bactericidal Material. Chem. Eng. J. 2021, 416, 129072. [Google Scholar] [CrossRef]
  40. Zhang, W.-J.; Li, S.; Vijayan, V.; Lee, J.; Park, S.; Cui, X.; Chung, I.; Lee, J.; Ahn, S.; Kim, J.; et al. ROS- and pH-Responsive Polydopamine Functionalized Ti3C2Tx MXene-Based Nanoparticles as Drug Delivery Nanocarriers with High Antibacterial Activity. Nanomaterials 2022, 12, 4392. [Google Scholar] [CrossRef]
  41. Guo, L.; Hu, K.; Wang, H. Antimicrobial and Mechanical Properties of Ag@Ti3C2Tx-Modified PVA Composite Hydrogels Enhanced with Quaternary Ammonium Chitosan. Polymers 2023, 15, 2352. [Google Scholar] [CrossRef] [PubMed]
  42. Salmi, M.S.; Ahmed, U.; Aslfattahi, N.; Rahman, S.; Hardy, J.G.; Anwar, A. Potent Antibacterial Activity of MXene–Functionalized Graphene Nanocomposites. RSC Adv. 2022, 12, 33142–33155. [Google Scholar] [CrossRef] [PubMed]
  43. Li, Y.; Lyu, Y.; Yang, Z.; Wang, T.; Zhang, Q.; Zhang, B. Preparation and Specific Adsorption of Dual Antibacterial BSA Surface Imprinted GO-PEI/MXene Wrinkled Microspheres. AIChE J. 2023, 69, e18157. [Google Scholar] [CrossRef]
  44. Fu, Y.; Cheng, Y.; Wei, Q.; Zhao, Y.; Zhang, W.; Yang, Y.; Li, D. Multifunctional Biomass Composite Aerogel Co-Modified by MXene and Ag Nanowires for Health Monitoring and Synergistic Antibacterial Applications. Appl. Surf. Sci. 2022, 598, 153783. [Google Scholar] [CrossRef]
  45. Zheng, Y.; Yan, Y.; Lin, L.; He, Q.; Hu, H.; Luo, R.; Xian, D.; Wu, J.; Shi, Y.; Zeng, F.; et al. Titanium Carbide MXene-Based Hybrid Hydrogel for Chemo-Photothermal Combinational Treatment of Localized Bacterial Infection. Acta Biomater. 2022, 142, 113–123. [Google Scholar] [CrossRef] [PubMed]
  46. Yan, B.; Zhou, M.; Liao, X.; Wang, P.; Yu, Y.; Yuan, J.; Wang, Q. Developing a Multifunctional Silk Fabric with Dual-Driven Heating and Rapid Photothermal Antibacterial Abilities Using High-Yield MXene Dispersions. ACS Appl. Mater. Interfaces 2021, 13, 43414–43425. [Google Scholar] [CrossRef] [PubMed]
  47. Liu, Y.; Chen, X.; Sun, J.; Xu, N.; Tang, Q.; Ren, J.; Chen, C.; Lei, W.; Zhang, C.; Liu, D. Large-Scale Production of MXenes as Nanoknives for Antibacterial Application. Nanoscale Adv. 2023, 5, 6572–6581. [Google Scholar] [CrossRef] [PubMed]
  48. Wang, J.; Xuan, J.; Liu, Y.; Li, Z.; Han, Y.; Wang, Z. NIR-Dependent Photothermal-Photodynamic Synergistic Antibacterial Mechanism for Titanium Carbide Nanosheets Intercalated and Delaminated by Tetramethylammonium Hydroxide. Biomater. Adv. 2023, 152, 213492. [Google Scholar] [CrossRef]
  49. Qu, X.; Guo, Y.; Xie, C.; Li, S.; Liu, Z.; Lei, B. Photoactivated MXene Nanosheets for Integrated Bone–Soft Tissue Therapy: Effect and Potential Mechanism. ACS Nano 2023, 17, 7229–7240. [Google Scholar] [CrossRef]
  50. Talreja, N.; Ashfaq, M.; Chauhan, D.; Mangalaraja, R.V. Cu-MXene: A Potential Biocide for the next-Generation Biomedical Application. Mater. Chem. Phys. 2023, 294, 127029. [Google Scholar] [CrossRef]
  51. Yu, Z.; Jiang, L.; Liu, R.; Zhao, W.; Yang, Z.; Zhang, J.; Jin, S. Versatile Self-Assembled MXene-Au Nanocomposites for SERS Detection of Bacteria, Antibacterial and Photothermal Sterilization. Chem. Eng. J. 2021, 426, 131914. [Google Scholar] [CrossRef]
  52. Li, L.; Ji, X.; Chen, K. Conductive, Self-Healing, and Antibacterial Ag/MXene-PVA Hydrogel as Wearable Skin-like Sensors. J. Biomater. Appl. 2023, 37, 1169–1181. [Google Scholar] [CrossRef] [PubMed]
  53. Li, Z.; Xu, D.; Deng, Z.; Yin, J.; Qian, Y.; Hou, J.-T.; Ding, X.; Shen, J.; He, X. Single-Atom-Catalyzed MXene-Based Nanoplatform with Photo-Enhanced Peroxidase-Like Activity Nanotherapeutics for Staphylococcus Aureus Infection. Chem. Eng. J. 2023, 452, 139587. [Google Scholar] [CrossRef]
  54. Qu, W.; Zhao, H.; Zhang, Q.; Xia, D.; Tang, Z.; Chen, Q.; He, C.; Shu, D. Multifunctional Au/Ti3C2 Photothermal Membrane with Antibacterial Ability for Stable and Efficient Solar Water Purification under the Full Spectrum. ACS Sustain. Chem. Eng. 2021, 9, 11372–11387. [Google Scholar] [CrossRef]
  55. Wen, S.; Xiong, Y.; Cai, S.; Li, H.; Zhang, X.; Sun, Q.; Yang, R. Plasmon-Enhanced Photothermal Properties of Au@Ti3C2Tx Nanosheets for Antibacterial Applications. Nanoscale 2022, 14, 16572–16580. [Google Scholar] [CrossRef] [PubMed]
  56. Jiang, L.; Yu, Z.; Zhao, W.; Yang, Z.; Peng, Y.; Zhou, Y.; Lin, X.; Jin, S. Self-Assembled MXene-Au Multifunctional Nanomaterials with Various Shapes for Label-Free SERS Detection of Pathogenic Bacteria and Photothermal Sterilization. Anal. Chem. 2023, 95, 1721–1730. [Google Scholar] [CrossRef] [PubMed]
  57. Tang, W.; Dong, Z.; Zhang, R.; Yi, X.; Yang, K.; Jin, M.; Yuan, C.; Xiao, Z.; Liu, Z.; Cheng, L. Multifunctional Two-Dimensional Core–Shell MXene@Gold Nanocomposites for Enhanced Photo–Radio Combined Therapy in the Second Biological Window. ACS Nano 2019, 13, 284–294. [Google Scholar] [CrossRef] [PubMed]
  58. Zhu, X.; Zhu, Y.; Jia, K.; Abraha, B.S.; Li, Y.; Peng, W.; Zhang, F.; Fan, X.; Zhang, L. A Near-Infrared Light-Mediated Antimicrobial Based on Ag/Ti3C2Tx for Effective Synergetic Antibacterial Applications. Nanoscale 2020, 12, 19129–19141. [Google Scholar] [CrossRef] [PubMed]
  59. Liu, M.; Zheng, L.; Zha, K.; Yang, Y.; Hu, Y.; Chen, K.; Wang, F.; Zhang, K.; Liu, W.; Mi, B.; et al. Cu(II)@MXene Based Photothermal Hydrogel with Antioxidative and Antibacterial Properties for the Infected Wounds. Front. Bioeng. Biotechnol. 2023, 11, 1308184. [Google Scholar] [CrossRef]
  60. Wang, W.; Liu, C.; Zhang, M.; Zhang, C.; Cao, L.; Zhang, C.; Liu, T.; Kong, D.; Li, W.; Chen, S. In Situ Synthesis of 2D/2D MXene-COF Heterostructure Anchored with Ag Nanoparticles for Enhancing Schottky Photocatalytic Antibacterial Efficiency under Visible Light. J. Colloid Interface Sci. 2022, 608, 735–748. [Google Scholar] [CrossRef]
  61. Jin, J.; Wu, S.; Wang, J.; Xu, Y.; Xuan, S.; Fang, Q. AgPd Nanocages Sandwiched between a MXene Nanosheet and PDA Layer for Photothermally Improved Catalytic Activity and Antibacterial Properties. Dalton Trans. 2023, 52, 2335–2344. [Google Scholar] [CrossRef] [PubMed]
  62. Liu, G.; Wang, H.; Xu, C.; Fang, Q.; Wang, H.; Xu, Y.; Sang, M.; Xuan, S.; Hao, L. A MXene@AgAu@PDA Nanoplatform Loaded with AgAu Nanocages for Enhancing Catalytic Activity and Antibacterial Performance. J. Mater. Chem. B 2023, 11, 10678–10691. [Google Scholar] [CrossRef] [PubMed]
  63. Lemire, J.A.; Harrison, J.J.; Turner, R.J. Antimicrobial Activity of Metals: Mechanisms, Molecular Targets and Applications. Nat. Rev. Microbiol. 2013, 11, 371–384. [Google Scholar] [CrossRef] [PubMed]
  64. Kong, F.; He, X.; Liu, Q.; Qi, X.; Sun, D.; Zheng, Y.; Wang, R.; Bai, Y. Further Surface Modification by Carbon Coating for In-Situ Growth of Fe3O4 Nanoparticles on MXene Ti3C2 Multilayers for Advanced Li-Ion Storage. Electrochim. Acta 2018, 289, 228–237. [Google Scholar] [CrossRef]
  65. Alsafari, I.A. Synthesis of CuO/MXene Nanocomposite to Study Its Photocatalytic and Antibacterial Properties. Ceram. Int. 2022, 48, 10960–10968. [Google Scholar] [CrossRef]
  66. Wang, W.; Feng, H.; Liu, J.; Zhang, M.; Liu, S.; Feng, C.; Chen, S. A Photo Catalyst of Cuprous Oxide Anchored MXene Nanosheet for Dramatic Enhancement of Synergistic Antibacterial Ability. Chem. Eng. J. 2020, 386, 124116. [Google Scholar] [CrossRef]
  67. Warsi, A.-Z.; Aziz, F.; Zulfiqar, S.; Haider, S.; Shakir, I.; Agboola, P.O. Synthesis, Characterization, Photocatalysis, and Antibacterial Study of WO3, MXene and WO3/MXene Nanocomposite. Nanomaterials 2022, 12, 713. [Google Scholar] [CrossRef] [PubMed]
  68. Jiang, A.; Chen, X.; Xu, Y.; Shah, K.J.; You, Z. One-Step Hydrothermal Generation of Oxygen-Deficient N-Doped Blue TiO2–Ti3C2 for Degradation of Pollutants and Antibacterial Properties. Environ. Res. 2023, 235, 116657. [Google Scholar] [CrossRef]
  69. Sukidpaneenid, S.; Chawengkijwanich, C.; Pokhum, C.; Isobe, T.; Opaprakasit, P.; Sreearunothai, P. Multi-Function Adsorbent-Photocatalyst MXene-TiO2 Composites for Removal of Enrofloxacin Antibiotic from Water. J. Environ. Sci. 2023, 124, 414–428. [Google Scholar] [CrossRef] [PubMed]
  70. Xu, S.; Zhao, C.; Li, G.; Shi, Z.; Liu, B. In Situ Oxidized TiO2/MXene Ultrafiltration Membrane with Photocatalytic Self-Cleaning and Antibacterial Properties. RSC Adv. 2023, 13, 15843–15855. [Google Scholar] [CrossRef]
  71. Zhao, Y.; Kang, H.; Xia, Y.; Sun, L.; Li, F.; Dai, H. 3D Printed Photothermal Scaffold Sandwiching Bacteria Inside and Outside Improves the Infected Microenvironment and Repairs Bone Defects. Adv. Healthc. Mater. 2024, 13, 2302879. [Google Scholar] [CrossRef] [PubMed]
  72. Yu, H.; Xu, X.; Xie, Z.; Huang, X.; Lin, L.; Jiao, Y.; Li, H. High-Efficiency Near-Infrared Light Responsive Antibacterial System for Synergistic Ablation of Bacteria and Biofilm. ACS Appl. Mater. Interfaces 2022, 14, 36947–36956. [Google Scholar] [CrossRef] [PubMed]
  73. Qian, G.; Zhang, L.; Shuai, Y.; Wu, X.; Zeng, Z.; Peng, S.; Shuai, C. 3D-Printed CuFe2O4-MXene/PLLA Antibacterial Tracheal Scaffold against Implantation-Associated Infection. Appl. Surf. Sci. 2023, 614, 156108. [Google Scholar] [CrossRef]
  74. Liu, G.; Xiong, Q.; Xu, Y.; Fang, Q.; Leung, K.C.-F.; Sang, M.; Xuan, S.; Hao, L. Magnetically Separable MXene@Fe3O4/Au/PDA Nanosheets with Photothermal-Magnetolytic Coupling Antibacterial Performance. Appl. Surf. Sci. 2022, 590, 153125. [Google Scholar] [CrossRef]
  75. Gao, F.; Fang, M.; Zhang, S.; Ni, M.; Cai, Y.; Zhang, Y.; Tan, X.; Kong, M.; Xu, W.; Wang, X. Symmetry-Breaking Induced Piezocatalysis of Bi2S3 Nanorods and Boosted by Alternating Magnetic Field. Appl. Catal. B Environ. 2022, 316, 121664. [Google Scholar] [CrossRef]
  76. Li, J.; Li, Z.; Liu, X.; Li, C.; Zheng, Y.; Yeung, K.W.K.; Cui, Z.; Liang, Y.; Zhu, S.; Hu, W.; et al. Interfacial Engineering of Bi2S3/Ti3C2Tx MXene Based on Work Function for Rapid Photo-Excited Bacteria-Killing. Nat. Commun. 2021, 12, 1224. [Google Scholar] [CrossRef] [PubMed]
  77. Feng, H.; Wang, W.; Wang, T.; Pu, Y.; Ma, C.; Chen, S. Interfacial Regulation of BiOI@Bi2S3/MXene Heterostructures for Enhanced Photothermal and Photodynamic Therapy in Antibacterial Applications. Acta Biomater. 2023, 171, 506–518. [Google Scholar] [CrossRef] [PubMed]
  78. Li, X.; Wang, G.; Li, Q.; Wang, Y.; Lu, X. Dual Optimized Ti3C2Tx MXene@ZnIn2S4 Heterostructure Based on Interface and Vacancy Engineering for Improving Electromagnetic Absorption. Chem. Eng. J. 2023, 453, 139488. [Google Scholar] [CrossRef]
  79. Shen, J.; Liu, J.; Fan, X.; Liu, H.; Bao, Y.; Hui, A.; Munir, H.A. Unveiling the Antibacterial Strategies and Mechanisms of MoS2: A Comprehensive Analysis and Future Directions. Biomater. Sci. 2024, 12, 596–620. [Google Scholar] [CrossRef]
  80. Yang, Z.; Fu, X.; Ma, D.; Wang, Y.; Peng, L.; Shi, J.; Sun, J.; Gan, X.; Deng, Y.; Yang, W. Growth Factor-Decorated Ti3C2 MXene/MoS2 2D Bio-Heterojunctions with Quad-Channel Photonic Disinfection for Effective Regeneration of Bacteria-Invaded Cutaneous Tissue. Small 2021, 17, 2103993. [Google Scholar] [CrossRef]
  81. Alimohammadi, F.; Sharifian Gh, M.; Attanayake, N.H.; Thenuwara, A.C.; Gogotsi, Y.; Anasori, B.; Strongin, D.R. Antimicrobial Properties of 2D MnO2 and MoS2 Nanomaterials Vertically Aligned on Graphene Materials and Ti3C2 MXene. Langmuir 2018, 34, 7192–7200. [Google Scholar] [CrossRef]
  82. Liu, Y.; Ji, M.; Wang, P. Recent Advances in Small Copper Sulfide Nanoparticles for Molecular Imaging and Tumor Therapy. Mol. Pharm. 2019, 16, 3322–3332. [Google Scholar] [CrossRef]
  83. Wang, W.; Cao, L.; Li, Q.; Du, C.; Chen, S. Copper Sulfide Anchored MXene Improving Photo-Responsive Self-Healing Polyurethane with Enhanced Mechanical and Antibacterial Properties. J. Colloid Interface Sci. 2023, 630, 511–522. [Google Scholar] [CrossRef]
  84. Sun, G.; Jiang, X.; Liu, C.; Song, S.; Zhang, J.; Shen, J. FeS@LAB-35@Ti3C2 as a High-Efficiency Nanozyme for near Infrared Light Induced Photothermal Enhanced Chemodynamic Antibacterial Activity and Wound Healing. Nano Res. 2023, 16, 2840–2850. [Google Scholar] [CrossRef]
  85. Xu, D.; Zhang, Z.; Tao, K.; Han, L. A Heterostructure of a 2D Bimetallic Metal–Organic Framework Assembled on an MXene for High-Performance Supercapacitors. Dalton Trans. 2023, 52, 2455–2462. [Google Scholar] [CrossRef]
  86. Zhang, Y.; Yang, H.; Cao, P.; He, Y.; Zhao, Y.; Song, P.; Wang, R. Construction of UiO-NH2@TiC Schottky Junction and Their Effectively Photocatalytic and Antibacterial Performance. J. Clust. Sci. 2023, 34, 373–383. [Google Scholar] [CrossRef]
  87. Nie, X.; Wu, S.; Huang, F.; Wang, Q.; Wei, Q. Smart Textiles with Self-Disinfection and Photothermochromic Effects. ACS Appl. Mater. Interfaces 2021, 13, 2245–2255. [Google Scholar] [CrossRef] [PubMed]
  88. Li, C.-C.; Gao, M.-Y.; Sun, X.-J.; Tang, H.-L.; Dong, H.; Zhang, F.-M. Rational Combination of Covalent-Organic Framework and Nano TiO2 by Covalent Bonds to Realize Dramatically Enhanced Photocatalytic Activity. Appl. Catal. B Environ. 2020, 266, 118586. [Google Scholar] [CrossRef]
  89. Liu, C.; Wang, W.; Zhang, M.; Zhang, C.; Ma, C.; Cao, L.; Kong, D.; Feng, H.; Li, W.; Chen, S. Synthesis of MXene/COF/Cu2O Heterojunction for Photocatalytic Bactericidal Activity and Mechanism Evaluation. Chem. Eng. J. 2022, 430, 132663. [Google Scholar] [CrossRef]
  90. Sass, P. Antibiotics: Precious Goods in Changing Times. In Antibiotics; Sass, P., Ed.; Methods in Molecular Biology; Springer New York: New York, NY, USA, 2017; Volume 1520, pp. 3–22. ISBN 978-1-4939-6632-5. [Google Scholar]
  91. Durante-Mangoni, E.; Zarrilli, R. Global Spread of Drug-Resistant Acinetobacter Baumannii: Molecular Epidemiology and Management of Antimicrobial Resistance. Future Microbiol. 2011, 6, 407–422. [Google Scholar] [CrossRef]
  92. Ndowa, F.; Lusti-Narasimhan, M.; Unemo, M. The Serious Threat of Multidrug-Resistant and Untreatable Gonorrhoea: The Pressing Need for Global Action to Control the Spread of Antimicrobial Resistance, and Mitigate the Impact on Sexual and Reproductive Health. Sex. Transm. Infect. 2012, 88, 317–318. [Google Scholar] [CrossRef] [PubMed]
  93. Xu, X.; Wang, S.; Wu, H.; Liu, Y.; Xu, F.; Zhao, J. A Multimodal Antimicrobial Platform Based on MXene for Treatment of Wound Infection. Colloids Surf. B Biointerfaces 2021, 207, 111979. [Google Scholar] [CrossRef] [PubMed]
  94. Park, H.; Kim, J.-U.; Kim, S.; Hwang, N.S.; Kim, H.D. Sprayable Ti3C2 MXene Hydrogel for Wound Healing and Drug Release System. Mater. Today Bio 2023, 23, 100881. [Google Scholar] [CrossRef] [PubMed]
  95. Yin, J.; Han, Q.; Zhang, J.; Liu, Y.; Gan, X.; Xie, K.; Xie, L.; Deng, Y. MXene-Based Hydrogels Endow Polyetheretherketone with Effective Osteogenicity and Combined Treatment of Osteosarcoma and Bacterial Infection. ACS Appl. Mater. Interfaces 2020, 12, 45891–45903. [Google Scholar] [CrossRef] [PubMed]
  96. Liang, C.; Wang, H.; Lin, Z.; Zhang, C.; Liu, G.; Hu, Y. Augmented Wound Healing Potential of Photosensitive GelMA Hydrogel Incorporating Antimicrobial Peptides and MXene Nanoparticles. Front. Bioeng. Biotechnol. 2023, 11, 1310349. [Google Scholar] [CrossRef] [PubMed]
  97. Li, Z.; Wei, W.; Zhang, M.; Guo, X.; Zhang, B.; Wang, D.; Jiang, X.; Liu, F.; Tang, J. Cryptotanshinone-Doped Photothermal Synergistic MXene@PDA Nanosheets with Antibacterial and Anti-Inflammatory Properties for Wound Healing. Adv. Healthc. Mater. 2023, 12, 2301060. [Google Scholar] [CrossRef] [PubMed]
  98. Fang, Z.; Zhou, Q.; Zhang, W.; Wang, J.; Liu, Y.; Yu, M.; Qiu, Y.; Ma, Z.; Liu, S. A Synergistic Antibacterial Study of Copper-Doped Polydopamine on Ti3C2Tx Nanosheets with Enhanced Photothermal and Fenton-like Activities. Materials 2023, 16, 7583. [Google Scholar] [CrossRef] [PubMed]
  99. Zhu, B.; Song, P.; Li, J.; Cao, S.; Shi, J. Ti3C2 MXene/Gold Nanorods-Based Hybrid Nanoparticles with Photodynamic Antibacterial Activities. J. Mater. Sci. 2022, 57, 19957–19971. [Google Scholar] [CrossRef]
  100. Zhang, L.; Zhang, H.; Zhou, H.; Tan, Y.; Zhang, Z.; Yang, W.; Zhao, L.; Zhao, Z. A Ti3C2 MXene-Integrated near-Infrared-Responsive Multifunctional Porous Scaffold for Infected Bone Defect Repair. J. Mater. Chem. B 2024, 12, 79–96. [Google Scholar] [CrossRef]
  101. Su, Y.; Zhang, X.; Wei, Y.; Gu, Y.; Xu, H.; Liao, Z.; Zhao, L.; Du, J.; Hu, Y.; Lian, X.; et al. Nanocatalytic Hydrogel with Rapid Photodisinfection and Robust Adhesion for Fortified Cutaneous Regeneration. ACS Appl. Mater. Interfaces 2023, 15, 6354–6370. [Google Scholar] [CrossRef]
  102. Li, Y.; Fu, R.; Duan, Z.; Zhu, C.; Fan, D. Artificial Nonenzymatic Antioxidant MXene Nanosheet-Anchored Injectable Hydrogel as a Mild Photothermal-Controlled Oxygen Release Platform for Diabetic Wound Healing. ACS Nano 2022, 16, 7486–7502. [Google Scholar] [CrossRef] [PubMed]
  103. Zhou, L.; Zheng, H.; Liu, Z.; Wang, S.; Liu, Z.; Chen, F.; Zhang, H.; Kong, J.; Zhou, F.; Zhang, Q. Conductive Antibacterial Hemostatic Multifunctional Scaffolds Based on Ti3C2Tx MXene Nanosheets for Promoting Multidrug-Resistant Bacteria-Infected Wound Healing. ACS Nano 2021, 15, 2468–2480. [Google Scholar] [CrossRef] [PubMed]
  104. Wei, H.; Yang, L.; Pang, C.; Lian, L.; Hong, L. Bacteria-Targeted Photothermal Therapy for Combating Drug-Resistant Bacterial Infections. Biomater. Sci. 2023, 11, 5634–5640. [Google Scholar] [CrossRef] [PubMed]
  105. 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] [PubMed]
  106. Diedkova, K.; Pogrebnjak, A.D.; Kyrylenko, S.; Smyrnova, K.; Buranich, V.V.; Horodek, P.; Zukowski, P.; Koltunowicz, T.N.; Galaszkiewicz, P.; Makashina, K.; et al. Polycaprolactone–MXene Nanofibrous Scaffolds for Tissue Engineering. ACS Appl. Mater. Interfaces 2023, 15, 14033–14047. [Google Scholar] [CrossRef] [PubMed]
  107. Zeng, W.; Cheng, N.; Liang, X.; Hu, H.; Luo, F.; Jin, J.; Li, Y. Electrospun Polycaprolactone Nanofibrous Membranes Loaded with Baicalin for Antibacterial Wound Dressing. Sci. Rep. 2022, 12, 10900. [Google Scholar] [CrossRef]
  108. Xu, S.; Du, C.; Zhang, M.; Wang, R.; Feng, W.; Wang, C.; Liu, Q.; Zhao, W. Electroactive and Antibacterial Wound Dressings Based on Ti3C2Tx MXene/Poly(ε-Caprolactone)/Gelatin Coaxial Electrospun Nanofibrous Membranes. Nano Res. 2023, 16, 9672–9687. [Google Scholar] [CrossRef]
  109. Yu, C.; Sui, S.; Yu, X.; Huang, W.; Wu, Y.; Zeng, X.; Chen, Q.; Wang, J.; Peng, Q. Ti3C2Tx MXene Loaded with Indocyanine Green for Synergistic Photothermal and Photodynamic Therapy for Drug-Resistant Bacterium. Colloids Surf. B Biointerfaces 2022, 217, 112663. [Google Scholar] [CrossRef]
  110. Liu, X.; Xie, H.; Zhuo, S.; Zhou, Y.; Selim, M.S.; Chen, X.; Hao, Z. Ru(II) Complex Grafted Ti3C2Tx MXene Nano Sheet with Photothermal/Photodynamic Synergistic Antibacterial Activity. Nanomaterials 2023, 13, 958. [Google Scholar] [CrossRef]
  111. Cheng, H.; Wang, J.; Yang, Y.; Shi, H.; Shi, J.; Jiao, X.; Han, P.; Yao, X.; Chen, W.; Wei, X.; et al. Ti3C2TX MXene Modified with ZnTCPP with Bacteria Capturing Capability and Enhanced Visible Light Photocatalytic Antibacterial Activity. Small 2022, 18, 2200857. [Google Scholar] [CrossRef]
  112. Wang, H.; Mu, N.; He, Y.; Zhang, X.; Lei, J.; Yang, C.; Ma, L.; Gao, Y. Ultrasound-Controlled MXene-Based Schottky Heterojunction Improves Anti-Infection and Osteogenesis Properties. Theranostics 2023, 13, 1669–1683. [Google Scholar] [CrossRef]
  113. Jin, C.; Sun, D.; Sun, Z.; Rao, S.; Wu, Z.; Cheng, C.; Liu, L.; Liu, Q.; Yang, J. Interfacial Engineering of Ni-Phytate and Ti3C2Tx MXene-Sensitized TiO2 toward Enhanced Sterilization Efficacy under 808 Nm NIR Light Irradiation. Appl. Catal. B Environ. 2023, 330, 122613. [Google Scholar] [CrossRef]
  114. Rozmysłowska-Wojciechowska, A.; Mitrzak, J.; Szuplewska, A.; Chudy, M.; Woźniak, J.; Petrus, M.; Wojciechowski, T.; Vasilchenko, A.S.; Jastrzębska, A.M. Engineering of 2D Ti3C2 MXene Surface Charge and Its Influence on Biological Properties. Materials 2020, 13, 2347. [Google Scholar] [CrossRef] [PubMed]
  115. Zhang, D.; Liu, L.; Zhao, B.; Wang, X.; Pang, H.; Yu, S. Highly Efficient Extraction of Uranium from Seawater by Polyamide and Amidoxime Co-Functionalized MXene. Environ. Pollut. 2023, 317, 120826. [Google Scholar] [CrossRef] [PubMed]
  116. Jakubczak, M.; Karwowska, E.; Rozmysłowska-Wojciechowska, A.; Petrus, M.; Woźniak, J.; Mitrzak, J.; Jastrzębska, A. Filtration Materials Modified with 2D Nanocomposites—A New Perspective for Point-of-Use Water Treatment. Materials 2021, 14, 182. [Google Scholar] [CrossRef]
  117. Wei, J.; Jia, S.; Wei, J.; Ma, C.; Shao, Z. Tough and Multifunctional Composite Film Actuators Based on Cellulose Nanofibers toward Smart Wearables. ACS Appl. Mater. Interfaces 2021, 13, 38700–38711. [Google Scholar] [CrossRef]
  118. Mao, Z.; Hao, W.; Wang, W.; Ma, F.; Ma, C.; Chen, S. BiOI@CeO2@Ti3C2 MXene Composite S-Scheme Photocatalyst with Excellent Bacteriostatic Properties. J. Colloid Interface Sci. 2023, 633, 836–850. [Google Scholar] [CrossRef]
  119. Rozmysłowska-Wojciechowska, A.; Karwowska, E.; Poźniak, S.; Wojciechowski, T.; Chlubny, L.; Olszyna, A.; Ziemkowska, W.; Jastrzębska, A.M. Influence of Modification of Ti3C2 MXene with Ceramic Oxide and Noble Metal Nanoparticles on Its Antimicrobial Properties and Ecotoxicity towards Selected Algae and Higher Plants. RSC Adv. 2019, 9, 4092–4105. [Google Scholar] [CrossRef] [PubMed]
  120. Feng, H.; Wang, W.; Zhang, M.; Zhu, S.; Wang, Q.; Liu, J.; Chen, S. 2D Titanium Carbide-Based Nanocomposites for Photocatalytic Bacteriostatic Applications. Appl. Catal. B Environ. 2020, 266, 118609. [Google Scholar] [CrossRef]
  121. Jin, C.; Rao, S.; Xie, J.; Sun, Z.; Gao, J.; Li, Y.; Li, B.; Liu, S.; Liu, L.; Liu, Q.; et al. Enhanced Photocatalytic Antibacterial Performance by Hierarchical TiO2/W18O49 Z-Scheme Heterojunction with Ti3C2Tx-MXene Cocatalyst. Chem. Eng. J. 2022, 447, 137369. [Google Scholar] [CrossRef]
  122. Asadi Tokmedash, M.; Nagpal, N.; Chen, P.; VanEpps, J.S.; Min, J. Stretchable, Nano-Crumpled MXene Multilayers Impart Long-Term Antibacterial Surface Properties. Adv. Mater. Interfaces 2023, 10, 2202350. [Google Scholar] [CrossRef]
  123. Guo, H.; Niu, H.-Y.; Wang, W.-J.; Wu, Y.; Xiong, T.; Chen, Y.-R.; Su, C.-Q.; Niu, C.-G. Schottky Barrier Height Mediated Ti3C2 MXene Based Heterostructure for Rapid Photocatalytic Water Disinfection: Antibacterial Efficiency and Reaction Mechanism. Sep. Purif. Technol. 2023, 312, 123412. [Google Scholar] [CrossRef]
  124. Dai, W.; Shu, R.; Yang, F.; Li, B.; Johnson, H.M.; Yu, S.; Yang, H.; Chan, Y.K.; Yang, W.; Bai, D.; et al. Engineered Bio-Heterojunction Confers Extra- and Intracellular Bacterial Ferroptosis and Hunger-Triggered Cell Protection for Diabetic Wound Repair. Adv. Mater. 2024, 36, 2305277. [Google Scholar] [CrossRef] [PubMed]
  125. Du, T.; Zhao, S.; Dong, W.; Ma, W.; Zhou, X.; Wang, Y.; Zhang, M. Surface Modification of Carbon Fiber-Reinforced Polyetheretherketone with MXene Nanosheets for Enhanced Photothermal Antibacterial Activity and Osteogenicity. ACS Biomater. Sci. Eng. 2022, 8, 2375–2389. [Google Scholar] [CrossRef] [PubMed]
  126. Zheng, K.; Li, S.; Jing, L.; Chen, P.; Xie, J. Synergistic Antimicrobial Titanium Carbide (MXene) Conjugated with Gold Nanoclusters. Adv. Healthc. Mater. 2020, 9, 2001007. [Google Scholar] [CrossRef] [PubMed]
  127. Yu, Y.; Sun, H.; Lu, Q.; Sun, J.; Zhang, P.; Zeng, L.; Vasilev, K.; Zhao, Y.; Chen, Y.; Liu, P. Spontaneous Formation of MXene-Oxidized Sono/Chemo-Dynamic Sonosensitizer/Nanocatalyst for Antibacteria and Bone-Tissue Regeneration. J. Nanobiotechnology 2023, 21, 193. [Google Scholar] [CrossRef] [PubMed]
  128. Nie, Y.; Huang, J.; Ma, S.; Li, Z.; Shi, Y.; Yang, X.; Fang, X.; Zeng, J.; Bi, P.; Qi, J.; et al. MXene-Hybridized Silane Films for Metal Anticorrosion and Antibacterial Applications. Appl. Surf. Sci. 2020, 527, 146915. [Google Scholar] [CrossRef]
  129. Hu, Y.; Xu, Z.; Pu, J.; Hu, L.; Zi, Y.; Wang, M.; Feng, X.; Huang, W. 2D MXene Ti3C2Tx Nanosheets in the Development of a Mechanically Enhanced and Efficient Antibacterial Dental Resin Composite. Front. Chem. 2022, 10, 1090905. [Google Scholar] [CrossRef] [PubMed]
  130. Cheng, X.; Qin, X.; Su, Z.; Gou, X.; Yang, Z.; Wang, H. Research on the Antibacterial Properties of MXene-Based 2D–2D Composite Materials Membrane. Nanomaterials 2023, 13, 2121. [Google Scholar] [CrossRef]
  131. Wu, Q.; Tan, L.; Liu, X.; Li, Z.; Zhang, Y.; Zheng, Y.; Liang, Y.; Cui, Z.; Zhu, S.; Wu, S. The Enhanced Near-Infrared Photocatalytic and Photothermal Effects of MXene-Based Heterojunction for Rapid Bacteria-Killing. Appl. Catal. B Environ. 2021, 297, 120500. [Google Scholar] [CrossRef]
  132. Ku, M.; Mao, C.; Wu, S.; Zheng, Y.; Li, Z.; Cui, Z.; Zhu, S.; Shen, J.; Liu, X. Lattice Strain Engineering of Ti3C2 Narrows Band Gap for Realizing Extraordinary Sonocatalytic Bacterial Killing. ACS Nano 2023, 17, 14840–14851. [Google Scholar] [CrossRef] [PubMed]
  133. Lin, B.; Yuen, A.C.Y.; Oliver, S.; Liu, J.; Yu, B.; Yang, W.; Wu, S.; Yeoh, G.H.; Wang, C.H. Dual Functionalisation of Polyurethane Foam for Unprecedented Flame Retardancy and Antibacterial Properties Using Layer-by-Layer Assembly of MXene Chitosan with Antibacterial Metal Particles. Compos. Part B Eng. 2022, 244, 110147. [Google Scholar] [CrossRef]
  134. Wang, H.; Dong, A.; Hu, K.; Sun, W.; Wang, J.; Han, L.; Mo, L.; Li, L.; Zhang, W.; Guo, Y.; et al. LBL Assembly of Ag@Ti3C2TX and Chitosan on PLLA Substrate to Enhance Antibacterial and Biocompatibility. Biomed. Mater. 2022, 17, 035006. [Google Scholar] [CrossRef] [PubMed]
  135. Tang, S.; Wu, Z.; Feng, G.; Wei, L.; Weng, J.; Ruiz-Hitzky, E.; Wang, X. Multifunctional Sandwich-like Composite Film Based on Superhydrophobic MXene for Self-Cleaning, Photodynamic and Antimicrobial Applications. Chem. Eng. J. 2023, 454, 140457. [Google Scholar] [CrossRef]
  136. Yang, Y.; Zhou, X.; Chan, Y.K.; Wang, Z.; Li, L.; Li, J.; Liang, K.; Deng, Y. Photo-Activated Nanofibrous Membrane with Self-Rechargeable Antibacterial Function for Stubborn Infected Cutaneous Regeneration. Small 2022, 18, 2105988. [Google Scholar] [CrossRef] [PubMed]
  137. Wang, Z.; Du, B.; Gao, X.; Huang, Y.; Li, M.; Yu, Z.; Li, J.; Shi, X.; Deng, Y.; Liang, K. Endogenous Oxygen-Evolving Bio-Catalytic Fabrics with Fortified Photonic Disinfection for Invasive Bacteria-Caused Refractory Cutaneous Regeneration. Nano Today 2022, 46, 101595. [Google Scholar] [CrossRef]
  138. He, J.; Li, A.; Wang, W.; Cui, C.; Jiang, S.; Chen, M.; Qin, W.; Tang, H.; Guo, R. Multifunctional Wearable Device Based on an Antibacterial and Hydrophobic Silver Nanoparticles/Ti3C2Tx MXene/Thermoplastic Polyurethane Fibrous Membrane for Electromagnetic Shielding and Strain Sensing. Ind. Eng. Chem. Res. 2023, 62, 9221–9232. [Google Scholar] [CrossRef]
  139. Feng, Q.; Zhan, Y.; Yang, W.; Dong, H.; Sun, A.; Li, L.; Chen, X.; Chen, Y. Ultra-High Flux and Synergistically Enhanced Anti-Fouling Ag@MXene Lamellar Membrane for the Fast Purification of Oily Wastewater through Nano-Intercalation, Photocatalytic Self-Cleaning and Antibacterial Effect. Sep. Purif. Technol. 2022, 298, 121635. [Google Scholar] [CrossRef]
  140. Liu, H.; Liu, Y.; Wang, L.; Qin, X.; Yu, J. Nanofiber Based Origami Evaporator for Multifunctional and Omnidirectional Solar Steam Generation. Carbon 2021, 177, 199–206. [Google Scholar] [CrossRef]
  141. Mayerberger, E.A.; Street, R.M.; McDaniel, R.M.; Barsoum, M.W.; Schauer, C.L. Antibacterial Properties of Electrospun Ti3C2Tz (MXene)/Chitosan Nanofibers. RSC Adv. 2018, 8, 35386–35394. [Google Scholar] [CrossRef]
  142. Zhang, S.; Ye, J.; Liu, X.; Wang, Y.; Li, C.; Fang, J.; Chang, B.; Qi, Y.; Li, Y.; Ning, G. Titanium Carbide/Zeolite Imidazole Framework-8/Polylactic Acid Electrospun Membrane for near-Infrared Regulated Photothermal/Photodynamic Therapy of Drug-Resistant Bacterial Infections. J. Colloid Interface Sci. 2021, 599, 390–403. [Google Scholar] [CrossRef] [PubMed]
  143. Zhang, H.; Lan, D.; Li, X.; Li, Z.; Dai, F. Conductive and Antibacterial Scaffold with Rapid Crimping Property for Application Prospect in Repair of Peripheral Nerve Injury. J. Appl. Polym. Sci. 2023, 140, e53426. [Google Scholar] [CrossRef]
  144. Zhang, H.; Lan, D.; Wu, B.; Chen, X.; Li, X.; Li, Z.; Dai, F. Electrospun Piezoelectric Scaffold with External Mechanical Stimulation for Promoting Regeneration of Peripheral Nerve Injury. Biomacromolecules 2023, 24, 3268–3282. [Google Scholar] [CrossRef]
  145. Wang, L.; Du, L.; Wang, M.; Wang, X.; Tian, S.; Chen, Y.; Wang, X.; Zhang, J.; Nie, J.; Ma, G. Chitosan for Constructing Stable Polymer-Inorganic Suspensions and Multifunctional Membranes for Wound Healing. Carbohydr. Polym. 2022, 285, 119209. [Google Scholar] [CrossRef]
  146. Yan, B.; Bao, X.; Liao, X.; Wang, P.; Zhou, M.; Yu, Y.; Yuan, J.; Cui, L.; Wang, Q. Sensitive Micro-Breathing Sensing and Highly-Effective Photothermal Antibacterial Cinnamomum camphora Bark Micro-Structural Cotton Fabric via Electrostatic Self-Assembly of MXene/HACC. ACS Appl. Mater. Interfaces 2022, 14, 2132–2145. [Google Scholar] [CrossRef]
  147. Liu, Y.; Xu, D.; Ding, Y.; Lv, X.; Huang, T.; Yuan, B.; Jiang, L.; Sun, X.; Yao, Y.; Tang, J. A Conductive Polyacrylamide Hydrogel Enabled by Dispersion-Enhanced MXene@chitosan Assembly for Highly Stretchable and Sensitive Wearable Skin. J. Mater. Chem. B 2021, 9, 8862–8870. [Google Scholar] [CrossRef]
  148. Li, J.; Ma, L.; Li, Z.; Liu, X.; Zheng, Y.; Liang, Y.; Liang, C.; Cui, Z.; Zhu, S.; Wu, S. Oxygen Vacancies-Rich Heterojunction of Ti3C2/BiOBr for Photo-Excited Antibacterial Textiles. Small 2022, 18, 2104448. [Google Scholar] [CrossRef] [PubMed]
  149. Mi, X.; Su, Z.; Fu, Y.; Li, S.; Mo, A. 3D Printing of Ti3C2-MXene-Incorporated Composite Scaffolds for Accelerated Bone Regeneration. Biomed. Mater. 2022, 17, 035002. [Google Scholar] [CrossRef]
  150. Tan, Y.; Sun, H.; Lan, Y.; Khan, H.M.; Zhang, H.; Zhang, L.; Zhang, F.; Cui, Y.; Zhang, L.; Huang, D.; et al. Study on 3D Printed MXene-Berberine-Integrated Scaffold for Photo-Activated Antibacterial Activity and Bone Regeneration. J. Mater. Chem. B 2024, 12, 2158–2179. [Google Scholar] [CrossRef]
  151. Nie, R.; Sun, Y.; Lv, H.; Lu, M.; Huangfu, H.; Li, Y.; Zhang, Y.; Wang, D.; Wang, L.; Zhou, Y. 3D Printing of MXene Composite Hydrogel Scaffolds for Photothermal Antibacterial Activity and Bone Regeneration in Infected Bone Defect Models. Nanoscale 2022, 14, 8112–8129. [Google Scholar] [CrossRef]
  152. Gong, M.; Yue, L.; Kong, J.; Lin, X.; Zhang, L.; Wang, J.; Wang, D. Knittable and Sewable Spandex Yarn with Nacre-Mimetic Composite Coating for Wearable Health Monitoring and Thermo- and Antibacterial Therapies. ACS Appl. Mater. Interfaces 2021, 13, 9053–9063. [Google Scholar] [CrossRef] [PubMed]
  153. 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, 36589–36597. [Google Scholar] [CrossRef] [PubMed]
  154. Liu, W.; Kang, S.; Zhang, Q.; Chen, S.; Yang, Q.; Yan, B. Self-Assembly Fabrication of Chitosan-Tannic Acid/MXene Composite Film with Excellent Antibacterial and Antioxidant Properties for Fruit Preservation. Food Chem. 2023, 410, 135405. [Google Scholar] [CrossRef] [PubMed]
  155. Li, Y.; Han, M.; Cai, Y.; Jiang, B.; Zhang, Y.; Yuan, B.; Zhou, F.; Cao, C. Muscle-Inspired MXene/PVA Hydrogel with High Toughness and Photothermal Therapy for Promoting Bacteria-Infected Wound Healing. Biomater. Sci. 2022, 10, 1068–1082. [Google Scholar] [CrossRef] [PubMed]
  156. Zhang, D.; Huang, L.; Sun, D.-W.; Pu, H.; Wei, Q. Bio-Interface Engineering of MXene Nanosheets with Immobilized Lysozyme for Light-Enhanced Enzymatic Inactivation of Methicillin-Resistant Staphylococcus Aureus. Chem. Eng. J. 2023, 452, 139078. [Google Scholar] [CrossRef]
  157. Wu, Y.; Zheng, W.; Xiao, Y.; Du, B.; Zhang, X.; Wen, M.; Lai, C.; Huang, Y.; Sheng, L. Multifunctional, Robust, and Porous PHBV—GO/MXene Composite Membranes with Good Hydrophilicity, Antibacterial Activity, and Platelet Adsorption Performance. Polymers 2021, 13, 3748. [Google Scholar] [CrossRef] [PubMed]
  158. Chen, J.; Liu, Y.; Cheng, G.; Guo, J.; Du, S.; Qiu, J.; Wang, C.; Li, C.; Yang, X.; Chen, T.; et al. Tailored Hydrogel Delivering Niobium Carbide Boosts ROS-Scavenging and Antimicrobial Activities for Diabetic Wound Healing. Small 2022, 18, 2201300. [Google Scholar] [CrossRef]
  159. Ding, M.; Zhang, S.; Wang, J.; Ding, Y.; Ding, C. Ultrasensitive Ratiometric Electrochemiluminescence Sensor with an Efficient Antifouling and Antibacterial Interface of PSBMA@SiO2-MXene for Oxytetracycline Trace Detection in the Marine Environment. Anal. Chem. 2023, 95, 16327–16334. [Google Scholar] [CrossRef] [PubMed]
  160. Wu, P.; Qin, Z.; Dassanayake, R.; Sun, Z.; Cao, M.; Fu, K.; Zhou, Y.; Liu, Y. Antimicrobial MXene-Based Conductive Alginate Hydrogels as Flexible Electronics. Chem. Eng. J. 2023, 455, 140546. [Google Scholar] [CrossRef]
  161. Fan, Y.; Kong, F.; Yang, J.; Xiong, X.; Gao, S.; Yuan, J.; Meng, S.; Chen, L. Flexible Wearable Sensor Based on SF/EEP/GR/MXene Nanocomposites. Appl. Phys. A 2023, 129, 551. [Google Scholar] [CrossRef]
  162. Li, C.; Zheng, A.; Zhou, J.; Huang, W.; Zhang, Y.; Han, J.; Cao, L.; Yang, D. A Self-Adhesive, Self-Healing and Antibacterial Hydrogel Based on PVA/MXene-Ag/Sucrose for Fast-Response, High-Sensitivity and Ultra-Durable Strain Sensors. New J. Chem. 2023, 47, 6621–6630. [Google Scholar] [CrossRef]
  163. He, J.; Zou, X.; Wang, W.; Chen, M.; Jiang, S.; Cui, C.; Tang, H.; Yang, L.; Guo, R. Highly Stretchable, Highly Sensitive, and Antibacterial Electrospun Nanofiber Strain Sensors with Low Detection Limit and Stable CNT/MXene/CNT Sandwich Conductive Layers for Human Motion Detection. Ind. Eng. Chem. Res. 2023, 62, 8327–8338. [Google Scholar] [CrossRef]
  164. Cao, Y.-M.; Li, Y.-F.; Dong, X.-X.; Chen, J.; Zhang, K.-Q.; Zhao, Y.-D.; Zhai, W.-Y.; Zheng, M.; Zheng, M.; Wang, Z.-S.; et al. Knitted Structural Design of MXene/Cu2O Based Strain Sensor for Smart Wear. Cellulose 2022, 29, 9453–9467. [Google Scholar] [CrossRef]
  165. Wang, H.; Zou, Y.; Ji, Y.; Zhong, K.; Du, X.; Du, Z.; Cheng, X.; Wang, S. Tough and Extremely Temperature-Tolerance Nanocomposite Organohydrogels as Ultrasensitive Wearable Sensors for Wireless Human Motion Monitoring. Compos. Part Appl. Sci. Manuf. 2022, 157, 106905. [Google Scholar] [CrossRef]
  166. Zhao, W.; Jiang, J.; Chen, W.; He, Y.; Lin, T.; Zhao, L. Radiation Synthesis of Rapidly Self-Healing, Durable, and Flexible Poly(Ionic Liquid)/MXene Gels with Anti-Freezing Property for Multi-Functional Strain Sensors. Chem. Eng. J. 2023, 468, 143660. [Google Scholar] [CrossRef]
  167. Li, X.; Liu, Y.; Ding, Y.; Zhang, M.; Lin, Z.; Hao, Y.; Li, Y.; Chang, J. Capacitive Pressure Sensor Combining Dual Dielectric Layers with Integrated Composite Electrode for Wearable Healthcare Monitoring. ACS Appl. Mater. Interfaces 2024, 16, 12974–12985. [Google Scholar] [CrossRef] [PubMed]
  168. Wang, Z.; Zhang, X.; Cao, T.; Wang, T.; Sun, L.; Wang, K.; Fan, X. Antiliquid-Interfering, Antibacteria, and Adhesive Wearable Strain Sensor Based on Superhydrophobic and Conductive Composite Hydrogel. ACS Appl. Mater. Interfaces 2021, 13, 46022–46032. [Google Scholar] [CrossRef] [PubMed]
  169. Ma, J.; Pan, Z.; Zhang, W.; Fan, Q.; Li, W.; Liang, H. High-Sensitivity Microchannel-Structured Collagen Fiber-Based Sensors with Antibacterial and Hydrophobic Properties. ACS Sustain. Chem. Eng. 2022, 10, 16814–16824. [Google Scholar] [CrossRef]
  170. Fu, Y.; Li, C.; Cheng, Y.; He, Y.; Zhang, W.; Wei, Q.; Li, D. Biomass Aerogel Composite Containing BaTiO3 Nanoparticles and MXene for Highly Sensitive Self-Powered Sensor and Photothermal Antibacterial Applications. Compos. Part Appl. Sci. Manuf. 2023, 173, 107663. [Google Scholar] [CrossRef]
  171. Wang, X.; Tao, Y.; Pan, S.; Fang, X.; Lou, C.; Xu, Y.; Wu, J.; Sang, M.; Lu, L.; Gong, X.; et al. Biocompatible and Breathable Healthcare Electronics with Sensing Performances and Photothermal Antibacterial Effect for Motion-Detecting. Npj Flex. Electron. 2022, 6, 95. [Google Scholar] [CrossRef]
  172. Zhu, T.; Jiang, W.; Shi, X.; Sun, D.; Yang, J.; Qi, X.; Wang, Y. MXene/Ag Doped Hydrated-Salt Hydrogels with Excellent Thermal/Light Energy Storage, Strain Sensing and Photothermal Antibacterial Performances for Intelligent Human Healthcare. Compos. Part Appl. Sci. Manuf. 2023, 170, 107526. [Google Scholar] [CrossRef]
  173. Li, M.; Zhang, Y.; Lian, L.; Liu, K.; Lu, M.; Chen, Y.; Zhang, L.; Zhang, X.; Wan, P. Flexible Accelerated-Wound-Healing Antibacterial MXene-Based Epidermic Sensor for Intelligent Wearable Human-Machine Interaction. Adv. Funct. Mater. 2022, 32, 2208141. [Google Scholar] [CrossRef]
  174. Wang, S.; Bi, S.; Zhang, L.; Liu, R.; Wang, H.; Gu, J. Skin-Inspired Antibacterial Conductive Hydrogels Customized for Wireless Flexible Sensor and Collaborative Wound Healing. J. Mater. Chem. A 2023, 11, 14096–14107. [Google Scholar] [CrossRef]
  175. Zhang, J.; Fu, Y.; Mo, A. Multilayered Titanium Carbide MXene Film for Guided Bone Regeneration. Int. J. Nanomed. 2019, 14, 10091–10103. [Google Scholar] [CrossRef]
  176. Mazinani, A.; Rastin, H.; Nine, M.J.; Lee, J.; Tikhomirova, A.; Tung, T.T.; Ghomashchi, R.; Kidd, S.; Vreugde, S.; Losic, D. Comparative Antibacterial Activity of 2D Materials Coated on Porous-Titania. J. Mater. Chem. B 2021, 9, 6412–6424. [Google Scholar] [CrossRef] [PubMed]
  177. Huang, S.; Fu, Y.; Mo, A. Electrophoretic-Deposited MXene Titanium Coatings in Regulating Bacteria and Cell Response for Peri-Implantitis. Front. Chem. 2022, 10, 991481. [Google Scholar] [CrossRef] [PubMed]
  178. Liu, Y.; Tian, Y.; Han, Q.; Yin, J.; Zhang, J.; Yu, Y.; Yang, W.; Deng, Y. Synergism of 2D/1D MXene/Cobalt Nanowire Heterojunctions for Boosted Photo-Activated Antibacterial Application. Chem. Eng. J. 2021, 410, 128209. [Google Scholar] [CrossRef]
  179. Wang, Q.; Zang, C.; Chan, Y.K.; Wang, S.; Yang, W.; Deng, Y.; Zhu, T.; He, M. CoFe2O4/MXene Nanosheets Modified Hydrogel on PEEK with Phototherapeutic and GPx-Mimetic Activities for Anti-Pathogens in Infectious Bone Defect Repairment. Mater. Chem. Phys. 2023, 308, 128269. [Google Scholar] [CrossRef]
  180. Li, K.; Chang, T.; Li, Z.; Yang, H.; Fu, F.; Li, T.; Ho, J.S.; Chen, P. Biomimetic MXene Textures with Enhanced Light-to-Heat Conversion for Solar Steam Generation and Wearable Thermal Management. Adv. Energy Mater. 2019, 9, 1901687. [Google Scholar] [CrossRef]
  181. Chen, J.; Li, B.; Hu, G.; Aleisa, R.; Lei, S.; Yang, F.; Liu, D.; Lyu, F.; Wang, M.; Ge, X.; et al. Integrated Evaporator for Efficient Solar-Driven Interfacial Steam Generation. Nano Lett. 2020, 20, 6051–6058. [Google Scholar] [CrossRef]
  182. Dixit, F.; Zimmermann, K.; Dutta, R.; Prakash, N.J.; Barbeau, B.; Mohseni, M.; Kandasubramanian, B. Application of MXenes for Water Treatment and Energy-Efficient Desalination: A Review. J. Hazard. Mater. 2022, 423, 127050. [Google Scholar] [CrossRef] [PubMed]
  183. Tao, Y.; Mi, Y.; Gao, S.; Wang, G.; Bai, J.; Ma, S.; Wang, B. High-Efficiency, Thermal Stable, and Self-Floating Silver Nanowires/Ti3C2Tx MXene/Aramid Nanofibers Composite Aerogel for Photothermal Water Evaporation and Antibacterial Application. Chem. Eng. J. 2023, 477, 147276. [Google Scholar] [CrossRef]
  184. Ma, X.; Tian, N.; Wang, G.; Wang, W.; Miao, J.; Fan, T. Biomimetic Vertically Aligned Aerogel with Synergistic Photothermal Effect Enables Efficient Solar-Driven Desalination. Desalination 2023, 550, 116397. [Google Scholar] [CrossRef]
  185. Li, Y.; Wu, T.; Shen, H.; Yang, S.; Qin, Y.; Zhu, Z.; Zheng, L.; Wen, X.; Xia, M.; Yin, X. Flexible MXene-Based Janus Porous Fibrous Membranes for Sustainable Solar-Driven Desalination and Emulsions Separation. J. Clean. Prod. 2022, 347, 131324. [Google Scholar] [CrossRef]
  186. Guan, H.-S.; Fan, T.-T.; Bai, H.-Y.; Su, Y.; Liu, Z.; Ning, X.; Yu, M.; Ramakrishna, S.; Long, Y.-Z. A Waste Biomass-Derived Photothermic Material with High Salt-Resistance for Efficient Solar Evaporation. Carbon 2022, 188, 265–275. [Google Scholar] [CrossRef]
  187. Rasheed, T.; Rasheed, A.; Munir, S.; Ajmal, S.; Muhammad Shahzad, Z.; Alsafari, I.A.; Ragab, S.A.; Agboola, P.O.; Shakir, I. A Cost-Effective Approach to Synthesize NiFe2O4/MXene Heterostructures for Enhanced Photodegradation Performance and Anti-Bacterial Activity. Adv. Powder Technol. 2021, 32, 2248–2257. [Google Scholar] [CrossRef]
  188. Rasool, K.; Mahmoud, K.A.; Johnson, D.J.; Helal, M.; Berdiyorov, G.R.; Gogotsi, Y. Efficient Antibacterial Membrane Based on Two-Dimensional Ti3C2Tx (MXene) Nanosheets. Sci. Rep. 2017, 7, 1598. [Google Scholar] [CrossRef] [PubMed]
  189. Qin, X.; Ding, C.; Tian, Y.; Dong, J.; Cheng, B. Multifunctional Ti3C2Tx MXene/Silver Nanowire Membranes with Excellent Catalytic, Antifouling, and Antibacterial Properties for Nitrophenol-Containing Water Purification. ACS Appl. Mater. Interfaces 2023, 15, 48154–48167. [Google Scholar] [CrossRef] [PubMed]
  190. Pandey, R.P.; Rasool, K.; Madhavan, V.E.; Aïssa, B.; Gogotsi, Y.; Mahmoud, K.A. Ultrahigh-Flux and Fouling-Resistant Membranes Based on Layered Silver/MXene (Ti3C2Tx) Nanosheets. J. Mater. Chem. A 2018, 6, 3522–3533. [Google Scholar] [CrossRef]
  191. Mansoorianfar, M.; Shahin, K.; Hojjati–Najafabadi, A.; Pei, R. MXene–Laden Bacteriophage: A New Antibacterial Candidate to Control Bacterial Contamination in Water. Chemosphere 2022, 290, 133383. [Google Scholar] [CrossRef]
  192. Mao, Z.; Peng, X.; Chen, H. Sunlight Propelled Two-Dimensional Nanorobots with Enhanced Mechanical Damage of Bacterial Membrane. Water Res. 2023, 235, 119900. [Google Scholar] [CrossRef] [PubMed]
  193. Liu, X.; Jin, X.; Li, L.; Wang, J.; Yang, Y.; Cao, Y.; Wang, W. Air-Permeable, Multifunctional, Dual-Energy-Driven MXene-Decorated Polymeric Textile-Based Wearable Heaters with Exceptional Electrothermal and Photothermal Conversion Performance. J. Mater. Chem. A 2020, 8, 12526–12537. [Google Scholar] [CrossRef]
  194. Liu, X.; Du, X.; Li, L.; Cao, Y.; Yang, Y.; Wang, W.; Wang, J. Multifunctional AgNW@MXene Decorated Polymeric Textile for Highly-Efficient Electro-/Photothermal Conversion and Triboelectric Nanogenerator. Compos. Part Appl. Sci. Manuf. 2022, 156, 106883. [Google Scholar] [CrossRef]
  195. Tang, L.; Lyu, B.; Gao, D.; Jia, Z.; Fu, Y.; Ma, J. A Janus Textile with Tunable Heating Modes toward Precise Personal Thermal Management in Cold Conditions. Small 2024, 20, 2308194. [Google Scholar] [CrossRef] [PubMed]
  196. Purbayanto, M.A.K.; Jakubczak, M.; Bury, D.; Nair, V.G.; Birowska, M.; Moszczyńska, D.; Jastrzębska, A. Tunable Antibacterial Activity of a Polypropylene Fabric Coated with Bristling Ti3C2Tx MXene Flakes Coupling the Nanoblade Effect with ROS Generation. ACS Appl. Nano Mater. 2022, 5, 5373–5386. [Google Scholar] [CrossRef]
  197. Deng, C.; Yu, Z.; Liang, F.; Liu, Y.; Seidi, F.; Yong, Q.; Liu, C.; Zhang, Y.; Han, J.; Xiao, H. Surface Nano-Engineering of Cellulosic Textiles for Superior Biocidal Performance and Effective Bacterial Detection. Chem. Eng. J. 2023, 473, 145492. [Google Scholar] [CrossRef]
  198. Zhai, W.; Cao, Y.; Li, Y.; Zheng, M.; Wang, Z. MoO3–x QDs/MXene (Ti3C2Tx) Self-Assembled Heterostructure for Multifunctional Application with Antistatic, Smoke Suppression, and Antibacterial on Polyester Fabric. J. Mater. Sci. 2022, 57, 2597–2609. [Google Scholar] [CrossRef]
  199. Su, X.; Sha, Q.; Gao, X.; Li, J.; Wu, Y.; Li, W.; Wu, W.; Han, N.; Zhang, X. Lightweight, Multifunctional Smart MXene@PET Non-Woven with Electric/Photothermal Conversion, Antibacterial and Flame Retardant Properties. Appl. Surf. Sci. 2023, 639, 158205. [Google Scholar] [CrossRef]
  200. Santos, X.; Álvarez, M.; Videira-Quintela, D.; Mediero, A.; Rodríguez, J.; Guillén, F.; Pozuelo, J.; Martín, O. Antibacterial Capability of MXene (Ti3C2Tx) to Produce PLA Active Contact Surfaces for Food Packaging Applications. Membranes 2022, 12, 1146. [Google Scholar] [CrossRef]
  201. Cao, Y.; Chang, T.; Fang, C.; Zhang, Y.; Liu, H.; Zhao, G. Inhibition Effect of Ti3C2Tx MXene on Ice Crystals Combined with Laser-Mediated Heating Facilitates High-Performance Cryopreservation. ACS Nano 2022, 16, 8837–8850. [Google Scholar] [CrossRef]
  202. Zhang, W.-J.; Li, S.; Yan, Y.-Z.; Park, S.S.; Mohan, A.; Chung, I.; Ahn, S.; Kim, J.R.; Ha, C.-S. Dual (pH- and ROS-) Responsive Antibacterial MXene-Based Nanocarrier for Drug Delivery. Int. J. Mol. Sci. 2022, 23, 14925. [Google Scholar] [CrossRef] [PubMed]
  203. Bose, N.; Danagody, B.; Rajappan, K.; Ramanujam, G.M.; Anilkumar, A.K. Sustainable Routed Mxene-Based Aminolyzed PU/PCL Film for Increased Oxidative Stress and a pH-Sensitive Drug Delivery System for Anticancer Therapy. ACS Appl. Bio Mater. 2024, 7, 379–393. [Google Scholar] [CrossRef] [PubMed]
  204. Qin, X.; Wu, Z.; Fang, J.; Li, S.; Tang, S.; Wang, X. MXene-Intercalated Montmorillonite Nanocomposites for Long-Acting Antibacterial. Appl. Surf. Sci. 2023, 616, 156521. [Google Scholar] [CrossRef]
  205. Liu, L.; Zhu, M.; Ma, Z.; Xu, X.; Mohesen Seraji, S.; Yu, B.; Sun, Z.; Wang, H.; Song, P. A Reactive Copper-Organophosphate-MXene Heterostructure Enabled Antibacterial, Self-Extinguishing and Mechanically Robust Polymer Nanocomposites. Chem. Eng. J. 2022, 430, 132712. [Google Scholar] [CrossRef]
Figure 1. General applications of Ti3C2Tx composites in antibacterial field.
Figure 1. General applications of Ti3C2Tx composites in antibacterial field.
Molecules 29 02902 g001
Figure 2. Ti3C2Tx/metal composites [59]. (A) Synthesis and characterization of Cu(II)@MXene complex. (a) Schematic diagram of the synthesis process of Cu(II)@MXene complex. (b) Zeta potential of MXene, CuCl2, and Cu(II)@MXene. (c) Concentration of Cu2+ in the supernatant before and after electrostatic self-assemble between Cu2+ and MXene. (d) The SEM images of Cu(II)@MXene at different time points. (B) Antibacterial properties of Cu(II)@MXene hydrogel and promotion of infected wound healing in vivo. Total view of bacterial colonies formed by E. coli (a) and S. aureus (b) after different hydrogel treatments. (c) Schematic diagram of the construction of mouse skin infection model. (C) In vitro cytocom-patibility and angiogenesis property of Cu(II)@MXene hydrogel. (a) Live/dead staining images of the L929 cells after being treated with different materials for 24 h. (b) CCK-8 results of the L929 cells after treatment with different materials for 24 h. (c) Tube formation assay of HUVECs treated by different materials. The quantitative analysis of formed (d) branch points and (e) tube length, respectively. Statistic results: * p < 0.05. ** p < 0.01.
Figure 2. Ti3C2Tx/metal composites [59]. (A) Synthesis and characterization of Cu(II)@MXene complex. (a) Schematic diagram of the synthesis process of Cu(II)@MXene complex. (b) Zeta potential of MXene, CuCl2, and Cu(II)@MXene. (c) Concentration of Cu2+ in the supernatant before and after electrostatic self-assemble between Cu2+ and MXene. (d) The SEM images of Cu(II)@MXene at different time points. (B) Antibacterial properties of Cu(II)@MXene hydrogel and promotion of infected wound healing in vivo. Total view of bacterial colonies formed by E. coli (a) and S. aureus (b) after different hydrogel treatments. (c) Schematic diagram of the construction of mouse skin infection model. (C) In vitro cytocom-patibility and angiogenesis property of Cu(II)@MXene hydrogel. (a) Live/dead staining images of the L929 cells after being treated with different materials for 24 h. (b) CCK-8 results of the L929 cells after treatment with different materials for 24 h. (c) Tube formation assay of HUVECs treated by different materials. The quantitative analysis of formed (d) branch points and (e) tube length, respectively. Statistic results: * p < 0.05. ** p < 0.01.
Molecules 29 02902 g002
Figure 3. Ti3C2Tx/organic frameworks [87]. (A,B) Design and characterization of S-CF@PCNx. (A) (a) Schematic illustration of the production of S-CF@PCNx. (be) XRD characterization of PCN-224, CF@PCN, Ti3AlC2 MAX, and Ti3C2 MXene. (f) Digital photograph showing the Tyndall scattering effect of the Ti3C2 MXene nanosheet suspension. (g) TEM image of Ti3C2 MXene nanosheets. (B) (ac) Digital photograph and SEM images of CF@PCN. (df) Digital photograph and SEM images of [email protected]. (g) Digital photograph of [email protected]. (hi) Surface and cross section observation of [email protected]. (C) Photodynamic antibacterial properties. (a,b) Antibacterial photodynamic inactivation study against E. coli and S. aureus. (c) Reusability of antibacterial activity of the materials against E. coli under 30 min illumination. (d) Antibacterial activity against E. coli after one and two washes under 30 min illumination. The light source for bacterial inactivation was provided by a Xe lamp (500 W, 15 cm sample distance, λ ≥ 420 nm). (D) Photothermal and photothermochromic effect. (a) Surface thermal and digital images under Xe lamp illumination (500 W, 15 cm sample distance, λ ≥ 420 nm). Left top: [email protected]; right top: [email protected]; left bottom: S-CF@PCN; and right bottom: CF. (b) Temperature rise curves of CF, S-CF@PCN, [email protected], and [email protected]. (c) Photothermal heating curves of CF, S-CF@PCN, [email protected], and [email protected] for five cycles. (d) Thermal diffusivity of all of the prepared fabrics.
Figure 3. Ti3C2Tx/organic frameworks [87]. (A,B) Design and characterization of S-CF@PCNx. (A) (a) Schematic illustration of the production of S-CF@PCNx. (be) XRD characterization of PCN-224, CF@PCN, Ti3AlC2 MAX, and Ti3C2 MXene. (f) Digital photograph showing the Tyndall scattering effect of the Ti3C2 MXene nanosheet suspension. (g) TEM image of Ti3C2 MXene nanosheets. (B) (ac) Digital photograph and SEM images of CF@PCN. (df) Digital photograph and SEM images of [email protected]. (g) Digital photograph of [email protected]. (hi) Surface and cross section observation of [email protected]. (C) Photodynamic antibacterial properties. (a,b) Antibacterial photodynamic inactivation study against E. coli and S. aureus. (c) Reusability of antibacterial activity of the materials against E. coli under 30 min illumination. (d) Antibacterial activity against E. coli after one and two washes under 30 min illumination. The light source for bacterial inactivation was provided by a Xe lamp (500 W, 15 cm sample distance, λ ≥ 420 nm). (D) Photothermal and photothermochromic effect. (a) Surface thermal and digital images under Xe lamp illumination (500 W, 15 cm sample distance, λ ≥ 420 nm). Left top: [email protected]; right top: [email protected]; left bottom: S-CF@PCN; and right bottom: CF. (b) Temperature rise curves of CF, S-CF@PCN, [email protected], and [email protected]. (c) Photothermal heating curves of CF, S-CF@PCN, [email protected], and [email protected] for five cycles. (d) Thermal diffusivity of all of the prepared fabrics.
Molecules 29 02902 g003
Figure 4. Ti3C2Tx/antibiotic composite materials [94]. (A) Preparation and characterization of the MXene/SA hydrogel (MX@Gel). (a) Temperature change of MX@Gel containing MXene at various concentrations (0, 10, 50, 100 μg/mL) under near-infrared (NIR) laser irradiation (808 nm, 0.5 W/cm2). (b) Heating and cooling curves of MX@Gel by laser on/off cycle. (c) MX@Gel temperature change graph according to various laser densities (0.50, 1.00, 1.50, 2.00 W/cm2). (d) Schematic of loading ciprofloxacin on MXene surface. (e) UV–vis absorption spectrum of CIP-loaded MXene (CIP-MX). (B) Biocompatibility and cell migration of the CIP-MX@Gel with antibacterial property. (a) Optical microscopy and Live/dead staining images of human dermal fibroblasts cultured on MX@Gel at 1, 4, and 7 days. (b) quantification of cell viability of MX@Gels. (c) Wound healing scratch assay of MX@Gel at different concentrations. (d) Quantification of cell migration (n = 3). (e) Gene expression profiles of MX@Gels (n = 3). (f) Antibacterial gross images of E. coli and S. aureus in CIP-MX@Gel using smear plate method. (g) Quantification of the zone of inhibition of CIP-MX@Gel (n = 3). All data represent mean ± SD. * p < 0.05, ** p < 0.01. The symbol * indicates comparisons with 0 μg/mL. Scale bars = 100 μm (a,c). (C) In vivo wound healing effect of CIP-MX@Gel in mice. (a) Schematic diagram of the formation process of sprayable CIP-MX@Gel, and drug release by NIR irradiation (b) Schematic diagram of the experimental healing process in vivo. (c) Wound images at days 0, 3, 7, 10 and 14 (Silicone isolator diameter = 9 mm). (d) quantification of relative wound area. (n ≥ 3) (e) Hematoxylin and eosin (H&E) staining images on day 7 and 14 of wound tissues treated in different groups. (f,g) Quantification of the panniculus gap at day 7 and day 14 in wound tissue treated with different groups. (h,i) Quantification of the granulation tissue thickness at day 7 and day 14 in wound tissue treated with different groups. All data represent mean ± SD. * p < 0.05, ** p < 0.01. The symbol * indicates comparisons with Sham. Scale bars = 1 mm (e), 100 μm (magnified images, e).
Figure 4. Ti3C2Tx/antibiotic composite materials [94]. (A) Preparation and characterization of the MXene/SA hydrogel (MX@Gel). (a) Temperature change of MX@Gel containing MXene at various concentrations (0, 10, 50, 100 μg/mL) under near-infrared (NIR) laser irradiation (808 nm, 0.5 W/cm2). (b) Heating and cooling curves of MX@Gel by laser on/off cycle. (c) MX@Gel temperature change graph according to various laser densities (0.50, 1.00, 1.50, 2.00 W/cm2). (d) Schematic of loading ciprofloxacin on MXene surface. (e) UV–vis absorption spectrum of CIP-loaded MXene (CIP-MX). (B) Biocompatibility and cell migration of the CIP-MX@Gel with antibacterial property. (a) Optical microscopy and Live/dead staining images of human dermal fibroblasts cultured on MX@Gel at 1, 4, and 7 days. (b) quantification of cell viability of MX@Gels. (c) Wound healing scratch assay of MX@Gel at different concentrations. (d) Quantification of cell migration (n = 3). (e) Gene expression profiles of MX@Gels (n = 3). (f) Antibacterial gross images of E. coli and S. aureus in CIP-MX@Gel using smear plate method. (g) Quantification of the zone of inhibition of CIP-MX@Gel (n = 3). All data represent mean ± SD. * p < 0.05, ** p < 0.01. The symbol * indicates comparisons with 0 μg/mL. Scale bars = 100 μm (a,c). (C) In vivo wound healing effect of CIP-MX@Gel in mice. (a) Schematic diagram of the formation process of sprayable CIP-MX@Gel, and drug release by NIR irradiation (b) Schematic diagram of the experimental healing process in vivo. (c) Wound images at days 0, 3, 7, 10 and 14 (Silicone isolator diameter = 9 mm). (d) quantification of relative wound area. (n ≥ 3) (e) Hematoxylin and eosin (H&E) staining images on day 7 and 14 of wound tissues treated in different groups. (f,g) Quantification of the panniculus gap at day 7 and day 14 in wound tissue treated with different groups. (h,i) Quantification of the granulation tissue thickness at day 7 and day 14 in wound tissue treated with different groups. All data represent mean ± SD. * p < 0.05, ** p < 0.01. The symbol * indicates comparisons with Sham. Scale bars = 1 mm (e), 100 μm (magnified images, e).
Molecules 29 02902 g004
Figure 5. Ti3C2Tx/antibiofilm composite [103]. (A) The fabrication and application of HPEM scaffolds in multidrug-resistant bacteria-infected wound healing. (B) Antibacterial activity and hemostatic ability assay. (a) Bacteria clones and (b) bacterial viability against E. coli, S. aureus, and MRSA treated with different samples. (c) Total blood loss and photographs from the damaged livers after 60 s of treatment with HPEM scaffolds and the control (* p < 0.05, ** p < 0.01, n = 3). (C) Cytotoxicity and cell proliferation evaluations of scaffolds. (a) Cell viability of HPEM scaffolds and each composite in L929 cells at different concentrations for 24 h. (b) LIVE/DEAD staining images of L929 cells at 1, 3, and 5 days after treated with HPEM scaffolds and controls (10 μg/mL) (live cells: green, dead cells: red, scale bar: 100 μm, n = 3). (c) Fluorescence intensity of L929 cells treatment with HPEM scaffolds and controls at 1, 3, and 5 days (* p < 0.05, n = 5). Cells without any treatment were used as a negative control (NC). (D) In vivo anti-infection and MRSA-infected wound healing. (a) Representative skin wound images at 0, 3, 7, and 14 days and (b) wound area size of the HPEM scaffolds and controls (* p < 0.05, ** p < 0.01). (c) Images of MRSA colonies growing on the agar plates derived from the homogenized infected tissues after various treatments at 3, 7, and 14 days. (d) Quantitative bacterial colonies densities based on (c). (e) IL-6 (red) immunofluorescence images at day 3, and H&E and Masson’s trichrome staining of wound tissues at day 14; the black arrows show the granulation layers in wound beds. Corresponding quantification of (f) fluorescence intensity of IL-6, (g) granulation tissue thickness, and (h) collagen content in wound tissues based on (e) (scale bar = 100 μm, n = 6; * p < 0.05 and ** p < 0.01).
Figure 5. Ti3C2Tx/antibiofilm composite [103]. (A) The fabrication and application of HPEM scaffolds in multidrug-resistant bacteria-infected wound healing. (B) Antibacterial activity and hemostatic ability assay. (a) Bacteria clones and (b) bacterial viability against E. coli, S. aureus, and MRSA treated with different samples. (c) Total blood loss and photographs from the damaged livers after 60 s of treatment with HPEM scaffolds and the control (* p < 0.05, ** p < 0.01, n = 3). (C) Cytotoxicity and cell proliferation evaluations of scaffolds. (a) Cell viability of HPEM scaffolds and each composite in L929 cells at different concentrations for 24 h. (b) LIVE/DEAD staining images of L929 cells at 1, 3, and 5 days after treated with HPEM scaffolds and controls (10 μg/mL) (live cells: green, dead cells: red, scale bar: 100 μm, n = 3). (c) Fluorescence intensity of L929 cells treatment with HPEM scaffolds and controls at 1, 3, and 5 days (* p < 0.05, n = 5). Cells without any treatment were used as a negative control (NC). (D) In vivo anti-infection and MRSA-infected wound healing. (a) Representative skin wound images at 0, 3, 7, and 14 days and (b) wound area size of the HPEM scaffolds and controls (* p < 0.05, ** p < 0.01). (c) Images of MRSA colonies growing on the agar plates derived from the homogenized infected tissues after various treatments at 3, 7, and 14 days. (d) Quantitative bacterial colonies densities based on (c). (e) IL-6 (red) immunofluorescence images at day 3, and H&E and Masson’s trichrome staining of wound tissues at day 14; the black arrows show the granulation layers in wound beds. Corresponding quantification of (f) fluorescence intensity of IL-6, (g) granulation tissue thickness, and (h) collagen content in wound tissues based on (e) (scale bar = 100 μm, n = 6; * p < 0.05 and ** p < 0.01).
Molecules 29 02902 g005
Figure 6. Ti3C2Tx/photosensitizer composite material [110]. (A) A sSchematic illustration of the preparation of Ru@MXene nanocomposites. (B) A sSchematic diagram of Ru@MXene as a photoconverting agent to produce heat and 1O2 for achieving PTT/PDT antibacterial activity. (C) Photothermal property assessments: photothermal heating curve of Ru@MXene (a) with gradient concentrations; and (b) with different power density. (c) Photothermal heating curve of various samples with a power density of 150 mW/cm2. (d) Real-time infrared thermal images of different samples. (e) Photothermal periodic curve of Ru@MXene. (f) Schematic diagram of simulated sunlight irradiation samples. (D) Antibacterial mechanism evaluations involving intracellular ROS level detection.: The fFluorescence intensity of intracellular ROS induced by different samples with or without xenon lamp illumination (150 mW/cm2, 30 min) in (a) E. coli and (b) S. aureus. (Scale bar: 50 μm). (E) Antibacterial mechanism evaluations involving GSH consumption: (a) Photographs of the color change in GSH solution after incubating with various materials with or without xenon lamp illumination (150 mW/cm2, 30 min). (b) Corresponding GSH consumption rate (n = 3). (c) Chemical reaction between DTNB and GSH.
Figure 6. Ti3C2Tx/photosensitizer composite material [110]. (A) A sSchematic illustration of the preparation of Ru@MXene nanocomposites. (B) A sSchematic diagram of Ru@MXene as a photoconverting agent to produce heat and 1O2 for achieving PTT/PDT antibacterial activity. (C) Photothermal property assessments: photothermal heating curve of Ru@MXene (a) with gradient concentrations; and (b) with different power density. (c) Photothermal heating curve of various samples with a power density of 150 mW/cm2. (d) Real-time infrared thermal images of different samples. (e) Photothermal periodic curve of Ru@MXene. (f) Schematic diagram of simulated sunlight irradiation samples. (D) Antibacterial mechanism evaluations involving intracellular ROS level detection.: The fFluorescence intensity of intracellular ROS induced by different samples with or without xenon lamp illumination (150 mW/cm2, 30 min) in (a) E. coli and (b) S. aureus. (Scale bar: 50 μm). (E) Antibacterial mechanism evaluations involving GSH consumption: (a) Photographs of the color change in GSH solution after incubating with various materials with or without xenon lamp illumination (150 mW/cm2, 30 min). (b) Corresponding GSH consumption rate (n = 3). (c) Chemical reaction between DTNB and GSH.
Molecules 29 02902 g006
Figure 8. Application of Ti3C2Tx composite antibacterial materials [46,152,153,154].
Figure 8. Application of Ti3C2Tx composite antibacterial materials [46,152,153,154].
Molecules 29 02902 g008
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chen, H.; Wang, Y.; Chen, X.; Wang, Z.; Wu, Y.; Dai, Q.; Zhao, W.; Wei, T.; Yang, Q.; Huang, B.; et al. Research Progress on Ti3C2Tx-Based Composite Materials in Antibacterial Field. Molecules 2024, 29, 2902. https://doi.org/10.3390/molecules29122902

AMA Style

Chen H, Wang Y, Chen X, Wang Z, Wu Y, Dai Q, Zhao W, Wei T, Yang Q, Huang B, et al. Research Progress on Ti3C2Tx-Based Composite Materials in Antibacterial Field. Molecules. 2024; 29(12):2902. https://doi.org/10.3390/molecules29122902

Chicago/Turabian Style

Chen, Huangqin, Yilun Wang, Xuguang Chen, Zihan Wang, Yue Wu, Qiongqiao Dai, Wenjing Zhao, Tian Wei, Qingyuan Yang, Bin Huang, and et al. 2024. "Research Progress on Ti3C2Tx-Based Composite Materials in Antibacterial Field" Molecules 29, no. 12: 2902. https://doi.org/10.3390/molecules29122902

Article Metrics

Back to TopTop