Next Article in Journal
Utilising Portable Laser-Induced Breakdown Spectroscopy for Quantitative Inorganic Water Testing
Previous Article in Journal
High-Performance MEMS Oxygen Sensors Based on Au/TiO2 Films
Previous Article in Special Issue
Heater Topology Influence on the Functional Characteristics of Thin-Film Gas Sensors Made by MEMS-Silicon Technology
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Design of Functional Ti3C2Tx MXene for Gas Sensors and Energy Harvesting: A Review

1
Department of Physics, Gachon University, 1342 Seongnamdaero, Sujeong-gu, Seongnam-si 13120, Gyeonggi-do, Republic of Korea
2
Department of Agriculture, Forestry and Bioresources, Seoul National University, 1 Gwanak-ro, Gwanak-gu, Seoul 08826, Republic of Korea
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Chemosensors 2023, 11(9), 477; https://doi.org/10.3390/chemosensors11090477
Submission received: 26 June 2023 / Revised: 29 July 2023 / Accepted: 13 August 2023 / Published: 1 September 2023
(This article belongs to the Special Issue Gas Sensors and Electronic Noses for the Real Condition Sensing)

Abstract

:
Two-dimensional (2D) inorganic compounds, MXenes, are the most promising candidate for chemical sensors and environmental remediation. Since the first synthesis of Ti3C2Tx MXene from the Ti3AlC2 MAX phase in 2011, 2D materials have been attracting significant attention from a wide range of scientific communities because of their unique physicochemical properties. The attractive properties of MXenes motivated us to explore the new wave of front-end research and applications. Over the past 12 years, there have been more than 10,000 theoretical and experimental studies on MXenes. All these publications have primarily focused on Ti3C2Tx MXene because of its fascinating material properties and tunability towards target applications. To provide readers with a fundamental understanding of this emerging 2D material, this review discusses the recent trends in the design of Ti3C2Tx MXene for gas sensors and energy harvesting applications. For the most updated information, this review focuses on important findings and applications reported in the last decade.

1. Introduction

Ever-growing power consumption, the accumulative demand for smart materials, and the need for green electronics for low-power and low-voltage devices have led to increased demand to develop alternatives for supercapacitors and batteries, which is exemplified by self-powered wearable microelectronic devices [1,2,3]. Therefore, it is crucial to find smart materials that can be applied to photocatalysis, electrocatalytic sensors, chemical adsorbents, energy conversion, gas sensors, and biomedical applications. To this end, by the end of 2004, Novoselov et al. [4] actively explored graphene, which is considered the pioneering work of two-dimensional (2D) nanomaterials. Subsequently, an increasing number of 2D nanomaterials such as phosphorene, transition metal dichalcogenides, transition metal carbides/nitrides, and hexagonal boron nitrides have been discovered [5,6,7]. Among these 2D materials, new 2D materials referred to as MXenes have been at the forefront and become suitable candidates for electrodes, detectors, additives, and separators or hosts for numerous applications because of their outstanding properties [8,9].
Since the first publication on 2D Ti3C2Tx from Drexel University, an increasing number of researchers have focused on producing Ti3C2Tx MXene, demonstrating its attractive qualities, and exploring its diverse applications [10]. The formula for MXenes is Mn+1XnTx, where M, X, and Tx represent an early transition metal (Ta, Hf, Mo, Ti, Mo, Nb, and V), carbon and/or nitrogen, and surface terminations (-OH, -F, =O), respectively; n = 1–4 [11,12,13]. MXenes have tremendous compositional variety and tunable qualities depending on their structural atomic arrangement, because transition-metal atoms are grouped in a layered pattern with carbon or nitrogen atoms [14,15,16]. Furthermore, in such atomic sandwiches, some of the transition-metal atoms are arranged in the surface layers while the remaining metal atoms are placed in the inner layers. Different transition metals can merge into MXenes, such as Mo2TiC2 or Mo2Ti2C3 [17,18].
As members of the “wonderful material family,” MXenes have been rapidly expanding because of the growing diversity of their parent materials, MAX phases, and advancements in their processing methods [19,20]. The ability to manipulate the structure and morphology of MXenes using multiple methodologies offers the possibility to create a variety of 2D materials with customizable compositions and favorable surface qualities [21,22,23].
Thus far, more than 35 stoichiometric MXenes have been explored as solid solutions with different combinations of M and X sites recorded [24,25]. From theoretical calculations, a considerably larger number of solid solutions have been obtained, and their experimental syntheses remain pending. A wide range of surface terminations, including Cl, F, O, OH, I, and Br, further expand the number of stoichiometric MXenes, endowing them with unique physical and chemical properties [26,27,28]. Thus, MXenes have been used in a variety of potential applications, including electromagnetic interference shielding, optoelectronics, photocatalysis, gas sensors, and energy storage [29,30,31]. Although metallic conductivity facilitates Ti3C2Tx MXene-based chemiresistive gas sensors running at room temperature, gas sensors prepared with pure Ti3C2Tx MXene suffer from poor selectivity (only NH3 or VOCs) and low response during gas detection [22,32]. Therefore, integrating Ti3C2Tx MXene with other materials is a promising strategy to enhance catalytic activity, facilitate efficient charge transfer, and reduce energy barriers [33,34,35,36]. These effects cooperatively contribute toward enhancing the selectivity of the gas-sensing system and enable operation at room temperature.
Among all MXenes, Ti3C2Tx MXenes offer an exceptionally appealing platform for a wide range of applications, especially for gas sensing and energy harvesting, because of their remarkable electrical qualities and unique tunability. The two applications are important and can stimulate the design of multifunctional Ti3C2Tx MXene-based flexible, high-performance, and cost-effective smart materials. Although there have been several comprehensive reviews on the synthesis, properties, and applications of Ti3C2Tx MXene [37,38,39,40,41,42,43], reviews on the detailed and prospective preparation and applications of multifunctional Ti3C2Tx MXene for gas sensors and energy harvesting are very rare. The need for batteries and electrical power sources can be dramatically overcome by combining energy harvesting technologies with gas sensor systems, which also makes the gas sensors more sustainable and easier to deploy in both remote and hard-to-reach locations. In addition, energy harvesting can reduce the maintenance costs of the gas sensors and enhance the overall lifetime of the devices [44,45,46].
To this end, we highlight recent advances in Ti3C2Tx MXene implementation for gas sensing and energy harvesting in various environments. Using a self-assembly method, rGO-assisted method, and framework-supported method to prepare Ti3C2Tx MXene can efficiently prevent it from stacking together, which promotes the diffusion of toxic gas molecules. Finally, we discuss future research difficulties and techniques to further investigate the unique properties and potential applications of Ti3C2Tx MXenes.

2. Synthesis of Ti3C2Tx MXene

2.1. Etching

The functional group of Ti3C2Tx MXene critically affects its electrical conduction properties, serving as covalent linking sites with other nanostructures and as active sites for adsorbing gas molecules. Ti3C2Tx MXenes are created by selectively etching specific atomic layers from their parent MAX phases, where A is an A group (groups 13 and 14, or IIIA, IVA) of the periodic table [47]. Selective etching is the main experimental method for synthesizing Ti3C2Tx MXene, which has a metallic Ti-Al bond strength weaker than that of Ti-C bonds [48]. Aqueous solutions with the fluoride group have been mostly used to remove Al layers from the Ti3AlC2 MAX phase, such as direct hydrofluoric acid (HF, 50%) or the in situ creation of HF from lithium fluoride (LiF) and hydrochloric acid (HCl), as shown in Figure 1 [49,50,51,52]. Both ammonium fluoride (NH4F) and ammonium hydrogen bifluoride (NH4HF2) have also been effectively applied for making Ti3C2Tx MXene. Only Al has been effectively removed from MAX phases from the 12 elements of groups 13 and 14. The chemical reactions that happen during the HF etching of the Ti3AlC2 MAX phase are given below [53,54,55]:
Ti3AlC2 + 3HF → Ti3C2 + 1.5H2 + AlF3
Ti3AlC2 + 2H2O → Ti3C2(OH/O)2 + H2
Ti3AlC2 + 2HF → Ti3C2F2 + H2
There is another technique to prepare MXenes from nonMAX phases, which is etching aluminum carbide (Al3C3) from a nonMAX phase. In this nonMAX phase precursor, Al-C units separate either M3C2 or M2C layers, e.g., U2Al3C3. Mo2CTx MXenes have also been prepared via Ga layer etching from Mo2Ga2C [10,56,57,58]. MXenes can be synthesized using molten fluoride salt at high temperatures, as demonstrated by the synthesis of Ti4N3 from Ti4AlN3 at 550 °C under Ar flow. The TiCx nanostructure was synthesized by annealing a mixture of molten LiF and Ti2AlC at 900 °C [59]. This method was not efficient and removed either the M or A element because of the specific treatment environment. These results and the morphologies of Ti3C2Tx MXene suggest that the aqueous fluoride-containing acidic technique is suitable for synthesizing Ti3C2Tx MXenes. Moreover, etching in the presence of a metal halide such as LiF tends to intercalate the cation (Li+) and water, increasing the distance between each layer [60,61,62,63].
Halogen etching is a new technique for preparing Ti3C2Tx MXene. In 2021, Jawaid et al. [64] explored an etching method to prepare Ti3C2Tx MXene using halogens (IBr, ICl, I2, Br2) in anhydrous media. The halogen-based etching is conducted in a glove box for 8 h with an inert ambient (H2O:O2 was around 1:3) at room temperature, as shown in Figure 2. The recommended mechanism implies that continual halogen addition will provide reaction productivity and increase the yield of halogenated Ti3C2Xn with halogen-terminated surfaces (~1% yield of MXene on 1 mg/mL). The etching process with different halogens led to a variety of halogen-terminated surfaces (X = Br, I, C, F) at room temperature. A low-dielectric organic solution was used to obtain the purified size-selected Ti3C2Xn MXene. In the same context, a fluoride-free Ti3C2Tx MXene with O-rich terminal groups was generated using an iodine-assisted etching method in anhydrous acetonitrile (CNCH3) [65]. The Al layer was removed from the MAX phase by iodine because the Ti-C bond is stronger than Ti-Al. The I2-etched MXene was washed several times using a 1 M HCl solution to dissolve AlI3 and exfoliate multilayer Ti3C2Tx MXene by manual shaking. The I2-etched Ti3C2Tx MXene has excellent conductivity (1250 S/cm), ultimate thickness (5 nm), and large size (1.8 µm) with oxygen-rich termination groups. These abundant oxygen functional groups are good for gas-sensing performance and energy storage because they act as active sites and take up hydrogen ions.
Yang et al. [66] prepared Ti3C2Tx using NH4Cl and tetramethylammonium hydroxide (TMA.OH) using electrochemical etching with low potential (+5 V). The exfoliated single-layer and bilayer Ti3C2Tx MXenes were collected with high yield (90%) because of the delamination of the positive charge ions (TMA+) and as Cl ions have a strong binding with Al. This technique does not require any dangerous F-containing agents, and the Al layers are replaced by hydroxide and the oxygen group. The average lateral size of the obtained Ti3C2Tx MXene is ~2 µm, which is larger than that of the traditional HF etching process. This method offers a safe technique to scale the preparation of Ti3C2Tx MXene. In 2018, Tengfei et al. [67] attempted to directly utilize 27.5 M NaOH to etch 100 mg of the Ti3AlC2 MAX phase in an Ar atmosphere with excellent efficiency (92 wt%). The obtained samples were synthesized at 270 °C via alkali treatment and had a (002) plane spacing of 23.4 Å, which was larger than that of Ti3C2Tx prepared using an HF-related technique. This technique offers an alkali etching process for Ti3C2Tx MXene in which the pure MAX phase must be uninvolved. Moreover, Ti3C2Tx MXene can be cationized with poly(diallyldimethylammonium chloride) (PDDA) as a stabilizer to better compound with negatively charged materials through self-reduction and/or electrostatic interaction [68].

2.2. Delamination

Delaminating any 2D material is important for achieving interlayer interaction. Ti3C2Tx MXenes can be intercalated and exfoliated in polar organic molecules, followed by manual shaking or mechanical vibration in water to obtain monolayer and few-layer Ti3C2Tx MXenes. Ultrasonication is the main tool for separating and delaminating Ti3C2Tx MXene sheets with precision to dramatically accelerate the procedure and process an advanced fraction of high-quality nanosheets. Ti3C2Tx can be exfoliated with dimethyl sulfoxide (DMSO) or tetrabutylammonium hydroxide (TBAOH) by manual shaking, and it can then be filtered to obtain the freestanding MXene “paper” [16,27,69]. Etching the Ti3AlC2 MAX phase in the presence of a fluoride salt makes it easy to obtain few-layer Ti3C2Tx MXene without any delamination step because of the effect of metal cations in halide salts (Figure 3) [70,71,72].
The resistance of Ti3C2Tx MXene without intercalation is smaller than that with intercalation because intercalation-induced interlayer swelling retards the out-of-plane electron transport. Metal ion intercalation is one of the effective techniques for improving the diffusion coefficients of gas molecules and ionic species [16,73]. Consequently, scientists focused on using metal ion intercalation to enhance the gas-sensing performance and the energy storage activity of Ti3C2Tx MXene-based composites. The enhanced gas-sensing mechanism of metallic Ti3C2Tx MXene is more sophisticated than the electron movement and surface adsorption in traditional 2D materials. Suitable cation exfoliation can contribute to the swelling of Ti3C2Tx MXene, which can improve the conductivity change when exposed to reducing gases or oxidizing gases.

3. Structure and Properties

Ti3AlC2 is the most commercially available, low-cost MAX phase for producing Ti3C2Tx MXene with superior metallic conductivity and low diffusion barriers. In terms of electrical properties, Ti3C2Tx MXenes were almost like multilayer graphene. The conventional HF-etched Ti3C2Tx MXene and additive-free Ti3C2Tx MXene film have shown volumetric capacitances around 300 F and 1500 F per cubic centimeter, respectively [74,75]. In addition, the multilayered structure of Ti3C2Tx MXene provides a high cycling rate and stability. Benefitting from the electronic sensitization and synergistic effect of Ti on MXenes, the charge carrier transportation, response, and other electrical parameters of Ti3C2Tx MXene-based electrical devices can be enhanced significantly.
Similar to its parent, Ti3C2Tx MXene has a hexagonal close-packed structure, and the C atoms occupy the octahedral interstitial sites composed of 6 Ti atoms. Ti3C2Tx MXenes are prepared using hydrofluoric acid and have surface functional groups Tx (-OH, -F, =O). Moreover, they can have another kind of Tx, such as -Cl, -Br, and -I, when different etching and delaminating agents are employed [76,77,78].
The etchant plays a crucial role in the delamination quality of MXene. For example, a HF-etched Ti3C2Tx has thicker multilayer lamellas and narrower interstack gaps than those of (LiF + HCl)-etched Ti3C2Tx. The space between each layer of (LiF + HCl)-etched Ti3C2Tx is around 2.8 Å due to the presence of metal cation, which is larger than that of HF-etched Ti3C2Tx MXene [79]. The product from a lower HF concentration yields a larger O:F ratio [80,81,82,83]. For the removal of the Al layer in the Ti3AlC2 MAX phase, at least 18 h of HF etching (30 wt%) is necessary, which was established by the shift of a main peak ((002) peak) from 9.5° to 9.0° in the XRD measurement, as shown in Figure 4. In fact, Ti3C2Tx MXenes with less open structures were collected when etched with HF solutions of 5 wt% and 10 wt% for 24 h and 18 h [5].
Multilayer Ti3C2Tx MXenes are not stable in an ambient environment with water and oxygen molecules; however, they are stable in dry air and oxygen-free degassed water. Consequently, it is highly recommended to store Ti3C2Tx MXene colloids at low temperatures in refrigerators and in dark environments. Further, the standard manufacturing procedure yields high-quality Ti3C2Tx MXene with high stability [84,85,86,87]. However, Ti3C2Tx MXene powders can be stored in a vacuum environment; vacuum drying or lyophilization can be active in preserving MXene powders, because they eliminate the solvent and reduce the potential for deprivation or agglomeration. In terms of suitability during experiments, using Ti3C2Tx MXene powder can have certain advantages over colloidal suspensions.

4. Applications of Ti3C2Tx MXenes in Gas Sensors

The main function of gas sensors is to detect gases at low concentrations (ppm, ppb, or even ppt level), which are lower than the recognition of the nose olfactory nerve. In industry, gas-sensing devices are utilized to control the concentration of toxic gases such as volatile organic compounds (VOCs), NO2, and H2S to guarantee that the working atmosphere is safe for workers. Moreover, it can also be used to control the concentration of explosive and flammable gases. The gas-sensing material plays the main role in the operation of gas sensors, which should be normally stable; however, they should sensitively respond to target gases at room temperature. The gas sensor response is determined by calculating the relative change in the sensor material’s resistance in an air environment (Rair) versus its resistance in the specific gas being tested (Rgas) after the injection of gas. The calculation is
R e s p o n s e % = R g a s R a i r R a i r × 100

4.1. MXene Gas Sensors

Multilayered Ti3C2Tx MXene has been considered a potential candidate for gas sensors because of its large surface area, high electrical conductivity, and wealthy surface functional groups that can act as active sites. Ti3C2Tx MXene provides an excellent path for electron transport at room temperature because of its metallic conductivity, which results in enhanced gas-sensing signals [87,88,89,90,91,92]. Table 1 shows that pure Ti3C2Tx MXene and Ti3C2Tx composites can interact with various toxic gas molecules with a good response and fast recovery time. The resistance of metallic Ti3C2Tx MXene was almost increased when exposed to a toxic gas in both oxidizing and reduced gas environments at room temperature [6,38,93,94]. One study by Lipatov et al. [94] examined the effect of different gas environments on the electrical conductivity of Ti3C2Tx MXene films. The authors discovered that exposure to ambient air, which can be classified as an oxidizing environment, led to an increase in the resistance of the MXene films because of the formation of a thin surface oxide layer. Unlike the rise in resistance in oxidizing environments, several studies have shown that Ti3C2Tx MXenes exhibit enhanced electrical conductivity in reducing gas environments [95,96,97]. For example, Doan et al. [95] revealed that the electrical conductivity of Ti3C2Tx MXene films massively increased upon exposure to hydrogen gas, which acts as a reducing agent. A convincing gas-sensing mechanism of metallic Ti3C2Tx MXene needs to be developed to fully understand the sensing process and create a breakthrough for further applications of Ti3C2Tx MXenes.
Among the members of the MXene family of 2D materials, a considerable amount of research has been performed on Ti3C2Tx MXene to establish its synthesis routine and unveil its qualities [116,117]. Lee et al. [118] developed a gas sensor using Ti3C2Tx MXene; the sensor displayed good sensing performance to ammonia, ethanol, methanol, and acetone at room temperature. For their study, Ti3C2Tx MXene was prepared using 9 M HCl and 1 g of LiF at 35 °C, and a flexible polyimide film was used as the gas-sensing substrate (Figure 5). The average response signals of the as-prepared material were 7.5, 11.5, 14.3, and 21% for 100 ppm of acetone, ethanol, methanol, and ammonia, respectively. All response signals of Ti3C2Tx had positive signs for all types of gases, suggesting that MXene behaved like a p-type material. Thus, the p-type behavior resulted from oxygen and water molecules introduced during the etching process, which may have acted as the main active sites and received electrons from the electron-donating gases, leading to a drop in the majority carrier population. Thanh et al. [95] constructed palladium-decorated Ti3C2Tx MXene using a polyol method to combine 1D Pd nanoparticles with multilayered Ti3C2Tx MXene for multifunctional operation as hydrogen storage and a H2 gas sensor. The responses of the Pd-Ti3C2Tx composite to 10, 50, and 100 ppm of H2 were recorded at 10, 20, and 25%, respectively. The enhanced responses of optimized Pd-Ti3C2Tx composite could be ascribed to the roles of Pd NPS and the aforementioned O functional groups of MXene, which may collaboratively supply H atoms to a nearby composite via a spill-over mechanism.
Kim et al. [32] reported that the Ti3C2Tx MXene gas sensor exhibits high sensitivity to volatile organic compounds with a very low detection limit (50 ppb) at room temperature. The sensor’s signal-to-noise ratio was extremely high, twice as large as that for other sensing materials. The response values of Ti3C2Tx films on exposure to 100 ppm of SO2, NO2, propanol, ammonia, ethanol, and acetone were 0.16, 0.21, 0.88, 0.80, 1.76, and 1.18%, respectively. The low noise of the as-prepared Ti3C2Tx MXene can be attributed to the high electrical conductivity and structural uniformity obtained by the vacuum-filtrated technique, which has a comparative advantage over other techniques such as drop-casting. Ti3C2Tx MXene with a small thickness (25 nm) showed good sensitivity compared to MXenes with 50 and 150 nm thicknesses; this is strongly correlated with the fraction of active sites exposed on the surface (Figure 6). They demonstrated that the highly conductive Ti3C2Tx MXene could be used to sense different VOCs with clear response signals and low electrical noise, realized by the support of the hydroxyl and oxygen groups. According to the DFT calculations, Ti3C2(OH)2 showed the toughest binding energy strengths (1.2 Å and −0.774 eV for acetone), stronger than those of other functional groups.
Sanjit et al. [119] prepared accordion-like-Ti3C2Tx MXene from Ti3AlC2 via 48% HF etching for 15 h at 60 °C. The as-prepared Ti3C2Tx MXene displayed a macroscopic layer separation of 200 nm. The resistance changed positively when the Ti3C2Tx sensor was exposed to acetone at room temperature. The responses to 100 ppm and 0.25 ppm of acetone vapor were 99%, and 17.3% with a fast response time (53 s), respectively. The hydrogen bonding on the surface of MXene facilitated acetone-molecule interactions with the sensor material. The O and C atoms in acetone gas strongly bound to the Ti3C2Tx MXene surface. Multilayered Ti3C2Tx MXene seems to promote a higher surface area and possibly a higher gas adsorption capacity. Low-layered Ti3C2Tx MXene offers faster response times, higher sensitivity, and better accessibility to gas molecules, but with a lower surface area and potential stability. The choice between these Ti3C2Tx MXene alternatives depends on the specific requirements of gas-sensing applications.
Winston et al. [120] studied functionalizing multilayered Ti3C2Tx with a superhydrophobic protection layer using fluoroalkylsilane functionalization. The optimized sample provided a superhydrophobic surface and enabled the sensor to operate normally even in humid environments. The sensor yielded responses of 3, 5, and 14% to 5, 15, and 120 ppm of ethanol, respectively. The improved responses could be attributed to local structure deformation and adjacent oxygen atoms.
In 2020, Shoumya et al. [121] doped Ti3C2Tx MXene with sulfur to design a room temperature VOCs sensor. In their report, Ti3C2Tx MXene was mixed with thiourea as a sulfur source, and then heated to 500 °C under Ar flow. The optimal S-doped Ti3C2Tx MXene showed an excellent response (12%) to 1 ppm toluene with long-term stability (1 month). The -S functional groups play a key role for enhanced gas sensing, which is modified upon doping, thereby increasing the binding energy of gas molecules substantially.
Wu et al. [22] used dimethyl sulfoxide to synthesize single-layered Ti3C2Tx MXene for toxic gas detection at room temperature. Among the eight different gases, the response signal for NH3 was 6.1%, which is four times the response for ethanol (1.5%). The selectivity of the NH3 chemiresistive gas sensor was higher than others because of the clean surface of the microstructure of Ti3C2Tx MXene, which had a higher adsorption area when using HCl and NaF as the etchant. The potential of Ti3C2Tx MXene for NH3 detection was demonstrated using first-principles calculation. The authors indicated that O-terminated Ti3C2Tx was the most energetically favorable compared to the F-terminated Ti3C2Tx and OH-terminated Ti3C2Tx. The calculations displayed that the adsorption energy of NH3 (−0.078 eV/atom) has a considerably stronger interaction with O-terminated Ti3C2Tx than that of CH4 (−0.022 eV/atom), H2S (−0.047 eV/atom). Further, the N atom in the NH3 molecules lies straight above the Ti atom in O-terminated Ti3C2Tx (Figure 7). The adsorbed geometry and adsorption energy results guarantee that the response of NH3 molecules is considerably better than those of other toxic gases; the O-terminated Ti3C2Tx MXene has a high selectivity to ammonia at room temperature.
Machine learning was applied to predict the appearance of VOCs molecules on the Ti3C2Tx MXene-based multivariable sensor [122]. Eight different gases with concentrations from 100 to 800 ppm were used as the original data set (Toluene (TOL), methanol (MeOH), hexane, ethanol (EtOH), dicholoromethane (DCM), acetone, acetonitrile (MeCN), and isopropanol (IPA)). The principal component analysis (PCA) and linear discrimination analysis (LDA) were executed as supervised pattern recognition tools. Therefore, the VOC mixtures were well separated and had high data dimensionality. The correct cross-validation value of 90.9% was reached, and it revealed that utilizing the Ti3C2Tx MXene-based multivariable gas sensor is a potential method for detecting pure VOCs.
Very few studies have focused on the gas-sensing performance of other kinds of MXenes than Ti3C2Tx MXene. For example, Lee et al. [123] demonstrated the great performance of metallic V2CTx MXene gas sensors towards methane and hydrogen with ultrahigh sensitivity at room temperature. Methane has high enthalpy (C-H bonds) and nonpolarity, therefore, it is tremendously difficult to sense the gas at room temperature. Multilayered V2CTx MXenes were prepared by Al etching from the V2AlC MAX phase with HF acid. Monolayer V2CTx MXene flakes were collected via tetra n-butyl ammonium cation delamination. The gas responses of monolayer V2CTx MXene flakes are estimated to be 1.67% and 0.5% for methane and hydrogen sulfide, respectively (Figure 8). The outstanding detection of V2CTx MXene is related to the oxygen-containing termination groups.
A toluene sensor using 2D MXene Mo2CTx was studied by Wenzhe et al. [124]. The as-obtained sample successfully detected VOCs with an LOD of 220 ppb at room temperature. The response to acetone, methanol, toluene, ethanol, and benzene was 0.14, 0.58, 2.81, 0.73, and 0.97%, respectively, indicating that Mo2CTx MXene was selective toward toluene compared to other VOCs. The detecting mechanism of the sensor material was based on the interaction between Mo2CTx flakes and the benzene ring in VOCs molecules, which reduces the number of charge carriers. The methyl group further enhance the activity of the phenyl ring in toluene molecules, supporting the evidence that toluene displays the highest response compared to other gases.

4.2. MXene Composites with Chalcogenides and Conductive Polymers

Le et al. [125] researched the effect of transition metal dichalcogenides on Ti3C2Tx MXene for gas-sensing activity. WO3/Ti3C2Tx and MoS2/Ti3C2Tx composites were prepared by conducting a hydrothermal reaction for 8 h at 170 °C. As shown in Figure 9, the MoS2-covered Ti3C2Tx MXene exhibited great sensitivity to 10 ppm of NO2 gas (response = 35%), which was two-fold higher than that of the WO3/Ti3C2Tx composite (16%) at the same conditions. The enhanced NO2-sensing capability of MoS2/Ti3C2Tx composites can be explained based on the excellent interaction between MoS2 and Ti3C2Tx MXene. Many S-terminated edges with high d electron density can facilitate active electronic interactions with NO2 molecules. The electron donation movement can be addressed as
NO2 + O2 + 2e → NO2 + 2O
The conductive polymer is suitable for improved gas sensing because of its high sensitivity, high flexibility, and facile fabrication. Therefore, combining Ti3C2Tx MXenes with conductive polymers to create nanocomposites is a good strategy in the field of wearable devices [107,126]. Further, heterojunctions and Schottky junctions inside composites are highly beneficial for gas sensor activities. Jun et al. synthesized the heterostructures of Ti3C2Tx MXene and poly (3,4 ethylenedioxythiophane) polystyrene sulfonate (PEDOT:PSS) using an in situ polymerization technique. The optimum heterostructure with 15% Ti3C2Tx MXene exhibited the best response of 36.7% to 100 ppm of NH3 gas with fast recovery (40 s) and response (117 s) times. From a comparison with previously researched NH3 sensors, they concluded that the direct charge transfers and chemisorption of oxygen play a crucial role in improving the sensing performance. In addition, the sensor displays a similar response signal regardless of the bending angles (240°, 180°, 120°, 60°), thereby representing its tremendous mechanical stability and potential to be utilized in wearable sensor devices, as indicated in Figure 10 [109].
Wang et al. [9] prepared nanocomposites of PEDOT:PSS/Ti3C2Tx MXene for a methanol sensor at room temperature. A gas sensor based on the composite with a mass ratio of 4:1 (Ti3C2Tx MXene 20 wt%) revealed the highest response (5.5%) to 300 ppm methanol compared to pure PEDOT:PSS (4.67%) and pristine Ti3C2Tx MXene (1.13%) [9]. The presence of Ti3C2Tx in the polymer matrix enlarges the interchain space of PEDOT:PSS, making electron hopping difficult. The high metallic behavior enables the composite sensors to run under 0.1 V with low power consumption. The formation of heterojunctions at the interfaces of the conductive polymers and Ti3C2Tx MXene significantly improves the mechanical flexibility and practicality of wearable gas sensors.

4.3. MXene-Semiconductive Metal Oxide Composites

Metal oxide semiconductors (MOS) are the most used gas-sensing materials for a variety of applications because of their facile synthesis, high sensitivity, and easy functionalization. However, the high working temperature and low selectivity were the main disadvantages of these MOS-based gas sensors. To solve this dilemma, a combination of metallic Ti3C2Tx MXene and semiconducting MOS can overcome the low selectivity and high operating temperature issues. The formation of the Schottky junction at the interfaces of the metal oxide semiconductor and Ti3C2Tx MXene can accelerate the movement of only one type of charge carriers across the interface.
There are several extensive reports on the gas-sensing activities of MOS/MXene heterostructures, where MXenes are a conductive layer and MOS is the main sensing material [127,128,129,130]. Ta et al. [131] constructed a ZnO/Ti3C2Tx MXene heterostructure, which consists of 1D ZnO nanoparticles and 2D few-layered Ti3C2Tx MXene, using a simple technique. The optimized heterostructure exhibited responses of 54% and 25% to 10 and 5 ppm of NO2, respectively. Partially oxidized TiO2/Ti3C2Tx composites were synthesized by oxygen plasma treatment at 350 °C, substantially increasing the responses to acetone (180% at 2000 ppb) and ethanol (22.5% at 100 ppm) [132,133]. Ti3C2Tx MXene is partially converted to TiO2 nanoparticles, and the oxygen termination groups play a pivotal role in the enhancement of gas-sensing performance compared to the pure MXene, as shown in Figure 11.
Wang et al. [8] synthesized a SnO-SnO2/Ti3C2Tx composite for an acetone sensor at room temperature using the hydrothermal method. The creation of p-n junctions and Schottky junctions in the SnO-SnO2/Ti3C2Tx composite was key to improving the acetone-sensing responses (Rg/Ra = 12.1, 100 ppm), which were 11 and 4 times higher than those of pure Ti3C2Tx and SnO-SnO2, respectively. After the connection, the electron moved from the n-type SnO2 and metallic Ti3C2Tx to the p-type SnO because of the differences in Fermi energy; the electron depletion layer (EDL) decreases when acetone enters. Therefore, the resistance of the sensing material increased, and the oxidation-reaction mechanism of the entire process can be explained as
O2 + e → O2 (ads)
CH3COCH3 + 4O2 → 3CO2 + 3H2O + 4e
Bimetallic oxide is also available for gas sensing because it has several oxygen vacancies. Zou et al. [115] prepared a Fe2(MoO4)3/Ti3C2Tx MXene heterostructure using a hydrothermal method and sensed 100 ppm of n-butanol at 120 °C (response = 43.1%). The nanocomposite showed a large specific surface area (47 m2/g) with a small pore size (10 nm), which was better than that of pure Fe2(MoO4)3 (11 m2/g, 20.8 nm). The composite with 1 wt% of Ti3C2Tx MXene showed a small pore size caused by nanoparticles blocking the hole. Rich active sites and abundant chemical oxygen species were the main factors for enhancing the gas-sensing performance with a short response time (18 s).

4.4. Self-Powered Gas Sensors

Gas sensors require external power sources, which restricts their application in wearable devices. Several researchers focused on self-powered gas sensors to tackle this problem. Zhang et al. [104] studied a Co3O4/Ti3C2Tx composite-based gas sensor, which was driven by the ZnO/Ti3C2Tx nanostructure array piezoelectric nanogenerator (PENG) using the titanium and aluminum foils as electrodes. ZnO/Ti3C2Tx nanowires on the Ti foil acted as the power source, converting the mechanical energy of human movement into electrical energy. The power generated from PENG could run the gas sensor device, which exhibited high sensitivity (Rg/Ra = 9.2) to 10 ppm of HCHO with a fast recovery speed (recovery time = 5 s) [104].
Wang et al. [134] studied CuO/Ti3C2Tx composites for a self-powered NH3 sensor at room temperature. The sensor was driven by a triboelectric nanogenerator (TENG) using latex and polytetrafluoroethylene (PTFE). The resistance of the CuO/Ti3C2Tx composite increased when exposed to NH3 gas (0–100 ppm) which resulted in a voltage–resistance (U–R) curve through the load, as shown in Figure 12. The output voltage versus gas concentration curves can be obtained via sensing signals, and then can be converted to resistance versus gas concentration curves using the corresponding U–R relationship. The CuO/Ti3C2Tx composite showed a good response (25%) to 100 ppm of NH3 and could detect the deterioration of pork because it releases NH3 as it rots. Here, CuO showed p-type semiconductor behavior with holes as the majority carriers; the hole concentration increased when it was in contact with oxygen molecules. Many electrons jump into the conduction band of CuO under exposure to an NH3 atmosphere, and this increases the resistance of CuO. In the same context, the resistance of organ-like MXene increased in an NH3 environment. The excellent interaction between negatively charged MXene and positively charged CuO led to an effective combination for improved NH3 sensing. A wearable TENG device had a short-circuit current of 34 µA (open-circuit voltage = 810 V) and could light 480 LEDs with a power density of 10.8 W·m−2.
Self-powered composites may be essential in the fields of Internet of Things (IoT) and flexible devices. In this regard, CuO/Ti3C2Tx composites have the potential to be integrated into wearable sensors powered by human motion. The improved chemiresistive gas sensor mechanisms can be explained by
O2 (gas) → O2 (ads)
O2 (ads) + e → O2 (ads)
4NH3 + 5O2 (ads) → 4NO + 6H2O + 5e
2NH3 + 3O → N2 + 3H2O + 3e
NH3 + OH → NH2 + H2O + e

5. Applications of Ti3C2Tx MXene for Energy Harvesting

5.1. MXene in Triboelectric Energy

Mechanical energy is converted into useful electrical energy through triboelectrification and electrostatic induction. This phenomenon, commonly observed in our everyday lives, causes random electrostatic charges. When two materials with different electron affinities come into continuous contact and separate, they become oppositely charged through contact electrification, which causes a transfer of electrons from one material surface to another [135,136]. Wang et al. [137] proposed an energy harvesting technique that provides a new approach for producing electricity through frictional energy such as body movements, ocean waves, and mechanical vibrations. The efficiency of power generation depends on the difference in the electron affinity of materials. Material pairs with large electron affinity differences are preferred over materials with poor electron affinities [138]. Most triboelectric devices work on four principal modes: contact-separation, lateral sliding, freestanding, and single electrode modes.
MXenes play a crucial role here because of their unique properties such as high surface area, electrical conductivity, and tunable electronegativity, which make Ti3C2Tx MXenes a suitable candidate for this approach to energy harvesting. Poly (tetfluroethylene) PTFE and fluorinated ethylene propylene (FEP) are the most used as negative friction layers; however, they exhibit poor conductivity. Dong et al. [139] reported the use of Ti3C2Tx MXenes as active electronegative material triboelectric nanogenerators (TENG) for generating high output powers, where Ti3C2Tx MXenes were considered to be as good as PTFE with the PET-ITO electrode (Figure 13) [139,140,141].
Further, MXenes can be used as nanofillers in other materials to improve mechanical, physical, and surface properties. The incorporation of only 4% Ti3C2Tx MXenes into a PVA (polyvinyl alcohol) hydrogel improves conductivity as well as the maximum output voltage of 230 V and output current of 270 nA [142,143]. This system is effectively used for self-recognizing handwriting systems with great sensitivity, reliability, and identifiability as a self-powered sensor. Moreover, MXene-based materials are being studied for use in wearable flexible devices. For example, a stretchable and shape-adaptive TENG system with MXenes and liquid metal electrodes was proposed to drive wearable electronics with open-circuit voltages of 300 V [143]. Overall, the combination of Ti3C2Tx MXene and liquid metal led to excellent fluidity and high electronegativity.
Further, a self-powered 3D-printed smart glove with a Ti3C2Tx MXene/ecoflex nanocomposite was developed for gaming and human-machine interfaces. Adding MXenes into the ecoflex led to high charge density, ultraflexibility, high output TENG performance, and wide pressure detection range (0–120 kPa). Interestingly, Ti3C2Tx MXenes flaxes with added polymer matrices displayed a superior charge retention capability and highly sustainable wearable TENG [144], which led to its application for operating small portable electronic devices such as LEDs, a calculator, and a stopwatch. Ti3C2Tx MXene-based nanocomposites also realized a high dielectric constant and high surface charge density in the poly(vinylidiene flouride-tetrafluoroethylene) PVDF-TrFE polymer matrix [145]. The maximum power obtained was 4.02 W/m2 under a 4 MΩ external resistance.

5.2. MXene in Piezoelectric Energy Harvesting

Piezoelectric energy harvesting is a relatively new technique for converting mechanical energy into electricity. Its simple manufacturing process makes it ideal for harvesting energy and powering small-scale devices. Various materials show piezoelectric effects in response to applied mechanical stress such as semiconductors, ceramics, and polymers. When external force is applied to a piezoelectric material, the dipole moment of the material changes and leads to imbalance at the two poles. The crystal lattice consists of ions with asymmetric charges or molecular groups, and it is responsible for this dipole moment. A typical example of piezoelectricity is the wurtzite structure of the ZnO crystal lattice. The generated charge is collected by charge-collector materials. There is no dipole moment on the ZnO crystal lattice when there is no external force because the two charge centers of anion and cations cancel each other. This original state becomes disturbed through compression or stretching with external force, which separates the cation and anion charge centers. This is how the piezoelectric potential is generated and induces the flow of free electrons through the external circuit to attain a new equilibrium state.
Tan et al. [146] reported the piezoelectricity of 2D Ti3C2Tx MXene in monolayer and multilayer structures with a low band gap. They revealed that the generation of continuous piezoelectric potential occurs in the direction of the armchair at an atomic structure level. The double-layer structure does not have a piezoelectric polarization phenomenon because the same and opposite polarization can cancel or superimpose each other. However, in the monolayer Ti3C2Tx MXene, an inverse piezoelectric effect is inevitable. Li-doped ZnO nanowires grown on MXenes result in the performance enhancement of PENG, where MXenes (Ti3C2Tx) served as ground for the more effective polarization of Li-doped ZnO [147], as shown in Figure 14.
Han et al. [148] designed an anisotropic PVDF/MXene device with directionally oriented foam as the directional sensor and claimed to have a highest sensitivity of 41.3 nA/kPa, which outperforms the other polymer-based piezoelectric sensors. Other reports showed that the incorporation of Ti3C2Tx MXene into the polymer matrix of PVDF enhances the β-phase, which directly affects the output piezoelectric performance [140]. Ma et al. [149] introduced a highly flexible piezoelectric sensor with the qualities of greatly changed interlayer distance under external pressure, high sensitivity, and a fast response up to 30 s. This Ti3C2Tx MXene-based piezoelectric sensor could also detect human activities with extraordinarily reversible compressibility.

5.3. MXene in Thermoelectric Energy Harvesting

This method involves directly converting heat into electricity using Ti3C2Tx MXenes as both bulk and additives in thermoelectric energy harvesting applications. The temperature gradient induced in the device resulted in the diffusion of charged carriers from the hot side to the cold side. One side of the n-type and p-type junction absorbs heat and the other side releases heat when a current is applied to the device. This utilizes energy harvesting and cooling applications without the use of moving parts. Here, Ti3C2Tx MXenes are promising candidates for thermoelectric materials because of their unique electronic structure.
Chang et al. [150] studied the thermoelectric properties of MXenes (ScYCT2) and showed that MXene structures have high thermoelectric properties. The n-type (ScYC(OH)2) showed a maximum power factor of 0.072 W/mK2 at 900 K and a ZT value exceeding 3 and superior electron transfer properties. Further, the thermoelectric properties of Ti3C2Tx MXene and its energy conversion efficiency can be improved effectively by surface modifications [151]. The addition of Ti3C2Tx MXene to the SWCNT film improved the thermoelectric conversion efficiency by 25 times because of its diverse electronic properties [143]. The heterostructure of Ti3C2TxMXene helped enhance the Seeback coefficient up to −32.2 µV K−1.

5.4. MXene for Hydroelectric Energy Harvesting

Ti3C2Tx MXenes are known to harvest hydroelectric energy because of their hydrophilic, high surface area, and three-dimensional (3D) nanosheet structure. Humidity is abundantly available in nature and can be used to harvest clean energy from the environment. Through water vapor pressure, the water molecules adsorb into the micro/nanochannels inside the material of interest and result in an electric charge transfer by creating a ionic concentration gradient inside the material.
Owing to this ionic dissociation of the hydrophilic functional group, the free protons (H+) or mobile charges move along the concentration gradient until it reaches its maximum value. This gradient disappears when the moisture supply is cut off [138,139]. Further, the largely functionalized surface group is a more active material for hydroelectric energy harvesting, therefore, MXenes are suitable materials for hydroelectric energy harvesting because of their availability to functionalized surface groups (-F, -OH, etc.). The porous structure of Ti3C2Tx MXene promotes the transport of water molecules [152]. Recently, the Ti3C2Tx aerogel was used as a moisture-driven energy harvester for electric power generation [153]. The 3D-structured bilayer of Ti3C2Tx MXene with a PAN aerogel resulted in high-output electric energy with a sustained voltage of around 430 mV in a wide range of humidity. The presence of abundant ions such as Ca2+ and Cl- ions helps facilitate the large moisture gradient. In addition, Ti3C2Tx MXenes were discovered to be good for transpiration-driven power generators using water and electrolyte solutions [154]. This was also used to charge a battery with a power density of 24.8 μW cm–2. Li et al. [155] studied Ti3C2Tx MXene-based composite membranes with directional proton diffusion to generate electricity with moisture stimuli. Here, a multifunctional soft actuator was developed, as shown in Figure 15.

5.5. MXene for Solar Energy Harvesting

MXenes have great potential as a future material in solar cells. The continuous improvement in solar cell systems is expected to replace conventional materials with 2D materials. Two-dimensional semiconductor materials such as graphene, transition-metal dichalcogenide, and black phosphorous are potential candidates for making next-generation ultrathin solar cells. Recently, perovskite solar cells (PSCs) with perovskite crystal structures were highly studied for their quick energy conversion efficiencies [156]. The distortions in the crystallographic phase transition in the perovskite crystal structure were induced by varying temperature from the cubic to the tetragonal phase. Ti3C2Tx MXenes have also been used as additives in PSCs to overcome the potential challenges with PSCs, including their high instability upon exposure to air. Ti3C2Tx MXene has also been used to enhance carrier mobility and to improve the performance of PSCs by retarding the crystallization rate [157]. The charge transfer is facilitated through high electrical conductivity and the mobility of Ti3C2Tx MXenes. Further, the addition of Ti3C2Tx MXene in the perovskite film can shift the work function [149]. Zheng et al. [158] showed the quality-controlling ability of SiO2/ Ti3C2Tx MXene in a perovskite film with defect density, crystal size, etc. Similarly, interfacial defects are reduced with the addition of Ti3C2Tx QDs (TQDs), and they improve the carrier mobility of perovskite film (Figure 16) [156,159]. Thus, Ti3C2Tx MXene has the potential to be used as an additive in perovskite film and improve the performance of devices.
Ti3C2Tx MXene is also used as a flexible electrode material in organic solar cells (OSCs). OSCs are promising technologies for wearable and super flexible future electronics because of their qualities such as flexibility, light weight, and portability. Although the indium tin-oxide (ITO) electrode is used because of its high conductivity and transparency, ITO-based electrodes need to be replaced with other flexible electrodes when developing ultraflexible wearable electronic devices. Nirmal et al. [160] introduced a multilayer Ti3C2Tx MXene/Ag/MXene structure as a flexible electrode with sustained photovoltaic performance and memory retention functionalities. This device showed reliable resistive switching behavior at low operating voltages of 0.60 and −0.33 V, and stable endurance performance (4 × 103). Further, a Ti3C2Tx MXene/Ag NWs/ colorless polyimide hybrid electrode with high conductivity was fabricated with considerably lower sheet resistance (13.08 Ohm sq−1) [160,161]. The designed CuSe/Ti3C2Tx CE counter electrode showed excellent electrocatalytic activity towards polysulfide redox reactions because of its unique three-dimensional structure. Further, Ti3C2Tx MXene when added to a zinc oxide (ZnO) precursor can also be utilized in the electron transport layer (ETL) [161,162]. The optimized concentration of Ti3C2Tx with 0.5 and 2 wt% in the ETL exhibited power conversion efficiencies of 14.1% and 13.7%, respectively. The addition of MXene materials is a promising approach for improving the efficiency and stability of OSCs.

6. Conclusions and Outlook

The review summarized the synthesis and applications of Ti3C2Tx MXenes and their composites for toxic gas sensors. A brief review of MXene-based energy harvesting from various ambient sources was also provided, including thermoelectric energy, piezoelectric and triboelectric energy, hydroelectric energy, and solar energy. The effects of etching techniques (HF etching, HF-free etching), intercalation, and exfoliation on the morphology and structure of MXenes was examined. However, the review focused on HF etching because it is well-established and easy to use.
The hydrophilicity, surface tunability, high electrical conductivity, high surface area, and good optical transmittance of Ti3C2Tx MXenes provide a tremendous opportunity for these materials to be utilized in various areas of gas sensors and energy harvesting for realizing advanced device performances. The high electrical conductivity and rich surface termination groups may be the key features of Ti3C2Tx MXene for applications in those fields. The hybridization of Ti3C2Tx MXene with other materials to form functional composites is a good strategy to boost performance to an even higher level. Although the proposed mechanisms of gas adsorption and the resulting electron transport remain debatable, we are indeed moving closer toward filling the gaps between known facts.
More MXenes and MXene-based composites must be explored in future to find their unparalleled qualities and expand their breadth of applications. Gas sensors are expected to become one of the key sectors that need these advanced nanomaterials.

Author Contributions

Conceptualization, methodology, validation, formal analysis, investigation, resources, data curation, Q.T.H.T.; writing—original draft preparation, D.T.; writing—review and editing, visualization, supervision, project administration, funding acquisition, J.-S.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by a National Research Foundation of Korea (NRF) grant funded by the Korean government (MSIT) (No. 2019R1A2C1008746).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cristoloveanu, S.; Wan, J.; Zaslavsky, A. A Review of Sharp-Switching Devices for Ultra-Low Power Applications. IEEE J. Electron Devices Soc. 2016, 4, 215–226. [Google Scholar] [CrossRef]
  2. Raza, U.; Kulkarni, P.; Sooriyabandara, M. Low Power Wide Area Networks: An Overview. IEEE Commun. Surv. Tutor. 2017, 19, 855–873. [Google Scholar] [CrossRef]
  3. Nazir, G.; Rehman, A.; Park, S.-J. Energy-Efficient Tunneling Field-Effect Transistors for Low-Power Device Applications: Challenges and Opportunities. ACS Appl. Mater. Interfaces 2020, 12, 47127–47163. [Google Scholar] [CrossRef] [PubMed]
  4. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A. Electric Field Effect in Atomically Thin Carbon Films. Science 2004, 306, 666–669. [Google Scholar] [CrossRef] [PubMed]
  5. Alhabeb, M.; Maleski, K.; Anasori, B.; Lelyukh, P.; Clark, L.; Sin, S.; Gogotsi, Y. Guidelines for Synthesis and Processing of Two-Dimensional Titanium Carbide (Ti3C2Tx MXene). Chem. Mater. 2017, 29, 7633–7644. [Google Scholar] [CrossRef]
  6. Zhang, H.-F.; Xuan, J.-Y.; Zhang, Q.; Sun, M.-L.; Jia, F.-C.; Wang, X.-M.; Yin, G.-C.; Lu, S.-Y. Strategies and Challenges for Enhancing Performance of MXene-Based Gas Sensors: A Review. Rare Met. 2022, 41, 3976–3999. [Google Scholar]
  7. Hong, L.; Guo, R.; Yuan, Y.; Ji, X.; Li, Z.; Lin, Z.; Pan, W. Recent Progress of Two-Dimensional MXenes in Photocatalytic Applications: A Review. Mater. Today Energy 2020, 18, 100521. [Google Scholar] [CrossRef]
  8. Wang, Z.; Wang, F.; Hermawan, A.; Asakura, Y.; Hasegawa, T.; Kumagai, H.; Kato, H.; Kakihana, M.; Zhu, J.; Yin, S. SnO-SnO2 Modified Two-Dimensional MXene Ti3C2Tx for Acetone Gas Sensor Working at Room Temperature. J. Mater. Sci. Technol. 2021, 73, 128–138. [Google Scholar] [CrossRef]
  9. Wang, X.; Sun, K.; Li, K.; Li, X.; Gogotsi, Y. Ti3C2Tx/PEDOT: PSS Hybrid Materials for Room-Temperature Methanol Sensor. Chin. Chem. Lett. 2020, 31, 1018–1021. [Google Scholar] [CrossRef]
  10. 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]
  11. Tran, N.M.; Ta, Q.T.H.; Noh, J.-S. Unusual Synthesis of Safflower-Shaped TiO2/Ti3C2 Heterostructures Initiated from Two-Dimensional Ti3C2 MXene. Appl. Surf. Sci. 2021, 538, 148023. [Google Scholar] [CrossRef]
  12. Sreedhar, A.; Ta, Q.T.H.; Noh, J.-S. 2D Ti3C2 MXene Interfaced ZnO/WO3 Thin Film Nanostructures towards Improved Photoelectrochemical Water Splitting. J. Electroanal. Chem. 2023, 940, 117509. [Google Scholar] [CrossRef]
  13. My Tran, N.; Thanh Hoai Ta, Q.; Sreedhar, A.; Noh, J.S. Ti3C2Tx MXene Playing as a Strong Methylene Blue Adsorbent in Wastewater. Appl. Surf. Sci. 2021, 537, 148006. [Google Scholar] [CrossRef]
  14. Nguyen, V.-H.; Nguyen, B.-S.; Hu, C.; Nguyen, C.C.; Nguyen, D.L.T.; Nguyen Dinh, M.T.; Vo, D.-V.N.; Trinh, Q.T.; Shokouhimehr, M.; Hasani, A. Novel Architecture Titanium Carbide (Ti3C2Tx) MXene Cocatalysts toward Photocatalytic Hydrogen Production: A Mini-Review. Nanomaterials 2020, 10, 602. [Google Scholar] [CrossRef] [PubMed]
  15. Mohammadi, A.V.; Rosen, J.; Gogotsi, Y. The World of Two-Dimensional Carbides and Nitrides (MXenes). Science 2021, 372, eabf1581. [Google Scholar] [CrossRef]
  16. Li, Q.; Li, Y.; Zeng, W. Preparation and Application of 2D MXene-Based Gas Sensors: A Review. Chemosensors 2021, 9, 225. [Google Scholar] [CrossRef]
  17. Wu, Z.; Shen, J.; Li, C.; Zhang, C.; Feng, K.; Wang, Z.; Wang, X.; Meira, D.M.; Cai, M.; Zhang, D. Mo2TiC2 MXene-Supported Ru Clusters for Efficient Photothermal Reverse Water–Gas Shift. ACS Nano 2022, 17, 1550–1559. [Google Scholar] [CrossRef]
  18. Gandla, D.; Zhang, F.; Tan, D.Q. Advantage of Larger Interlayer Spacing of a Mo2Ti2C3 MXene Free-Standing Film Electrode toward an Excellent Performance Supercapacitor in a Binary Ionic Liquid–Organic Electrolyte. ACS Omega 2022, 7, 7190–7198. [Google Scholar] [CrossRef]
  19. Tran, V.A.; Tran, N.T.; Doan, V.D.; Nguyen, T.-Q.; Thi, H.H.P.; Vo, G.N.L. Application Prospects of MXenes Materials Modifications for Sensors. Micromachines 2023, 14, 247. [Google Scholar] [CrossRef]
  20. Ji, H.; Zeng, W.; Li, Y. Gas Sensing Mechanisms of Metal Oxide Semiconductors: A Focus Review. Nanoscale 2019, 11, 22664–22684. [Google Scholar] [CrossRef]
  21. Ta, Q.T.H.; Tran, N.M.; Tri, N.N.; Sreedhar, A.; Noh, J.S. Highly Surface-Active Si-Doped TiO2/Ti3C2Tx Heterostructure for Gas Sensing and Photodegradation of Toxic Matters. Chem. Eng. J. 2021, 425, 131437. [Google Scholar] [CrossRef]
  22. Wu, M.; He, M.; Hu, Q.; Wu, Q.; Sun, G.; Xie, L.; Zhang, Z.; Zhu, Z.; Zhou, A. Ti3C2 MXene-Based Sensors with High Selectivity for NH3 Detection at Room Temperature. ACS Sens. 2019, 4, 2763–2770. [Google Scholar] [CrossRef] [PubMed]
  23. Ta, Q.T.H.; Tran, N.M.; Noh, J.-S. Rice Crust-Like ZnO/Ti3C2Tx MXene Hybrid Structures for Improved Photocatalytic Activity. Catalysts 2020, 10, 1140. [Google Scholar] [CrossRef]
  24. Gogotsi, Y. MXenes: From Discovery to Applications of Two-Dimensional Metal Carbides and Nitrides; CRC Press: Boca Raton, FL, USA, 2023; ISBN 1000618943. [Google Scholar]
  25. Bhati, V.S.; Kumar, M.; Banerjee, R. Gas Sensing Performance of 2D Nanomaterials/Metal Oxide Nanocomposites: A Review. J. Mater. Chem. C 2021, 9, 8776–8808. [Google Scholar] [CrossRef]
  26. Cao, M.; Wang, F.; Wang, L.; Wu, W.; Lv, W.; Zhu, J. Room Temperature Oxidation of Ti3C2 MXene for Supercapacitor Electrodes. J. Electrochem. Soc. 2017, 164, A3933–A3942. [Google Scholar] [CrossRef]
  27. Cao, F.; Zhang, Y.; Wang, H.; Khan, K.; Tareen, A.K.; Qian, W.; Zhang, H.; Ågren, H. Recent Advances in Oxidation Stable Chemistry of 2D MXenes. Adv. Mater. 2022, 34, 2107554. [Google Scholar] [CrossRef] [PubMed]
  28. Pei, Y.; Zhang, X.; Hui, Z.; Zhou, J.; Huang, X.; Sun, G.; Huang, W. Ti3C2TX MXene for Sensing Applications: Recent Progress, Design Principles, and Future Perspectives. ACS Nano 2021, 15, 3996–4017. [Google Scholar] [CrossRef]
  29. Zhang, H.; Li, M.; Zhu, C.; Tang, Q.; Kang, P.; Cao, J. Preparation of Magnetic α-Fe2O3/ZnFe2O4@Ti3C2 MXene with Excellent Photocatalytic Performance. Ceram. Int. 2020, 46, 81–88. [Google Scholar] [CrossRef]
  30. Shahzad, A.; Rasool, K.; Nawaz, M.; Miran, W.; Jang, J.; Moztahida, M.; Mahmoud, K.A.; Lee, D.S. Heterostructural TiO2/Ti3C2Tx (MXene) for Photocatalytic Degradation of Antiepileptic Drug Carbamazepine. Chem. Eng. J. 2018, 349, 748–755. [Google Scholar] [CrossRef]
  31. Thi, Q.V.; Han, J.; Park, J.; Kim, S.-S.; Jeon, D.-Y.; Joo, Y. Printable and Recyclable Carbon Nanotube Electronics with Degradable Soybean Oil-Based Polycationic Substrate as Gate Dielectrics. Carbon 2023, 212, 118089. [Google Scholar] [CrossRef]
  32. Kim, S.J.; Koh, H.J.; Ren, C.E.; Kwon, O.; Maleski, K.; Cho, S.Y.; Anasori, B.; Kim, C.K.; Choi, Y.K.; Kim, J.; et al. Metallic Ti3C2Tx MXene Gas Sensors with Ultrahigh Signal-to-Noise Ratio. ACS Nano 2018, 12, 986–993. [Google Scholar] [CrossRef] [PubMed]
  33. 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]
  34. Hong, S.; Al Marzooqi, F.; El-Demellawi, J.K.; Al Marzooqi, N.; Arafat, H.A.; Alshareef, H.N. Ion-Selective Separation Using MXene-Based Membranes: A Review. ACS Mater. Lett. 2023, 5, 341–356. [Google Scholar] [CrossRef]
  35. Verma, R.; Thakur, P.; Chauhan, A.; Jasrotia, R.; Thakur, A. A Review on MXene and Its’ Composites for Electromagnetic Interference (EMI) Shielding Applications. Carbon 2023, 208, 170–190. [Google Scholar] [CrossRef]
  36. Li, Y.; Huang, S.; Peng, S.; Jia, H.; Pang, J.; Ibarlucea, B.; Hou, C.; Cao, Y.; Zhou, W.; Liu, H. Toward Smart Sensing by MXene. Small 2023, 19, 2206126. [Google Scholar] [CrossRef] [PubMed]
  37. Huang, W.-X.; Li, Z.-P.; Li, D.-D.; Hu, Z.-H.; Wu, C.; Lv, K.-L.; Li, Q. Ti3C2 MXene: Recent Progress in Its Fundamentals, Synthesis, and Applications. Rare Met. 2022, 41, 3268–3300. [Google Scholar] [CrossRef]
  38. Li, J.; Chen, X.; Zhu, X.; Jiang, Y.; Chang, X.; Sun, S. Two-Dimensional Transition Metal MXene-Based Gas Sensors: A Review. Chinese Chem. Lett. 2023, 108286. [Google Scholar] [CrossRef]
  39. Li, H.; Sun, B.; Gao, T.; Li, H.; Ren, Y.; Zhou, G. Ti3C2 MXene Co-Catalyst Assembled with Mesoporous TiO2 for Boosting Photocatalytic Activity of Methyl Orange Degradation and Hydrogen Production. Chin. J. Catal. 2022, 43, 461–471. [Google Scholar] [CrossRef]
  40. Barsoum, M.W.; Gogotsi, Y. Removing Roadblocks and Opening New Opportunities for MXenes. Ceram. Int. 2023, 49, 24112–24122. [Google Scholar] [CrossRef]
  41. Simonenko, E.P.; Simonenko, N.P.; Mokrushin, A.S.; Simonenko, T.L.; Gorobtsov, P.Y.; Nagornov, I.A.; Korotcenkov, G.; Sysoev, V.V.; Kuznetsov, N.T. Application of Titanium Carbide MXenes in Chemiresistive Gas Sensors. Nanomaterials 2023, 13, 850. [Google Scholar] [CrossRef]
  42. Khan, K.; Tareen, A.K.; Iqbal, M.; Ye, Z.; Xie, Z.; Mahmood, A.; Mahmood, N.; Zhang, H. Recent Progress in Emerging Novel MXenes Based Materials and Their Fascinating Sensing Applications. Small 2023, 19, e2206147. [Google Scholar] [CrossRef] [PubMed]
  43. Janjhi, F.A.; Ihsanullah, I.; Bilal, M.; Castro-Muñoz, R.; Boczkaj, G.; Gallucci, F. MXene-Based Materials for Removal of Antibiotics and Heavy Metals from Wastewater–A Review. Water Resour. Ind. 2023, 29, 100202. [Google Scholar] [CrossRef]
  44. Liu, X.; Li, M.; Zhang, Y.; Zhu, X.; Guan, Y. Self-Powered Wireless Sensor Node Based on RF Energy Harvesting and Management Combined Design. Energy Convers. Manag. 2023, 292, 117393. [Google Scholar] [CrossRef]
  45. Miao, L.; Huo, Z.-M.; Rong, Y.; Mu, H.-W.; Sun, Z.-X. IoT Adaptive Threshold Energy Management Algorithm Based on Energy Harvesting. Ad Hoc Netw. 2023, 149, 103241. [Google Scholar] [CrossRef]
  46. Su, Y.; Chen, S.; Liu, B.; Lu, H.; Luo, X.; Chen, C.; Li, W.; Long, Y.; Tai, H.; Xie, G. Maxwell Displacement Current Induced Wireless Self-Powered Gas Sensor Array. Mater. Today Phys. 2023, 30, 100951. [Google Scholar] [CrossRef]
  47. Ghidiu, M.; Lukatskaya, M.R.; Zhao, M.Q.; Gogotsi, Y.; Barsoum, M.W. Conductive Two-Dimensional Titanium Carbide “clay” with High Volumetric Capacitance. Nature 2015, 516, 78–81. [Google Scholar] [CrossRef] [PubMed]
  48. Shuck, C.E.; Han, M.; Maleski, K.; Hantanasirisakul, K.; Kim, S.J.; Choi, J.; Reil, W.E.B.; Gogotsi, Y. Effect of Ti3AlC2 MAX Phase on Structure and Properties of Resultant Ti3C2Tx MXene. ACS Appl. Nano Mater. 2019, 2, 3368–3376. [Google Scholar] [CrossRef]
  49. Shahzad, F.; Alhabeb, M.; Hatter, C.B.; Anasori, B.; Hong, S.M.; Koo, C.M.; Gogotsi, Y. Electromagnetic Interference Shielding with 2D Transition Metal Carbides (MXenes). Science 2016, 353, 1137–1140. [Google Scholar] [CrossRef]
  50. Cai, L.; Huang, Z.; Hu, W.; Hao, S.; Zhai, H.; Zhou, Y. Fabrication, Mechanical Properties, and Tribological Behaviors of Ti2AlC and Ti2AlSn0.2C Solid Solutions. J. Adv. Ceram. 2017, 6, 90–99. [Google Scholar] [CrossRef]
  51. Wang, X.; Mathis, T.S.; Li, K.; Lin, Z.; Vlcek, L.; Torita, T.; Osti, N.C.; Hatter, C.; Urbankowski, P.; Sarycheva, A.; et al. Influences from Solvents on Charge Storage in Titanium Carbide MXenes. Nat. Energy 2019, 4, 241–248. [Google Scholar] [CrossRef]
  52. Uzun, S.; Seyedin, S.; Stoltzfus, A.L.; Levitt, A.S.; Alhabeb, M.; Anayee, M.; Strobel, C.J.; Razal, J.M.; Dion, G.; Gogotsi, Y. Knittable and Washable Multifunctional MXene-Coated Cellulose Yarns. Adv. Funct. Mater. 2019, 29, 1905015. [Google Scholar] [CrossRef]
  53. Sreedhar, A.; Ta, Q.T.H.; Noh, J.-S. Role of Pn Junction Initiated Mixed-Dimensional 0D/2D, 1D/2D, and 2D/2D BiOX (X= Cl, Br, and I)/TiO2 Nanocomposite Interfaces for Environmental Remediation Applications: A Review. Chemosphere 2022, 305, 135478. [Google Scholar] [CrossRef]
  54. Perera, A.; Madhushani, K.A.U.; Punchihewa, B.T.; Kumar, A.; Gupta, R.K. MXene-Based Nanomaterials for Multifunctional Applications. Materials 2023, 16, 1138. [Google Scholar] [CrossRef] [PubMed]
  55. Nahirniak, S.; Ray, A.; Saruhan, B. Challenges and Future Prospects of the MXene-Based Materials for Energy Storage Applications. Batteries 2023, 9, 126. [Google Scholar] [CrossRef]
  56. Li, Z.; Yu, L.; Milligan, C.; Ma, T.; Zhou, L.; Cui, Y.; Qi, Z.; Libretto, N.; Xu, B.; Luo, J.; et al. Two-Dimensional Transition Metal Carbides as Supports for Tuning the Chemistry of Catalytic Nanoparticles. Nat. Commun. 2018, 9, 5258. [Google Scholar] [CrossRef]
  57. Naguib, M.; Mashtalir, O.; Carle, J.; Presser, V.; Lu, J.; Hultman, L.; Gogotsi, Y.; Barsoum, M.W. Two-Dimensional Transition Metal Carbides. ACS Nano 2012, 6, 1322–1331. [Google Scholar] [CrossRef]
  58. Xiao, X.; Wang, H.; Urbankowski, P.; Gogotsi, Y. Topochemical Synthesis of 2D Materials. Chem. Soc. Rev. 2018, 47, 8744–8765. [Google Scholar] [CrossRef]
  59. Shuck, C.E.; Gogotsi, Y. Taking MXenes from the Lab to Commercial Products. Chem. Eng. J. 2020, 401, 125786. [Google Scholar] [CrossRef]
  60. Zhu, J.F.; Qi, G.Q.; Wang, F.; Yang, H.B. Synthesis of Ti3AlC2/Al2O3 Nanopowders by Mechano-Chemical Reaction. Adv. Powder Technol. 2010, 21, 578–581. [Google Scholar] [CrossRef]
  61. Wang, X.; Wu, L.; Gao, H.; Zhang, X. Synthesis of Ti3AlC2 and electrochemical performance of Ti3C2Tx nanosheet electrode. Sci. Sin. Chim. 2018, 48, 289–297. [Google Scholar] [CrossRef]
  62. Shahin, N.; Kazemi, S.; Heidarpour, A. Mechanochemical Synthesis Mechanism of Ti3AlC2 MAX Phase from Elemental Powders of Ti, Al and C. Adv. Powder Technol. 2016, 27, 1775–1780. [Google Scholar] [CrossRef]
  63. An, H.; Habib, T.; Shah, S.; Gao, H.; Radovic, M.; Green, M.J.; Lutkenhaus, J.L. Surface-Agnostic Highly Stretchable and Bendable Conductive MXene Multilayers. Sci. Adv. 2018, 4, eaaq0118. [Google Scholar] [CrossRef] [PubMed]
  64. Jawaid, A.; Hassan, A.; Neher, G.; Nepal, D.; Pachter, R.; Kennedy, W.J.; Ramakrishnan, S.; Vaia, R.A. Halogen Etch of Ti3AlC2 MAX Phase for MXene Fabrication. ACS Nano 2021, 15, 2771–2777. [Google Scholar]
  65. Shi, H.; Zhang, P.; Liu, Z.; Park, S.; Lohe, M.R.; Wu, Y.; Shaygan Nia, A.; Yang, S.; Feng, X. Ambient-stable Two-dimensional Titanium Carbide (MXene) Enabled by Iodine Etching. Angew. Chemie Int. Ed. 2021, 60, 8689–8693. [Google Scholar]
  66. Yang, S.; Zhang, P.; Wang, F.; Ricciardulli, A.G.; Lohe, M.R.; Blom, P.W.M.; Feng, X. Fluoride-free Synthesis of Two-dimensional Titanium Carbide (MXene) Using a Binary Aqueous System. Angew. Chemie 2018, 130, 15717–15721. [Google Scholar]
  67. Li, T.; Yao, L.; Liu, Q.; Gu, J.; Luo, R.; Li, J.; Yan, X.; Wang, W.; Liu, P.; Chen, B. Fluorine-free Synthesis of High-purity Ti3C2Tx (T= OH, O) via Alkali Treatment. Angew. Chemie Int. Ed. 2018, 57, 6115–6119. [Google Scholar] [CrossRef] [PubMed]
  68. Peng, W.; Kuai, Z.; Peng, J.; Guo, W.; Li, H.; Peng, C.; Chen, S. MXene-Based Gold Nanoparticle Catalyst with High Activity and Stability for 4-Nitrophenol Reduction. J. Solid State Chem. 2023, 325, 124166. [Google Scholar]
  69. Hope, M.A.; Forse, A.C.; Griffith, K.J.; Lukatskaya, M.R.; Ghidiu, M.; Gogotsi, Y.; Grey, C.P. NMR Reveals the Surface Functionalisation of Ti3C2 MXene. Phys. Chem. Chem. Phys. 2016, 18, 5099–5102. [Google Scholar] [CrossRef]
  70. Venkateshalu, S.; Grace, A.N. MXenes—A New Class of 2D Layered Materials: Synthesis, Properties, Applications as Supercapacitor Electrode and Beyond. Appl. Mater. Today 2020, 18, 100509. [Google Scholar]
  71. Ding, L.; Wei, Y.; Li, L.; Zhang, T.; Wang, H.; Xue, J.; Ding, L.X.; Wang, S.; Caro, J.; Gogotsi, Y. MXene Molecular Sieving Membranes for Highly Efficient Gas Separation. Nat. Commun. 2018, 9, 155. [Google Scholar] [CrossRef]
  72. Malaki, M.; Maleki, A.; Varma, R.S. MXenes and Ultrasonication. J. Mater. Chem. A 2019, 7, 10843–10857. [Google Scholar] [CrossRef]
  73. Zhang, P.; Wang, F.; Yu, M.; Zhuang, X.; Feng, X. Two-Dimensional Materials for Miniaturized Energy Storage Devices: From Individual Devices to Smart Integrated Systems. Chem. Soc. Rev. 2018, 47, 7426–7451. [Google Scholar] [PubMed]
  74. Hemanth, N.R.; Kandasubramanian, B. Recent Advances in 2D MXenes for Enhanced Cation Intercalation in Energy Harvesting Applications: A Review. Chem. Eng. J. 2020, 392, 123678. [Google Scholar]
  75. Lukatskaya, M.R.; Kota, S.; Lin, Z.; Zhao, M.-Q.; Shpigel, N.; Levi, M.D.; Halim, J.; Taberna, P.-L.; Barsoum, M.W.; Simon, P. Ultra-High-Rate Pseudocapacitive Energy Storage in Two-Dimensional Transition Metal Carbides. Nat. Energy 2017, 2, 17105. [Google Scholar]
  76. Li, X.; Bai, Y.; Shi, X.; Su, N.; Nie, G.; Zhang, R.; Nie, H.; Ye, L. Applications of MXene (Ti3C2Tx) in Photocatalysis: A Review. Mater. Adv. 2021, 2, 1570–1594. [Google Scholar] [CrossRef]
  77. Zhu, J.; Tang, Y.; Yang, C.; Wang, F.; Cao, M. Composites of TiO2 Nanoparticles Deposited on Ti3C2 MXene Nanosheets with Enhanced Electrochemical Performance. J. Electrochem. Soc. 2016, 163, A785–A791. [Google Scholar] [CrossRef]
  78. Ding, X.; Li, Y.; Li, C.; Wang, W.; Wang, L.; Feng, L.; Han, D. 2D Visible-Light-Driven TiO2@Ti3C2/g-C3N4 Ternary Heterostructure for High Photocatalytic Activity. J. Mater. Sci. 2019, 54, 9385–9396. [Google Scholar] [CrossRef]
  79. Anasori, B.; Lukatskaya, M.R.; Gogotsi, Y. 2D Metal Carbides and Nitrides (MXenes) for Energy Storage. Nat. Rev. Mater. 2017, 2, 16098. [Google Scholar]
  80. Chae, Y.; Kim, S.J.; Cho, S.-Y.; Choi, J.; Maleski, K.; Lee, B.-J.; Jung, H.-T.; Gogotsi, Y.; Lee, Y.; Ahn, C.W. An Investigation into the Factors Governing the Oxidation of Two-Dimensional Ti3C2 MXene. Nanoscale 2019, 11, 8387–8393. [Google Scholar]
  81. Li, K.; Wang, X.; Li, S.; Urbankowski, P.; Li, J.; Xu, Y.; Gogotsi, Y. An Ultrafast Conducting Polymer@MXene Positive Electrode with High Volumetric Capacitance for Advanced Asymmetric Supercapacitors. Small 2020, 16, e1906851. [Google Scholar] [CrossRef]
  82. Hu, T.; Li, Z.; Hu, M.; Wang, J.; Hu, Q.; Li, Q.; Wang, X. Chemical Origin of Termination-Functionalized MXenes: Ti3C2T2 as a Case Study. J. Phys. Chem. C 2017, 121, 19254–19261. [Google Scholar]
  83. Wang, D.D.; Tian, W.B.; Lu, C.J.; Ding, J.X.; Zhu, Y.F.; Zhang, M.; Zhang, P.G.; Sun, Z.M. Comparison of the Interfacial Reactions and Properties between Ag/Ti3AlC2 and Ag/Ti3SiC2 Electrical Contact Materials. J. Alloys Compd. 2021, 857, 157588. [Google Scholar] [CrossRef]
  84. Son, W.; Gao, H.; Duong, T.; Talapatra, A.; Radovic, M.; Arróyave, R. Effect of A Mixing on Elastic Modulus, Cleavage Stress, and Shear Stress in the Ti3(SixAl1-x) C2 MAX Phase. Phys. Rev. B 2017, 95, 235131. [Google Scholar] [CrossRef]
  85. Husairi, F.S.; Eswar, K.A.; Azlinda, A.; Rusop, M.; Abdullah, S. Effect of Urea as a Stabiliser in the Thermal Immersion Method to Synthesise Zinc Oxide Nanostructures on Porous Silicon Nanostructures. Adv. Mater. Res. 2014, 832, 644–648. [Google Scholar] [CrossRef]
  86. Wang, H.; Sun, Y.; Wu, Y.; Tu, W.; Wu, S.; Yuan, X.; Zeng, G.; Xu, Z.J.; Li, S.; Chew, J.W. Electrical Promotion of Spatially Photoinduced Charge Separation via Interfacial-Built-in Quasi-Alloying Effect in Hierarchical Zn2In2S5/Ti3C2(O, OH)x Hybrids toward Efficient Photocatalytic Hydrogen Evolution and Environmental Remediation. Appl. Catal. B Environ. 2019, 245, 290–301. [Google Scholar]
  87. Junkaew, A.; Arróyave, R. Enhancement of the Selectivity of MXenes (M2C, M = Ti, V, Nb, Mo) via Oxygen-Functionalization: Promising Materials for Gas-Sensing and -Separation. Phys. Chem. Chem. Phys. 2018, 20, 6073–6082. [Google Scholar] [CrossRef] [PubMed]
  88. Qian, Y.; Wei, H.; Dong, J.; Du, Y.; Fang, X.; Zheng, W.; Sun, Y.; Jiang, Z. Fabrication of Urchin-like ZnO-MXene Nanocomposites for High-Performance Electromagnetic Absorption. Ceram. Int. 2017, 43, 10757–10762. [Google Scholar] [CrossRef]
  89. Zhao, Q.; Zhu, Q.; Miao, J.; Zhang, P.; Wan, P.; He, L.; Xu, B. Flexible 3D Porous MXene Foam for High-Performance Lithium-Ion Batteries. Small 2019, 15, e1904293. [Google Scholar] [CrossRef]
  90. Zhang, M.; Qin, J.; Rajendran, S.; Zhang, X.; Liu, R. Heterostructured D-Ti3 C2/TiO2/g-C3N4 Nanocomposites with Enhanced Visible-Light Photocatalytic Hydrogen Production Activity. ChemSusChem 2018, 11, 4226–4236. [Google Scholar] [CrossRef]
  91. Gasso, S.; Mahajan, A. Development of Highly Sensitive and Humidity Independent Room Temeprature NO2 Gas Sensor Using Two Dimensional Ti3C2Tx Nanosheets and One Dimensional WO3 Nanorods Nanocomposite. ACS Sens. 2022, 7, 2454–2464. [Google Scholar] [CrossRef]
  92. Liu, X.; Zhang, H.; Song, Y.; Shen, T.; Sun, J. Facile Solvothermal Synthesis of ZnO/Ti3C2Tx MXene Nanocomposites for NO2 Detection at Low Working Temperature. Sens. Actuators B Chem. 2022, 367, 132025. [Google Scholar] [CrossRef]
  93. Hantanasirisakul, K.; Gogotsi, Y. Electronic and Optical Properties of 2D Transition Metal Carbides and Nitrides (MXenes). Adv. Mater. 2018, 30, 1804779. [Google Scholar] [CrossRef] [PubMed]
  94. Lipatov, A.; Alhabeb, M.; Lukatskaya, M.R.; Boson, A.; Gogotsi, Y.; Sinitskii, A. Effect of Synthesis on Quality, Electronic Properties and Environmental Stability of Individual Monolayer Ti3C2 MXene Flakes. Adv. Electron. Mater. 2016, 2, 1600255. [Google Scholar] [CrossRef]
  95. Doan, T.H.P.; Hong, W.G.; Noh, J.-S. Palladium Nanoparticle-Decorated Multi-Layer Ti3C2Tx Dual-Functioning as a Highly Sensitive Hydrogen Gas Sensor and Hydrogen Storage. RSC Adv. 2021, 11, 7492–7501. [Google Scholar] [CrossRef]
  96. Zhu, Z.; Liu, C.; Jiang, F.; Liu, J.; Ma, X.; Liu, P.; Xu, J.; Wang, L.; Huang, R. Flexible and Lightweight Ti3C2Tx MXene@ Pd Colloidal Nanoclusters Paper Film as Novel H2 Sensor. J. Hazard. Mater. 2020, 399, 123054. [Google Scholar] [CrossRef]
  97. Nam, M.-S.; Kim, J.-Y.; Mirzaei, A.; Kim, H.W.; Kim, S.S. Self-Heated, Flexible, and Selective H2 Gas Sensors Based on Pd-Decorated Ti3c2Tx Mxenes. Available online: http://dx.doi.org/10.2139/ssrn.4367681 (accessed on 26 June 2023).
  98. Yuan, W.; Yang, K.; Peng, H.; Li, F.; Yin, F. A Flexible VOCs Sensor Based on a 3D Mxene Framework with a High Sensing Performance. J. Mater. Chem. A 2018, 6, 18116–18124. [Google Scholar] [CrossRef]
  99. Sehab, N.; Amara, A.; Mebdoua, Y. Mechanical Activation-Assisted Electro-Thermal Explosion Synthesis of the Ti2AlC Phase. Mater. Manuf. Process. 2017, 32, 1260–1264. [Google Scholar] [CrossRef]
  100. Hermawan, A.; Zhang, B.; Taufik, A.; Asakura, Y.; Hasegawa, T.; Zhu, J.; Shi, P.; Yin, S. CuO Nanoparticles/Ti3C2Tx MXene Hybrid Nanocomposites for Detection of Toluene Gas. ACS Appl. Nano Mater. 2020, 3, 4755–4766. [Google Scholar] [CrossRef]
  101. Liu, Z.; Lv, H.; Xie, Y.; Wang, J.; Fan, J.; Sun, B.; Jiang, L.; Zhang, Y.; Wang, R.; Shi, K. A 2D/2D/2D Ti3C2TX@TiO2@MoS2 Heterostructure as an Ultrafast and High-Sensitivity NO2 Gas Sensor at Room-Temperature. J. Mater. Chem. A 2022, 10, 11980–11989. [Google Scholar] [CrossRef]
  102. He, T.; Liu, W.; Lv, T.; Ma, M.; Liu, Z.; Vasiliev, A.; Li, X. MXene/SnO2 Heterojunction Based Chemical Gas Sensors. Sens. Actuators B Chem. 2021, 329, 129275. [Google Scholar] [CrossRef]
  103. Ta, Q.T.H.; Tri, N.N.; Noh, J.-S. Improved NO2 Gas Sensing Performance of 2D MoS2/Ti3C2Tx MXene Nanocomposite. Appl. Surf. Sci. 2022, 604, 154624. [Google Scholar]
  104. Zhang, D.; Mi, Q.; Wang, D.; Li, T. MXene/Co3O4 Composite Based Formaldehyde Sensor Driven by ZnO/MXene Nanowire Arrays Piezoelectric Nanogenerator. Sensors Actuators B Chem. 2021, 339, 129923. [Google Scholar] [CrossRef]
  105. Yang, Z.; Jiang, L.; Wang, J.; Liu, F.; He, J.; Liu, A.; Lv, S.; You, R.; Yan, X.; Sun, P. Flexible Resistive NO2 Gas Sensor of Three-Dimensional Crumpled MXene Ti3C2Tx/ZnO Spheres for Room Temperature Application. Sens. Actuators B Chem. 2021, 326, 128828. [Google Scholar] [CrossRef]
  106. Guo, X.; Ding, Y.; Kuang, D.; Wu, Z.; Sun, X.; Du, B.; Liang, C.; Wu, Y.; Qu, W.; Xiong, L. Enhanced Ammonia Sensing Performance Based on MXene-Ti3C2Tx Multilayer Nanoflakes Functionalized by Tungsten Trioxide Nanoparticles. J. Colloid Interface Sci. 2021, 595, 6–14. [Google Scholar] [CrossRef] [PubMed]
  107. Zhao, L.; Wang, K.; Wei, W.; Wang, L.; Han, W. High-performance Flexible Sensing Devices Based on Polyaniline/MXene Nanocomposites. InfoMat 2019, 1, 407–416. [Google Scholar] [CrossRef]
  108. Sun, S.; Wang, M.; Chang, X.; Jiang, Y.; Zhang, D.; Wang, D.; Zhang, Y.; Lei, Y. W18O49/Ti3C2Tx Mxene Nanocomposites for Highly Sensitive Acetone Gas Sensor with Low Detection Limit. Sens. Actuators B Chem. 2020, 304, 127274. [Google Scholar] [CrossRef]
  109. Jin, L.; Wu, C.; Wei, K.; He, L.; Gao, H.; Zhang, H.; Zhang, K.; Asiri, A.M.; Alamry, K.A.; Yang, L. Polymeric Ti3C2T x MXene Composites for Room Temperature Ammonia Sensing. ACS Appl. Nano Mater. 2020, 3, 12071–12079. [Google Scholar] [CrossRef]
  110. Liu, M.; Wang, Z.; Song, P.; Yang, Z.; Wang, Q. In2O3 Nanocubes/Ti3C2Tx MXene Composites for Enhanced Methanol Gas Sensing Properties at Room Temperature. Ceram. Int. 2021, 47, 23028–23037. [Google Scholar] [CrossRef]
  111. Chen, T.; Yan, W.; Wang, Y.; Li, J.; Hu, H.; Ho, D. SnS2/MXene Derived TiO2 Hybrid for Ultra-Fast Room Temperature NO2 Gas Sensing. J. Mater. Chem. C 2021, 9, 7407–7416. [Google Scholar] [CrossRef]
  112. Chen, W.Y.; Jiang, X.; Lai, S.-N.; Peroulis, D.; Stanciu, L. Nanohybrids of a MXene and Transition Metal Dichalcogenide for Selective Detection of Volatile Organic Compounds. Nat. Commun. 2020, 11, 1302. [Google Scholar] [CrossRef] [PubMed]
  113. Lee, S.H.; Eom, W.; Shin, H.; Ambade, R.B.; Bang, J.H.; Kim, H.W.; Han, T.H. Room-Temperature, Highly Durable Ti3C2Tx MXene/Graphene Hybrid Fibers for NH3 Gas Sensing. ACS Appl. Mater. Interfaces 2020, 12, 10434–10442. [Google Scholar] [CrossRef] [PubMed]
  114. Sun, B.; Lv, H.; Liu, Z.; Wang, J.; Bai, X.; Zhang, Y.; Chen, J.; Kan, K.; Shi, K. Co3O4@PEI/Ti3C2Tx MXene Nanocomposites for a Highly Sensitive NOx Gas Sensor with a Low Detection Limit. J. Mater. Chem. A 2021, 9, 6335–6344. [Google Scholar] [CrossRef]
  115. Zou, S.; Gao, J.; Liu, L.; Lin, Z.; Fu, P.; Wang, S.; Chen, Z. Enhanced Gas Sensing Properties at Low Working Temperature of Iron Molybdate/MXene Composite. J. Alloys Compd. 2020, 817, 152785. [Google Scholar] [CrossRef]
  116. Gogotsi, Y.; Huang, Q. MXenes: Two-Dimensional Building Blocks for Future Materials and Devices. ACS Nano 2021, 15, 5775–5780. [Google Scholar] [CrossRef]
  117. Lim, K.R.G.; Shekhirev, M.; Wyatt, B.C.; Anasori, B.; Gogotsi, Y.; Seh, Z.W. Fundamentals of MXene Synthesis. Nat. Synth. 2022, 1, 601–614. [Google Scholar] [CrossRef]
  118. Lee, E.; VahidMohammadi, A.; Prorok, B.C.; Yoon, Y.S.; Beidaghi, M.; Kim, D.-J. Room Temperature Gas Sensing of Two-Dimensional Titanium Carbide (MXene). ACS Appl. Mater. Interfaces 2017, 9, 37184–37190. [Google Scholar] [CrossRef]
  119. Majhi, S.M.; Ali, A.; Greish, Y.E.; El-Maghraby, H.F.; Qamhieh, N.N.; Hajamohideen, A.R.; Mahmoud, S.T. Accordion-like-Ti3C2 MXene-Based Gas Sensors with Sub-Ppm Level Detection of Acetone at Room Temperature. ACS Appl. Electron. Mater. 2022, 4, 4094–4103. [Google Scholar] [CrossRef]
  120. Chen, W.Y.; Lai, S.-N.; Yen, C.-C.; Jiang, X.; Peroulis, D.; Stanciu, L.A. Surface Functionalization of Ti3C2Tx MXene with Highly Reliable Superhydrophobic Protection for Volatile Organic Compounds Sensing. ACS Nano 2020, 14, 11490–11501. [Google Scholar] [CrossRef]
  121. Shuvo, S.N.; Ulloa Gomez, A.M.; Mishra, A.; Chen, W.Y.; Dongare, A.M.; Stanciu, L.A. Sulfur-Doped Titanium Carbide MXenes for Room-Temperature Gas Sensing. ACS Sens. 2020, 5, 2915–2924. [Google Scholar] [CrossRef]
  122. Li, D.; Qu, M.; Zhang, Q.; Xie, J. Multivariable Sensor Based on Mxene and Machine Learning for Selective Detections of VOCs. In Proceedings of the 2021 IEEE 34th International Conference on Micro Electro Mechanical Systems (MEMS), Virtual, 25–29 January 2021; IEEE: Piscataway, NJ, USA, 2021; pp. 802–805. [Google Scholar]
  123. Lee, E.; VahidMohammadi, A.; Yoon, Y.S.; Beidaghi, M.; Kim, D.-J. Two-Dimensional Vanadium Carbide MXene for Gas Sensors with Ultrahigh Sensitivity toward Nonpolar Gases. ACS Sens. 2019, 4, 1603–1611. [Google Scholar] [CrossRef] [PubMed]
  124. Guo, W.; Surya, S.G.; Babar, V.; Ming, F.; Sharma, S.; Alshareef, H.N.; Schwingenschlogl, U.; Salama, K.N. Selective Toluene Detection with Mo2CTx MXene at Room Temperature. ACS Appl. Mater. Interfaces 2020, 12, 57218–57227. [Google Scholar] [CrossRef] [PubMed]
  125. Vasseghian, Y.; Doan, V.D.; Nguyen, T.T.T.; Vo, T.-T.T.; Do, H.H.; Vu, K.B.; Vu, Q.H.; Dai Lam, T.; Tran, V.A. Flexible and High-Sensitivity Sensor Based on Ti3C2–MoS2 MXene Composite for the Detection of Toxic Gases. Chemosphere 2022, 291, 133025. [Google Scholar]
  126. Zhu, L.; Zeng, W. Room-Temperature Gas Sensing of ZnO-Based Gas Sensor: A Review. Sens. Actuators A Phys. 2017, 267, 242–261. [Google Scholar] [CrossRef]
  127. Yang, D.; Fan, X.; Zhao, D.; An, Y.; Hu, Y.; Luo, Z. Sc2CO2 and Mn-Doped Sc2CO2 as Gas Sensor Materials to NO and CO: A First-Principles Study. Phys. E Low-Dimens. Syst. Nanostruct. 2019, 111, 84–90. [Google Scholar] [CrossRef]
  128. Sun, L.; Sun, N.; Liu, Y.; Jiang, C. Anisotropic Frictional Properties between Ti3C2Tx MXene/SiO2 Layer-Dependent Heterojunctions. J. Sci. Adv. Mater. Devices 2021, 6, 488–493. [Google Scholar] [CrossRef]
  129. Zhao, Q.-N.; Zhang, Y.-J.; Duan, Z.-H.; Wang, S.; Liu, C.; Jiang, Y.-D.; Tai, H.-L. A Review on Ti3C2Tx-Based Nanomaterials: Synthesis and Applications in Gas and Humidity Sensors. Rare Met. 2021, 40, 1459–1476. [Google Scholar] [CrossRef]
  130. Devaraj, M.; Rajendran, S.; Hoang, T.K.A.; Soto-Moscoso, M. A Review on MXene and Its Nanocomposites for the Detection of Toxic Inorganic Gases. Chemosphere 2022, 302, 134933. [Google Scholar] [CrossRef]
  131. Ta, Q.T.H.; Thakur, D.; Noh, J.-S. Enhanced Gas Sensing Performance of ZnO/Ti3C2Tx MXene Nanocomposite. Micromachines 2022, 13, 1710. [Google Scholar] [CrossRef]
  132. Pazniak, H.; Plugin, I.A.; Loes, M.J.; Inerbaev, T.M.; Burmistrov, I.N.; Gorshenkov, M.; Polcak, J.; Varezhnikov, A.S.; Sommer, M.; Kuznetsov, D. V Partially Oxidized Ti3C2Tx MXenes for Fast and Selective Detection of Organic Vapors at Part-per-Million Concentrations. ACS Appl. Nano Mater. 2020, 3, 3195–3204. [Google Scholar] [CrossRef]
  133. Hou, M.; Guo, S.; Yang, L.; Gao, J.; Hu, T.; Wang, X.; Li, Y. Improvement of Gas Sensing Property for Two-Dimensional Ti3C2Tx Treated with Oxygen Plasma by Microwave Energy Excitation. Ceram. Int. 2021, 47, 7728–7737. [Google Scholar] [CrossRef]
  134. Wang, D.; Zhang, D.; Yang, Y.; Mi, Q.; Zhang, J.; Yu, L. Multifunctional Latex/Polytetrafluoroethylene-Based Triboelectric Nanogenerator for Self-Powered Organ-like MXene/Metal–Organic Framework-Derived CuO Nanohybrid Ammonia Sensor. ACS Nano 2021, 15, 2911–2919. [Google Scholar] [CrossRef] [PubMed]
  135. Zheng, Q.; Shi, B.; Li, Z.; Wang, Z.L. Recent Progress on Piezoelectric and Triboelectric Energy Harvesters in Biomedical Systems. Adv. Sci. 2017, 4, 1700029. [Google Scholar] [CrossRef]
  136. Ahmed, A.; Hassan, I.; El-Kady, M.F.; Radhi, A.; Jeong, C.K.; Selvaganapathy, P.R.; Zu, J.; Ren, S.; Wang, Q.; Kaner, R.B. Integrated Triboelectric Nanogenerators in the Era of the Internet of Things. Adv. Sci. 2019, 6, 1802230. [Google Scholar] [CrossRef]
  137. Yang, Y.; Lin, L.; Zhang, Y.; Jing, Q.; Hou, T.-C.; Wang, Z.L. Self-Powered Magnetic Sensor Based on a Triboelectric Nanogenerator. ACS Nano 2012, 6, 10378–10383. [Google Scholar] [CrossRef]
  138. Zhou, Y.; Deng, W.; Xu, J.; Chen, J. Engineering Materials at the Nanoscale for Triboelectric Nanogenerators. Cell Reports Phys. Sci. 2020, 1, 100142. [Google Scholar] [CrossRef]
  139. Dong, Y.; Mallineni, S.S.K.; Maleski, K.; Behlow, H.; Mochalin, V.N.; Rao, A.M.; Gogotsi, Y.; Podila, R. Metallic MXenes: A New Family of Materials for Flexible Triboelectric Nanogenerators. Nano Energy 2018, 44, 103–110. [Google Scholar] [CrossRef]
  140. Zhang, S.; Rana, S.M.S.; Bhatta, T.; Pradhan, G.B.; Sharma, S.; Song, H.; Jeong, S.; Park, J.Y. 3D Printed Smart Glove with Pyramidal MXene/Ecoflex Composite-Based Toroidal Triboelectric Nanogenerators for Wearable Human-Machine Interaction Applications. Nano Energy 2023, 106, 108110. [Google Scholar] [CrossRef]
  141. Chang, L.; Xiao, X. The Review of MXenes for Osmotic Energy Harvesting: Synthesis and Properties. Diam. Relat. Mater. 2023, 136, 109971. [Google Scholar] [CrossRef]
  142. Luo, X.; Zhu, L.; Wang, Y.; Li, J.; Nie, J.; Wang, Z.L. A Flexible Multifunctional Triboelectric Nanogenerator Based on MXene/PVA Hydrogel. Adv. Funct. Mater. 2021, 31, 2104928. [Google Scholar] [CrossRef]
  143. Cao, W.; Ouyang, H.; Xin, W.; Chao, S.; Ma, C.; Li, Z.; Chen, F.; Ma, M. A Stretchable Highoutput Triboelectric Nanogenerator Improved by MXene Liquid Electrode with High Electronegativity. Adv. Funct. Mater. 2020, 30, 2004181. [Google Scholar] [CrossRef]
  144. Kim, K.N.; Kim, S.Y.; Choi, S.H.; Lee, M.; Song, W.; Lim, J.; Lee, S.S.; Myung, S. All-Printed Wearable Triboelectric Nanogenerator with Ultra-Charged Electron Accumulation Polymers Based on MXene Nanoflakes. Adv. Electron. Mater. 2022, 8, 2200819. [Google Scholar] [CrossRef]
  145. Rana, S.M.S.; Rahman, M.T.; Salauddin, M.; Sharma, S.; Maharjan, P.; Bhatta, T.; Cho, H.; Park, C.; Park, J.Y. Electrospun PVDF-TrFE/MXene Nanofiber Mat-Based Triboelectric Nanogenerator for Smart Home Appliances. ACS Appl. Mater. Interfaces 2021, 13, 4955–4967. [Google Scholar] [CrossRef] [PubMed]
  146. Tan, D.; Sun, N.; Chen, L.; Bu, J.; Jiang, C. Piezoelectricity in Monolayer and Multilayer Ti3C2Tx MXenes: Implications for Piezoelectric Devices. ACS Appl. Nano Mater. 2021, 5, 1034–1046. [Google Scholar] [CrossRef]
  147. Cao, V.A.; Kim, M.; Lee, S.; Kim, C.G.; Cao Van, P.; Thi, T.N.; Jeong, J.-R.; Nah, J. Enhanced Output Performance of a Flexible Piezoelectric Nanogenerator Realized by Lithium-Doped Zinc Oxide Nanowires Decorated on MXene. ACS Appl. Mater. Interfaces 2022, 14, 26824–26832. [Google Scholar] [CrossRef] [PubMed]
  148. Han, C.; Zhang, H.; Chen, Q.; Li, T.; Kong, L.; Zhao, H.; He, L. A Directional Piezoelectric Sensor Based on Anisotropic PVDF/MXene Hybrid Foam Enabled by Unidirectional Freezing. Chem. Eng. J. 2022, 450, 138280. [Google Scholar] [CrossRef]
  149. Auliya, R.Z.; Ooi, P.C.; Sadri, R.; Talik, N.A.; Yau, Z.Y.; Mohammad Haniff, M.A.S.; Goh, B.T.; Dee, C.F.; Aslfattahi, N.; Al-Bati, S. Exploration of 2D Ti3C2 MXene for All Solution Processed Piezoelectric Nanogenerator Applications. Sci. Rep. 2021, 11, 17432. [Google Scholar] [CrossRef]
  150. Chang, W.-L.; Sun, Z.-Q.; Zhang, Z.-M.; Wei, X.-P.; Tao, X. Thermoelectric Properties of Two-Dimensional Double Transition Metal MXenes: ScYCT2 (T = F, OH). J. Phys. Chem. Solids 2023, 176, 111210. [Google Scholar] [CrossRef]
  151. Bark, H.; Thangavel, G.; Liu, R.J.; Chua, D.H.C.; Lee, P.S. Effective Surface Modification of 2D MXene toward Thermal Energy Conversion and Management. Small Methods 2023, 7, 2300077. [Google Scholar] [CrossRef]
  152. Fei, J.; Koh, S.W.; Tu, W.; Ge, J.; Rezaeyan, H.; Hou, S.; Duan, H.; Lam, Y.C.; Li, H. Functionalized MXene Enabled Sustainable Water Harvesting and Desalination. Adv. Sustain. Syst. 2020, 4, 2000102. [Google Scholar] [CrossRef]
  153. Zhao, K.; Lee, J.W.; Yu, Z.G.; Jiang, W.; Oh, J.W.; Kim, G.; Han, H.; Kim, Y.; Lee, K.; Lee, S. Humidity-Tolerant Moisture-Driven Energy Generator with MXene Aerogel–Organohydrogel Bilayer. ACS Nano 2023, 17, 5472–5485. [Google Scholar] [CrossRef] [PubMed]
  154. Bae, J.; Kim, M.S.; Oh, T.; Suh, B.L.; Yun, T.G.; Lee, S.; Hur, K.; Gogotsi, Y.; Koo, C.M.; Kim, I.-D. Towards Watt-Scale Hydroelectric Energy Harvesting by Ti3C2Tx-Based Transpiration-Driven Electrokinetic Power Generators. Energy Environ. Sci. 2022, 15, 123–135. [Google Scholar] [CrossRef]
  155. Li, P.; Su, N.; Wang, Z.; Qiu, J. A Ti3C2Tx MXene-Based Energy-Harvesting Soft Actuator with Self-Powered Humidity Sensing and Real-Time Motion Tracking Capability. ACS Nano 2021, 15, 16811–16818. [Google Scholar] [CrossRef] [PubMed]
  156. Kim, J.Y.; Lee, J.-W.; Jung, H.S.; Shin, H.; Park, N.-G. High-Efficiency Perovskite Solar Cells. Chem. Rev. 2020, 120, 7867–7918. [Google Scholar] [CrossRef] [PubMed]
  157. Guo, Z.; Gao, L.; Xu, Z.; Teo, S.; Zhang, C.; Kamata, Y.; Hayase, S.; Ma, T. High Electrical Conductivity 2D MXene Serves as Additive of Perovskite for Efficient Solar Cells. Small 2018, 14, 1802738. [Google Scholar] [CrossRef]
  158. Zheng, H.; Wang, Y.; Niu, B.; Ge, R.; Lei, Y.; Yan, L.; Si, J.; Zhong, P.; Ma, X. Controlling the Defect Density of Perovskite Films by MXene/SnO2 Hybrid Electron Transport Layers for Efficient and Stable Photovoltaics. J. Phys. Chem. C 2021, 125, 15210–15222. [Google Scholar] [CrossRef]
  159. Wang, Y.; Xiang, P.; Ren, A.; Lai, H.; Zhang, Z.; Xuan, Z.; Wan, Z.; Zhang, J.; Hao, X.; Wu, L. MXene-Modulated Electrode/SnO2 Interface Boosting Charge Transport in Perovskite Solar Cells. ACS Appl. Mater. Interfaces 2020, 12, 53973–53983. [Google Scholar] [CrossRef]
  160. Shin, E.-Y.; Choi, S.B.; Lee, J.H.; Yoo, B.; Han, C.J.; Park, S.H.; Noh, J.H.; Kim, J.-W.; Son, H.J. An Inverted Layer-by-Layer Process to Enable Ultrasmooth MXene–Ag Nanowire Hybrid Electrode for Organic Photovoltaics. Sol. RRL 2023, 7, 2201130. [Google Scholar] [CrossRef]
  161. Chen, Y.; Wang, D.; Lin, Y.; Zou, X.; Xie, T. In Suit Growth of CuSe Nanoparticles on MXene (Ti3C2) Nanosheets as an Efficient Counter Electrode for Quantum Dot-Sensitized Solar Cells. Electrochim. Acta 2019, 316, 248–256. [Google Scholar] [CrossRef]
  162. Aryal, U.K.; Pazniak, H.; Kumari, T.; Weber, M.; Johansson, F.O.L.; Vannucchi, N.; Witkowski, N.; Turkovic, V.; Di Carlo, A.; Madsen, M. 2D MXene-Based Electron Transport Layers for Nonhalogenated Solvent-Processed Stable Organic Solar Cells. ACS Appl. Energy Mater. 2023, 6, 4549–4558. [Google Scholar] [CrossRef]
Figure 1. Schematic of Ti3C2Tx MXene preparation from the Ti3AlC2 MAX phase. Reprinted and adapted with permission from [11]. Copyright 2021, Elsevier.
Figure 1. Schematic of Ti3C2Tx MXene preparation from the Ti3AlC2 MAX phase. Reprinted and adapted with permission from [11]. Copyright 2021, Elsevier.
Chemosensors 11 00477 g001
Figure 2. (AE) Halogen etching procedure of the MAX phase. Tetrahydrofuran was used as the stable colloidal solution. Reprinted and adapted with permission from [64]. Copyright 2021 American Chemical Society.
Figure 2. (AE) Halogen etching procedure of the MAX phase. Tetrahydrofuran was used as the stable colloidal solution. Reprinted and adapted with permission from [64]. Copyright 2021 American Chemical Society.
Chemosensors 11 00477 g002
Figure 3. Schematic of an increase in the d-spacing of Ti3C2Tx MXene with intercalation using cations. Reprinted and adapted with permission from [70].
Figure 3. Schematic of an increase in the d-spacing of Ti3C2Tx MXene with intercalation using cations. Reprinted and adapted with permission from [70].
Chemosensors 11 00477 g003
Figure 4. (ac) XRD results of Ti3C2Tx MXene prepared with different concentrations of HF solution. Reprinted and adapted with permission from [5]. Copyright 2017 American Chemical Society.
Figure 4. (ac) XRD results of Ti3C2Tx MXene prepared with different concentrations of HF solution. Reprinted and adapted with permission from [5]. Copyright 2017 American Chemical Society.
Chemosensors 11 00477 g004
Figure 5. (a,c,d) Schematic of the Ti3C2Tx gas-sensing material and (b) its sensing signals of various toxic gases of 100 ppm at room temperature. Reprinted and adapted with permission from [118]. Copyright 2017 American Chemical Society.
Figure 5. (a,c,d) Schematic of the Ti3C2Tx gas-sensing material and (b) its sensing signals of various toxic gases of 100 ppm at room temperature. Reprinted and adapted with permission from [118]. Copyright 2017 American Chemical Society.
Chemosensors 11 00477 g005
Figure 6. Thickness-dependent responses of Ti3C2Tx MXene to various gases at a fixed concentration of 100 ppm. Reprinted and adapted with permission from [32].
Figure 6. Thickness-dependent responses of Ti3C2Tx MXene to various gases at a fixed concentration of 100 ppm. Reprinted and adapted with permission from [32].
Chemosensors 11 00477 g006
Figure 7. (a,b) Side and top views of the optimized structure of eight gas molecules adsorbed on Ti3C2Tx MXene. Reprinted and adapted with permission from [22]. Copyright 2019 American Chemical Society.
Figure 7. (a,b) Side and top views of the optimized structure of eight gas molecules adsorbed on Ti3C2Tx MXene. Reprinted and adapted with permission from [22]. Copyright 2019 American Chemical Society.
Chemosensors 11 00477 g007
Figure 8. (af) Gas response variation of monolayer V2CTx MXene flakes to different types of gases. Reprinted and adapted with permission from [123]. Copyright 2019 American Chemical Society.
Figure 8. (af) Gas response variation of monolayer V2CTx MXene flakes to different types of gases. Reprinted and adapted with permission from [123]. Copyright 2019 American Chemical Society.
Chemosensors 11 00477 g008
Figure 9. (a,b) Comparison of the response behaviors of MoS2/Ti3C2Tx and WO3/Ti3C2Tx composites to NO2 gas. (c) NO2-sensing mechanism of MoS2/Ti3C2Tx composite. Reprinted and adapted with permission from [125].
Figure 9. (a,b) Comparison of the response behaviors of MoS2/Ti3C2Tx and WO3/Ti3C2Tx composites to NO2 gas. (c) NO2-sensing mechanism of MoS2/Ti3C2Tx composite. Reprinted and adapted with permission from [125].
Chemosensors 11 00477 g009
Figure 10. Gas-sensing signal of the as-prepared PEDOT:PSS/Ti3C2Tx composite to 100 ppm NH3 gas at a variety of bending angles (60°, 120°, 180°, 240°). Reprinted and adapted with permission from [109]. Copyright 2020 American Chemical Society.
Figure 10. Gas-sensing signal of the as-prepared PEDOT:PSS/Ti3C2Tx composite to 100 ppm NH3 gas at a variety of bending angles (60°, 120°, 180°, 240°). Reprinted and adapted with permission from [109]. Copyright 2020 American Chemical Society.
Chemosensors 11 00477 g010
Figure 11. VOC-sensing activity of oxidized TiO2/Ti3C2Tx composite. Reprinted and adapted with permission from [132]. Copyright 2020 American Chemical Society.
Figure 11. VOC-sensing activity of oxidized TiO2/Ti3C2Tx composite. Reprinted and adapted with permission from [132]. Copyright 2020 American Chemical Society.
Chemosensors 11 00477 g011
Figure 12. Schematic and experimental results of a self-powered ammonia sensor driven by TENG. Reprinted and adapted with permission from [134]. Copyright 2021 American Chemical Society.
Figure 12. Schematic and experimental results of a self-powered ammonia sensor driven by TENG. Reprinted and adapted with permission from [134]. Copyright 2021 American Chemical Society.
Chemosensors 11 00477 g012
Figure 13. (a) Structure of Ti3C2Tx MXene and MXene triboelectric nanogenerator (MXene TENG) assembled as shown in the schematics. (b) MXene/PVA hydrogel flexible multifunctional TENG. (c) Mechanical stability and reliability of EAPs-based TENG with diverse applications. (d) The 50 LEDs, calculator, and stopwatch were powered by the TENG. Reprinted and adapted with permission from [139,142,143,144].
Figure 13. (a) Structure of Ti3C2Tx MXene and MXene triboelectric nanogenerator (MXene TENG) assembled as shown in the schematics. (b) MXene/PVA hydrogel flexible multifunctional TENG. (c) Mechanical stability and reliability of EAPs-based TENG with diverse applications. (d) The 50 LEDs, calculator, and stopwatch were powered by the TENG. Reprinted and adapted with permission from [139,142,143,144].
Chemosensors 11 00477 g013
Figure 14. (af) Output performance of PENG fabricated with Li:ZnO/Ti3C2Tx composites. Reprinted and adapted with permission from [147]. Copyright 2022 American Chemical Society.
Figure 14. (af) Output performance of PENG fabricated with Li:ZnO/Ti3C2Tx composites. Reprinted and adapted with permission from [147]. Copyright 2022 American Chemical Society.
Chemosensors 11 00477 g014
Figure 15. MXene for harvesting energy with moisture stimuli. Reprinted and adapted with permission from [155]. Copyright 2021 American Chemical Society.
Figure 15. MXene for harvesting energy with moisture stimuli. Reprinted and adapted with permission from [155]. Copyright 2021 American Chemical Society.
Chemosensors 11 00477 g015
Figure 16. (a) Perovskite solar cell, (b) Interface boosting charge transport. Reprinted and adapted with permission from [156,159]. Copyright 2020 American Chemical Society.
Figure 16. (a) Perovskite solar cell, (b) Interface boosting charge transport. Reprinted and adapted with permission from [156,159]. Copyright 2020 American Chemical Society.
Chemosensors 11 00477 g016
Table 1. Gas-sensing activities of pure Ti3C2Tx MXenes and their composites.
Table 1. Gas-sensing activities of pure Ti3C2Tx MXenes and their composites.
MaterialsGasTemperature (°C)Concentration (ppm)Response (%)tRes/tRec (s)Reference
Ti3C2TxAcetoneRT53.791/103[98]
Ti3C2TxNH3RT10028.881/201[99]
Ti3C2Tx/CuOToluene2505011.5270/10[100]
Ti3C2Tx@TiO2@MoS2NO2RT5055.11.8/70[101]
Ti3C2Tx/SnO2NH3RT504136/45[102]
Ti3C2Tx/MoS2NO2RT1025.1530/140[103]
Ti3C2Tx/MoS2NO2RT2040.1525/155[103]
Ti3C2Tx/Co3O4HCHORT109.384/6[104]
Ti3C2Tx/ZnONO2RT2022.334/103[105]
Ti3C2Tx/WO3NH3RT122.4120/229[106]
Ti3C2Tx/PANIEthanolRT200420.4/0.5[107]
Ti3C2Tx/W18O49Acetone3000.181.45.7/7[108]
Ti3C2Tx/PEDOT:PSSNH3RT10037117/40[109]
Ti3C2Tx/In2O3MethaneRT5306.6/4[110]
Ti3C2Tx/SnS2NO2RT100011665/10[111]
Ti3C2Tx/WSe2EthanolRT402510/6[112]
Ti3C2Tx/GONH3RT1008-/-[113]
Ti3C2Tx/Co3O4/PEINOxRT100302/74[114]
Ti3C2Tx/Fe2(MnO4)3N-butaneRT1004419/25[115]
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

Ta, Q.T.H.; Thakur, D.; Noh, J.-S. Design of Functional Ti3C2Tx MXene for Gas Sensors and Energy Harvesting: A Review. Chemosensors 2023, 11, 477. https://doi.org/10.3390/chemosensors11090477

AMA Style

Ta QTH, Thakur D, Noh J-S. Design of Functional Ti3C2Tx MXene for Gas Sensors and Energy Harvesting: A Review. Chemosensors. 2023; 11(9):477. https://doi.org/10.3390/chemosensors11090477

Chicago/Turabian Style

Ta, Qui Thanh Hoai, Deepika Thakur, and Jin-Seo Noh. 2023. "Design of Functional Ti3C2Tx MXene for Gas Sensors and Energy Harvesting: A Review" Chemosensors 11, no. 9: 477. https://doi.org/10.3390/chemosensors11090477

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

Article Metrics

Back to TopTop