Next Article in Journal
A Study of the Optical and Structural Properties of SnO2 Nanoparticles Synthesized with Tilia cordata Applied in Methylene Blue Degradation
Previous Article in Journal
A Transcriptome- and Interactome-Based Analysis Identifies Repurposable Drugs for Human Breast Cancer Subtypes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Surface Terminations of MXene: Synthesis, Characterization, and Properties

1
Key Laboratory of Functional Materials Physics and Chemistry of the Ministry of Education, The Joint Laboratory of MXene Materials, Jilin Normal University, Changchun 130103, China
2
Key Laboratory of Automobile Materials MOE, School of Materials Science & Engineering, Jilin Provincial International Cooperation Key Laboratory of High-Efficiency Clean Energy Materials, Electron Microscopy Center, and International Center of Future Science, Jilin University, Changchun 130012, China
3
School of Optoelectronic Science, Changchun College of Electronic Technology, Changchun 130114, China
*
Authors to whom correspondence should be addressed.
Symmetry 2022, 14(11), 2232; https://doi.org/10.3390/sym14112232
Submission received: 7 September 2022 / Revised: 13 October 2022 / Accepted: 17 October 2022 / Published: 24 October 2022
(This article belongs to the Section Chemistry: Symmetry/Asymmetry)

Abstract

:
MXene, 2D transition metal carbides, nitrides, and carbonitrides with a unique 2D structure, inspired a series of function applications related to energy storage and conversion, biometrics and sensing, lighting, purification, and separation. Its surface terminations are confined by the adjacent MXene layers, and form the 2D planar space with symmetrical surfaces, which is similar to a 2D nanoreactor that can be utilized and determined MXene’s function. Based on the working principle, surface and interface play critical roles in the ion intercalation, physical/chemical adsorption, and chemical reaction process, and show significant effects on MXene’s properties and functions. Although there have been some reviews on MXene, less attention has been paid to the underlying principle of the involved surface chemistry, controllable design, and resultant properties. Herein, the regulation methods, characterization techniques, and the effects on properties of MXene surface terminations were summarized to understand the surface effects, and the relationship between the terminations and properties. We expected this review can offer the route for a series of ongoing studies to address the MXene surface environment and the guidelines for MXene’s application.

1. Introduction

Since the first 2D layered Ti3C2Tx in 2011, transition metal carbides, nitrides, or carbonitrides (MXene), as a family of 2D layered structure materials, have rigorously emerged [1]. It is well-established that the ternary carbides and nitrides with a Mn+1AXn formula, M is an early transition metal, A is an A-group (mostly groups 13 and 14) element, and X is C and/or N, forming laminated structures with anisotropic properties. These, the so-called MAX phases are layered hexagonal (space group P63/mmc), with two formula units per unit cell. The near-close-packed M-layers are interleaved with pure A-group element layers, with the X-atoms filling the octahedral sites between the former. Over 60 MAX phases are currently known to exist. Upon the development of the initial synthesis methods and the new synthetic precursor of MAX, a variety of stable phases MXene (more than 200 kinds, mainly Ti3C2Tx, Ti2CTx, Nb2CTx, Mo2CTx, Ti4N3Tx, Ta4C3Tx, Cr2TiC2Tx, V2CTx, Zr3C2Tx, (Nb0.8Zr0.2)4C3Tx and so on) have been determined and predicted [2], as shown in Figure 1. Owing to the unusual electronic [3], mechanical [4] and optical properties, numerous functions were explored in batteries [5], supercapacitors [6], photocatalysts [7], catalysts [8], transparent conducting films [9], electromagnetic interference shielding [10], sensors [11], adsorption agents [12], and flexible high-strength composites [13].
However, due to the presence of a large number of exposed metal atoms on the surface of MXene, the structure of MXene is easily destroyed by oxidation [17]. MXene is also prone to irreversible lamellar stacking in preparation and application [18]. Mechanical strength is insufficient [19], poor solubility in non-polar polymer or weakly polar polymer [20], low energy density, poor multiplier performance [21], and poor environmental stability [22].
MXene surface terminations have a significant influence on the stability, electronic structure, and other physical and chemical properties of MXene [23]. The controllable regulation of MXene surface terminations is an effective prerequisite to obtaining MXene materials with differential structural characteristics and properties. At present, the regulation of MXene surface terminations mainly starts from material preparation and post-treatment, and then selects different types or proportions of functional groups according to different application scenarios. For example, the lithium-ion storage capacity of MXene is largely determined by the properties of MXene surface terminations [23]. MXene surface terminations can be modified to remove -F and -OH surface groups that hinder the transport of Li ions, to obtain electrode materials with high lithium storage capacity and high-rate performance. In addition, by modifying MXene surface terminations Tx (-OH, -F, and -O) to adjust the carrier transport behavior of Ti2CTx MXene with high mobility and appropriate bandgap can expand the application range of MXene [8,9,10,11]. Moreover, high surface electrical loading of MXene, the physical modification of organic groups, or the loading functional agents on the surface of MXene have proved that the appropriate surface modification and functionalization can improve the performance of MXene in nanomedicine and broaden its potential biomedical application [24]. Moreover, the surface functional groups can affect the interlayer distance, energy band structure, and work function of MXene, which determine several properties and functions. The targeted surface modification and functionalization have proved that it can improve the performances of MXene in nanomedicine, electrocatalysis, thermodynamics, electromagnetism, biology, tribological and fluid mechanics. For example, the -O functional group is HER active site, and the high -F surface coverage is more conductive. Therefore, adjusting the type and proportion of functional groups is a strategy to improve the performance of electrocatalysis.
MXene surface terminations play an important role in interface-related applications such as thermodynamics, electromagnetism, biology, tribological and fluid mechanics. The different surface terminations of MXene own the preparation method and the experimental requirement, which need to be strictly controlled. Moreover, the successful preparation of MXene with different surface terminations requires a large number of comparison tests and characterization, and there may be differences in physical and chemical properties due to the uncertainty of the number of functional groups. Therefore, the key point of using MXene surface terminations regulation is to develop its regulatory methods, characterization techniques, and understand the structure-activity relationship between MXene terminations and properties and performance.
This review summarizes the synthesis route and characterization methods of MXene with differentiated terminations, as well as the effects of different terminations on the mechanical, optical, and magnetic properties of MXene. In addition, the role of MXene termination for its application function is also discussed. This is summarized in Figure 2. At the same time, the key research of MXene surface termination is prospected.

2. Surface Termination Regulation

MXene prepared by fluoride-containing etching method can obtain these surface terminations of -O, -F, and -OH. Owing to the preparation method, the types of MXene surface terminations are very simple, which limits the research on MXene surface terminations and the interaction between MXene surface terminations and properties [25,26]. Therefore, the key to expanding MXene surface termination is to develop the regulation method of MXene surface termination and new MXene preparation technology. With the development of the fluorine-free etching method and Lewis acid molten salt etching method [27], Ti3C2 MXene with single -Cl, -Br, -I, -O, and -S as terminations were developed. Ti3C2 MXene with binary (-ClBr, -ClI, and -BrI), and ternary (-ClBrI) halogen terminations were also synthesized through the Lewis-acidic-melt etching route [28]. In addition, the surface termination of MXene can be eliminated, replaced, and esterified by hydrothermal and solvothermal methods. Ti3C2, Ti2C, and Nb2C MXene with the surface terminations of -Se, -Te, -NH, and phosphorous and ester were developed [29,30]. As shown in Figure 3, the most common processes for preparing MXene are summarized. The development of the preparation technology of MXene terminations provides material science basis for the research on MXene surface terminations, and also provides more technical routes for the performance regulation of MXene as functional materials.

2.1. HF/Fluoride Etching

In 2011, as a sign of MXene materials emergence, Yury Gogotsi and Michel W. Barsoum reported the pioneering work of preparing multilayer Ti3C2F2 and Ti3C2(OH)2 by etching Ti3AlC2 MAX phase ceramic materials with HF [31]. This is also one of the most used methods to synthesize MXene. As shown in Figure 4a, the HF etching reaction and surface functionalization mechanism are described as follows [25,32]:
M n + 1 AX n + 3 HF AF 3 + 1.5 H 2 + M n + 1 X n
M n + 1 X n + 2 H 2 O M n + 1 X n OH 2 + H 2
M n + 1 X n + 2 HF M n + 1 X n F 2 + H 2
In Equation (1), the M-A in the MAX phase ceramics is a metal bond, which is weaker than the M-X bond. Due to the difference in the binding energy, HF reacts with the A-layer elements in the MAX phase to generate AF3, H2, and the corresponding Mn+1Xn phase [24]. However, Mn+1Xn, which has lost elements in the A-layer, reacts with the surrounding water or HF to form Mn+1Xn(OH)2 with OH or/and F functional groups, corresponding to Equations (2) and (3). Figure 4b shows the representative steps of Ti3C2Tx MXene synthesis from Ti3AlC2, and the structure after surface functional group modification [33].
Figure 4. (a) Schematic diagram of the synthesis of MXene by HF etching [24]. (b) The synthesis and structure diagram of Ti3C2Tx MXene [25].
Figure 4. (a) Schematic diagram of the synthesis of MXene by HF etching [24]. (b) The synthesis and structure diagram of Ti3C2Tx MXene [25].
Symmetry 14 02232 g004
Another reaction method is to use acid/fluoride salt solution etching MAX phase ceramics, such as HCl/LiF [34,35,36], HCl/NH4F, HCl/FeF3 [37,38] and H2SO4/LiF [39,40]. As shown in Figure 5, Lipatov et al. added LiF powder to concentrated HCl and stirred it to dissolve completely. Subsequently, Ti3AlC2 MAX powder was slowly added to the LiF/HCl dispersion, and the etching process was carried out in a water bath at 35 °C for 24 h under magnetic stirring. The resulting mixture was washed repeatedly with deionized water until a dark supernatant with a pH of approximately 6. After mild ultrasonic treatment, the mixture was centrifuged, and the supernatant was collected and freeze-dried to obtain MXene powder with -O, -OH, and -F functional groups [41].
Although the direct etching of precursor MAX phase by HF is the most widely used method to prepare MXene, it is potentially dangerous to humans and will produce a large amount of waste acid and liquid, which is not conducive to environmental protection. Therefore, it is necessary to develop a more moderate preparation method. Moreover, in an aqueous liquid environment, it is difficult for M of MXene to avoid contact with -OH and -F in a liquid environment during the reaction process, so the surface terminations formed by the discussed above are mostly -O, -OH, and -F. The ratio of these three functional groups is difficult to control. Due to the competition of metal ions, the content of termination -F on the surface of MXene via fluoride salt is relatively less. In addition, the fresher the sample is, the higher the -OH content is, with the increase of storage time, some -OH transform into -O.

2.2. Molten Salt Etching Method

To date, tremendous efforts have been made to develop synthetic methods of MXene. Based on the principle that transition metal halides are electron acceptors and can undergo displacement reaction with atomic layer A of MAX phase in the molten state, MXene materials can be prepared by etching MAX phase with transition metal halides in a molten state. As shown in Figure 6, taking Ti3AlC2 MAX as an example, the transition element atom (Zn) occupies the A site in the preexisting MAX phase structure. A new MAX phase Ti3ZnC2 was synthesized by the substitution reaction between the Zn element in molten ZnCl2 and the Al element in the precursor of MAX phase (Ti3AlC2). The formation mechanism can be expressed as the following simplified reaction:
Ti 3 AlC 2 + 1.5 ZnCl 2 Ti 3 ZnC 2 + 0.5 Zn + AlCl 3
Ti 3 AlC 2 + 1.5 ZnCl 2 Ti 3 C 2 + 1.5 Zn + AlCl 3
Ti 3 C 2 + Zn Ti 3 ZnC 2
It is well known that the melting point of ZnCl2 is about 280 °C, and in the molten state, it ionizes into Zn2+ and ZnCl42− tetrahedra [42,43]. Coordination unsaturated Zn2+ is a strong acceptor of Cl and electrons, which act as Lewis acidic in ZnCl2 molten salts. In this acidic environment, the weakly bonded Al atoms in Ti3AlC2 can be easily converted to Al3+ by a redox reaction (reaction 5). The resulting Al3+ will further combine with Cl to form AlCl3, which has a boiling point of about 180 °C and is expected to evaporate rapidly at the reaction temperature (550 °C). At the same time, the in-situ reduced zinc atoms intercalated into the Ti3C2 MXene layer and filled the A site of the MAX phase, thus forming Ti3ZnC2. The evaporation of AlCl3 provides the driving force for the outward diffusion of Al atoms, while the a-site vacancies of the MAX phase enable the insertion of Zn atoms into the Ti3C2 MXene layer. When excessive amounts of ZnCl2 are used, Ti3ZnC2 subsequently spall off to form -Cl terminating MXene (Ti3C2Cl2) due to the strong Lewis acidity of molten ZnCl2. Figure 7 shows the theoretical calculation of the synthesis mechanism of Ti3C2Cl2 by electrochemical corrosion assisted by molten salt.
Compared with HF etching method, the chemical process of forming eutectic in Al-Zn system is safer and cleaner [27]. This kind of MXene with -Cl as termination prepared by molten salt also provides an experimental and theoretical basis for the subsequent development of more functional groups of MXene. Based on obtaining -Cl termination MXene, it can also be changed into other functional groups by the molten salt method [44]. As shown in Figure 8, -Cl termination MXene was firstly prepared by LiCl-KCl salt, heating, and electrolysis, etc. Then Li2O or Li2S was added to the melting system, and -O or -S termination MXene was obtained after high-temperature cleaning. The synthesis mechanism is as follows:
Ti 3 AlC 2 + 3 Cl AlCl 3 + Ti 3 C 2 + 3 e
Ti 3 C 2 + 2 Cl Ti 3 C 2 Cl 2 + 2 e
Cathode : Li + K + + e Li K
Ti 3 AlC 2 + 5 LiCl KCl Ti 3 C 2 Cl 2 + 5 Li K + AlCl 3
This simple and environmentally friendly process to prepare MXene with controllable functional groups is extremely important. Since electrons are used as reactants during the reaction, cathodic reduction and anodic etching can be spatially isolated. Therefore, there are no metal impurities in Ti3C2Cl2 MXene. In addition, surface terminations were modified in situ from -Cl to -O and/or -S by adding various inorganic salts, which greatly shortened the modification step and enriched the variety of surface terminations. In addition, no acidic waste liquid is produced after synthesis, and the salt can be reused, making MS-E etching a sustainable method [45,46,47,48].
In addition, MXene with single or multiple halogen terminations can be prepared by the molten salt method. Taking Ti3C2 MXene powder as an example, Ti3C2 MXene with halogen termination group was prepared by molten salt reaction between Ti3AlC2 and the corresponding copper halide molten salt. Table 1 lists the composition of the initial material. Thoroughly mix the starting material with mortar under nitrogen protection in the glove box. The resulting powder mixture was then removed from the glove box and placed in an alumina crucible. The alumina crucible was loaded into a tubular furnace and heat treated at 700 °C for 7 h under argon protection to complete the reaction. Note that the reaction product is a mixture of MXene, residual salt, and copper as a by-product. The product was washed with deionized water to remove residual CuCl2 and CuBr2, and then further washed with NH4Cl/NH3·H2O solution to remove copper and CuI. The copper removal process had no significant effect on the composition and morphology of MXene. After washing the NH4Cl/NH3·H2O solution, the products were dried under vacuum at 40 °C to obtain Ti3C2 MXene with halogen terminations.
Talapin et al. used Lewis acidic CdBr2 to extend the molten salt etching route beyond chloride, and Ti3C2Br2 and Ti2CBr2 were prepared. The experimental results show that MXene with -Cl and -Br as surface terminations can exchange with other free ions, so that the surface chemical environment of MXene materials can be regulated. MXene surface exchange reactions typically require temperatures of 300 °C to 600 °C, which are difficult to achieve by using conventional solvents. In contrast, these molten alkali metal halides as solvents can overcome this problem and have high solubility for various ionic compounds. For example, Ti3C2Br2 dispersed in CsBr/KBr/LiBr eutectic reacts with Li2Te and Li2S to form new Ti3C2Te and Ti3C2S MXene, respectively. Ti3C2Cl2 and Ti3C2Br2 react with Li2Se, Li2O and NaNH2 to produce Ti3C2Se, Ti3C2O and Ti3C2(NH) MXene [29], respectively. The reactions of Ti3C2Br2 and Ti2CBr2 with LiH at 300 °C produced bare Ti3C22 and Ti2C□2, where □ stands for the vacancy site. The process can be described as the reduction and elimination of hydride groups after the exchange reaction. Finally, MXene with -O, -NH, -S, -Cl, -Se, -Br, and -Te surface terminations and without surface terminations were successfully synthesized by substitution and elimination reactions. The strategies for synthesis and removal of surface groups are shown in Figure 9.
The preparation of MXene by the molten salt method has mild experimental conditions, but it needs to be etched at high temperatures for a long time under the protection of inert gas, which is not conducive to large-scale preparation. Therefore, it is still of great research significance to explore a safe and efficient preparation method for MXene.

2.3. Ammoniation

Peng et al. introduced the termination of amino (-NH2) on the surface of MXene, and the schematic is shown in Figure 10. First, in a Teflon beaker, dissolve LiF in HCl and stir. Then, Ti3AlC2 powder was slowly added to the LiF+HCl solution and stirred continuously at 35 °C for 24 h. The resulting mixture was then centrifuged and continuously rinsed with deionized water until the pH of 6. Subsequently, the ammonia solution was added to the obtained grout and sonicated after manual shaking. The suspension was centrifuged again to remove unpeeled Ti3C2 MXene. The suspension of layered aminated Ti3C2 MXene was sonicated under nitrogen protection. Subsequently, the mixer was transferred to an autoclave and heated to 70 °C for 4 h. Finally, the supernatant was centrifuged and continuously rinsed with deionized water until the pH of the supernatant was approximately 7. The sediment was freeze-dried to obtain MXene-NH2 nanosheets [49,50]. The team proposed a substitution reaction to explain the amination process, and speculated that the positive charge in NH4+ could partially replace the -OH termination group. It is well known that the formation energy of Ti-N is higher [29,51]. Considering the moderate amination process, Ti3C2 cannot directly obtain the Ti-N bond in an ammonia solution. In addition, the surface of MXene was coated with -OH, -O, and -F after etching with a mixed solution of LiF and HCl. According to previous reports, the binding energy of Ti-OH was the lowest [26]. It can be inferred that under mild conditions it is easily replaced by other groups. Therefore, the substitution of the hydroxyl group by an amino group is a possible reaction pathway for the formation of Ti-NH2. In Figure 9, the positively charged NH4+ attacks the negatively charged -OH group to the Ti element with lower binding energy. Finally, part of the -OH group is replaced by the -NH2 group to form H2O [49].

2.4. Displacement and Esterification Method

The surface modification can also be carried out by substitution reaction, esterification reaction, and other methods. For example, Jacob et al. reported an important surface modification method by covalently attaching the polymer of PEG6-COOH to MXene nanosheets through an esterification reaction, as shown in Figure 11. Firstly, the aqueous dispersions of Ti3C2 MXene were transferred to dimethylformamide (DMF) by repeated operations of ultrasound and centrifugation until completely dispersed. DMF dispersion of Ti3C2 MXene binds PEG6-COOH and 4-dimethyl aminopyridine (DMAP). The combination of these components results in the appearance of small particles due to the assembly of -OH and -F formed by the intermolecular hydrogen bonds between PEG6-COOH and the Ti3C2 surface. Then, under the environment of nitrogen, a chloroform solution of dicyclohexyl carbon diimide (DCC) was added to initiate esterification. After continuous agitation, the small particles disappear, forming an optically transparent solution. Covalent surface modification with ω-functionalized PEG6 ligands (ω-PEG6-COOH, ω: -NH2, -N3, -CH=CH2) can regulate the types and quantities of MXene functional groups. Moreover, covalent ester bonds are chemically more stable than nitrogen bonds and electrostatic interactions [38]. On the surface of PEG6-functionalized Ti3C2, most of the hydrogen bond interactions between the sheets disappear due to ester bond formation.

2.5. Phase Transfer Method

Daesin et al. prepared stable Ti3C2Tx MXene dispersions in non-polar organic solvents by interfacial chemical grafting reaction and phase transfer method [30]. As shown in Figure 12, the alkyl phosphonic acid (CnPA) ligand was chemically grafted to the -OH terminus of Ti3C2Tx MXene sheet at room temperature through interface nucleophilic addition and sequential condensation reaction to form covalent Ti-O-P bond:
Ti OH + P OH Ti O P + H 2 O  
Unlike conventional surface chemical modification methods with complex steps, this is a simple synthesis process that can prepare surface-modified Ti3C2Tx MXene in a variety of organic solvents ranging from polar to non-polar. The dispersion of Ti3C2Tx MXene in non-polar chloroform also has strong oxidation resistance and stable long-term storage.

3. Characterization of MXene’s Terminations

Accurate characterization of MXene surface terminations is the prerequisite for studying the regulatory methods and the influence of properties and performance. Because the surface termination of MXene presents two-dimensional plane distribution [52] and the collision probability of atoms, molecules, electrons, and optical signals is low. MXene prepared by chemical methods has certain impurity content [53,54]. In addition, most surface terminations do not exist in a stable state, which leads to the difficulty of the accurate characterization of MXene surface terminations. At present, XPS, EDX, XAS and EELS are often used for qualitative and quantitative analysis of MXene surface terminations.

3.1. XRD

Although X-ray diffraction (XRD) cannot directly capture the differences of elements in the surface terminations of MXene, the evolution of MXene interlayer space induced by the dissimilar surface terminations can be deduced via the calculated MXene lattice parameters. As a basic crystal structure characterization technology, the statistical analysis of MXene crystal structure by XRD can indirectly reflect the modulating action of MXene surface termination on MXene crystal structure and the interlayer environment. Based on the information of new peak appearance, peak position shift, and half-peak width change, the surface termination atom arrangement and its induced structural evolution can be analyzed indirectly.
Figure 13a shows the XRD patterns of the initial phase Ti3AlC2 and intermediate product Ti3ZnC2. The XRD peaks of Ti3ZnC2, such as (103) and (104), shift to a lower angle than those of Ti3AlC2, indicating that the substitution of Al atoms by Zn atoms results in a larger lattice constant. The relative intensities of peaks (004) and (006) increase, while the relative intensities of peaks (002) decrease, resulting from structural factor changes caused by the substitution of A atoms [27]. Figure 13b shows the XRD patterns of the reaction products Ti3C2Cl2 and HCl rinsed products. Except for the diffraction peak characteristics of zinc, the XRD patterns of the reaction products are similar to those of Ti3C2 MXene prepared by the HF etching method. The (104) peak of Ti3AlC2 is significantly weakened or disappeared, in addition to the downward shift of the (002) peak to a lower Angle, attributed to the increased C lattice parameters. The C lattice parameter is consistent with the theoretical value calculated by DFT and larger than Ti3C2Tx generated by HF etching [2]. Note that unlike the typical wide (00L) peak of Ti3C2Tx produced by HF etching, the (00L) peak of Ti3C2Cl2 is sharp and intense, indicating an ordered crystal structure. Therefore, MXene prepared by this method with -Cl as the surface termination has a stronger XRD peak and a more orderly crystal structure. Other MXene with different functional groups also have different manifestations in the XRD. As shown in Figure 13c, the (002) characteristic peaks of Ti3C2-O and Ti3C2-S obtained by treating molten salt-assisted electrochemical (MS-E) etched Ti3C2Cl2 with Li2O and Li2S were further shifted from 8.18° to lower angles of 7.48° and 6.88°. This is attributed to the C lattice parameters ranging from 21.9 Å to 23.6 Å and 25.9 Å [44].
Figure 13d shows the XRD patterns of MXene and Ti3C2 (OF) at halogen terminations. It is found that Ti3C2Cl2 and Ti3C2 (OF) have strong (002) peaks, corresponding to the C values of 20.90 and 22.22 Å, respectively. The (002) peaks of Ti3C2Br2 and Ti3C2I2 are not very clear, but the (004) peaks are strong enough. Sufficient to determine their respective C values of 23.33 and 25.00 Å, which are slightly larger than the theoretical values calculated by density functional theory (DFT) (23.02 and 24.24 Å). This indicates that the layer spacing of MXene connecting halogen groups increases successively with the radius of halogen groups. The increased layer spacing in Ti3C2Br2 and Ti3C2I2 can be interpreted as an increase in the atomic radius of the halogen group from fluorine to iodine. Different termination atoms will lead to different structure factors of MXene, resulting in different relative intensities of the (00L) peak. The c-value of binary and ternary halogen-terminated MXene is between that of monoid halogen-terminated MXene. For example, the C value of Ti3C2 (BrI) (24.06 Å) is larger than that of Ti3C2Br2, but smaller than that of Ti3C2I2 [28].
Figure 13e shows the XRD patterns of conventional MXene and MXene with amino surface functionalization. Compared with the peak (002) of MXene at 6.21°, the peak (002) shifted to a large angle of 6.63° after -NH2 functionalization. It shows that their D-spacing decreases from 1.42 nm to 1.33 nm. The decrease in D-spacing is due to the introduction of positively charged amino groups on the surface of MXene-NH2, which reduces the repulsive force of the original negative charge on the MXene nanosheet. Therefore, the small D-spacing in MXene-NH2 can be judged by XRD [55].

3.2. SEM and EDS

Figure 14a,b show SEM images of Ti3AlC2 and Ti3ZnC2 powders. The Ti3AlC2 precursor exhibits a typical MAX phase layered structure. In contrast, the layered structure of Ti3ZnC2 becomes less pronounced, probably due to the dissolution of the powder edges in the molten salt. Ti3C2Cl2 was obtained by adding excessive zinc chloride. Figure 14c,d shows the semi-quantitative analysis of SEM of Ti3C2Cl2 and the corresponding EDS (energy dispersive spectroscopy). Typical accordion-like morphology was found, and the main elements were Ti (43.2 at.%), C (21.5 at.%), and Cl (25.3 at.%). As well, small amounts of Zn (0.7 at.%), Al (2.9 at.%), and O (6.3 at.%). The presence of Zn, Al, and O is plausible because ubiquitous O-containing compounds, such as Al(OH)3, are hydrolysates of AlCl3.
Shen et al. prepared -O termination MXene and -S termination MXene by adding Li2O or Li2S to the melting system to modify Ti3C2Tx powder. SEM shows that the modified Ti3C2Tx powder still maintained accordion-like multilayer morphology (Figure 14e,f) and was tightly stacked. In EDS analysis, element Cl was not detected, and elements O and S were the main components of surface terminations.
Huang et al. prepared Ti3C2Cl2, Ti3C2Br2, Ti3C2I2, Ti3C2(ClBr), Ti3C2(ClI), Ti3C2(BrI) and Ti3C2(ClBrI) using halogen atoms as terminations by molten salt method. The morphology and elemental composition were characterized by SEM and EDS, as shown in Figure 14g–m. The SEM confirmed that MXene has typical accordion morphology, and EDS analysis obtained their main composition elements (Ti, C) and end-group elements (Cl, Br, I), which can judge the morphology and preliminary functional success of MXene, indicating that halogen atoms almost completely occupy the termination position of the surface.
Based on the combination of SEM and EDX, the surface terminations of MXene in the microzone can be observed, and the information on element species, distribution, and content can be preliminarily obtained. As a fast and efficient means of characterization, SEM+EDX can achieve high-throughput data collection, which provides support for the study of MXene surface terminations.

3.3. XPS

XPS can accurately measure the inner electron binding energy and chemical shift of atoms, provide information on molecular structure and valence state, and provide information on element composition and content, chemical state, and molecular structure. In addition, because the X-ray beam incident on the sample surface is a photon beam, the damage to the sample is very small. XPS is used as the most conventional means to analyze the surface terminations of MXene. With the accurate representation of XPS elements, the valence states, chemical bonds, and chemical shifts of surface termination elements are collected. Meanwhile, the mechanism of the surface termination formation process can be obtained by comparing the XPS high-resolution spectrogram. Providing support for the development of differentiated surface terminations is one of the most powerful characterization techniques in MXene surface termination research.
XPS spectra of Ti 2p and Cl 2p of Ti3C2Cl2 are shown in Figure 15a. The peaks at 454.4 eV and 455.7 eV are assigned to the Ti-C(I) (2p3/2) and Ti-C (II) (2p3/2) bonds [56,57]. The peak at 458.1 eV belongs to the high-valence Ti compound, which is assigned to the Ti-Cl (2p3/2) bond [58,59]. In addition, the peaks at 460.3 eV, 461.8 eV and 464.1 eV are attributed to Ti-C (I) (2p1/2), Ti-C (II) (2p1/2) and Ti-Cl (2p1/2) bonds, respectively. The right picture of Figure 14a shows its Cl 2p spectrum. The peaks at 198.6 eV and 200.1 eV are consistent with the positions of Cl-Ti (2p1/2) and Cl-Ti (2p3/2) bonds [58,59]. This confirms the existence of Ti-Cl bonds. The Ti-Cl ratio determined by XPS analysis was 2.94:2.
High-resolution Ti 2p spectra (Figure 15b) also show that Ti3C2Cl2 forms a Ti-O (2P3/2) bond after treatment in Licl-KCl-Li2O, so that there is no Ti-Cl bond at 458.9 eV and a new peak occurs. For Ti3C2Tx treated with Li2S, the Ti-Cl bond was also not visible. There are two peaks at 459.2 eV and 456.8 eV, which are attributed to Ti-O (2p3/2) and Ti-S bond (2p3/2), respectively [29]. These results indicate that -Cl surface terminations change to surface terminations with -O and/or -S.
XPS can also confirm the amino groups on the surface of MXene, as shown in Figure 15c. It can be seen that there are only four major elements C, F, Ti, and O in the XPS spectrum of conventional MXene. However, the N 1s peak appears in the measured spectra of MXene-NH2 films and CMC/MXene-NH2 composite films. The evolution of the N 1s spectrum shows five peaks that can be attributed to the N-Ti, N-C, -NH2, -NH4+ and -NH3+ bonds, respectively. -NH4+ and -NH3+ should result from the protonation of the NH3·H2O and -NH2 groups, respectively. The Ti 2p high-resolution spectrum shows two peaks at 455.3 eV and 463.2 eV, originating from the Ti-N and Ti-F bonds, respectively. The C-N bond in MXene-NH2 was also shown. Thus, the N element in MXene-NH2 consists of Ti-N, C-N, -NH2, and -NH3+ forms. Importantly, the appearance of -NH2, -NH3+ and Ti-N bonds would be attributed to the introduction of amino-termini on the surface of MXene.
The XPS measurement spectra of Ti3C2Br2 and Ti3C2Te MXene were compared, as shown in Figure 15d. After surface group exchange in CsBr/LiBr/KBr molten salt, the Br peak was replaced by the Te peak. The Cs 3D peak corresponds to the intercalated Cs+ ions or residual Cs+. Figure 15e shows the high-resolution XPS spectrum of Ti3C2Br2 MXene and Ti3C2Te MXene. The binding energy of the Ti-C component of Ti3C2Br2 MXene shifts to a lower value after Br- replaces Te2-. This result is consistent with the fact that Te is less electronegative than Br. Figure 15f shows the XPS spectrum of Ti3C2 (NH) MXene, measured with the highlighted N 1s region. The elemental analysis of the measured spectrum shows that the Ti:N ratio is 3:1.1, and the analysis of the high-resolution N 1s spectrum shows that there are three components. The 396.2 eV peak (59%) belongs to the β-N chemisorbed on the Ti surface [61]. The 397.5 eV peak (26%) corresponds to α-N2 chemically adsorbed on the titanium surface. The 399.9 eV peak (16%) may correspond to the N-H bond. Figure 15g shows the high-resolution XPS spectra of Ti3C22 MXene. Compared with Ti3C2Br2 MXene in Figure 14e, the Ti-C component binding energy to a lower value after Br- has been eliminated with H-.
In addition, Chen et al. compared the XPS spectra of original Ti3C2Tx MXene and MXene modified by -SO3H (Ti3C2Tx-PSS), which prepared Ti3C2Tx MXene composite membranes by intercalating MXene nanosheets with sodium 4-styrene sulfonate (PSS), a spacer containing sulfonic acid group (SO3H). The spacer introduced abundant -SO3H groups into the membrane, as shown in Figure 16a,b. Compared with the F 1s spectra of Ti3C2Tx MXene, the signals of F-C and F-Ti move from 687.00 eV and 684.98 eV to 686.36 eV and 684.94 eV, respectively, indicating that these sites bind protons from the PSS. In addition, the S 2p XPS spectrum of MXene-PSS shows two peaks representing S 2P1/2 and S 2p3/2, in contrast to the S 2p spectrum of MXenewithout any peaks. This indicates that a large number of -SO3H groups were successfully introduced into the membrane.
XPS is used to analyze the MXene surface termination, realizing the accurate confirmation of complex multi-element and hybrid terminations. With the help of high-resolution characteristic element spectrography analysis, combined with computational simulation, the fine structure of MXene surface termination can be obtained. From qualitative to quantitative, provide technical support for MXene surface analysis. However, it is also limited to surface analysis, and is difficult to obtain termination information. It is easy to be affected by the external environment and this peak fitting is subjective.

3.4. TEM

TEM and spherical aberration-corrected electron microscopy can directly observe the atomic composition of MXene at the atomic scale, and the surface terminations of heavy elements can be directly observed [60]. With the help of sample preparation techniques such as frozen section and FBI, ultrathin samples can be obtained, and the arrangement of spatial terminations in the MXene layer can be observed. Depending on the support of testing methods such as EDX and EELS, atomic identification, valence analysis, and coordination relationship judgment can be realized at the atomic scale. In addition, the structure analysis of MXene can be realized by selecting the comparison between electron diffraction and XRD.
Figure 17a is a TEM image of Ti3C2Cl2 MXene with organ-like morphology. Figure 17b is a TEM image of MXene-NH2 similar to graphene. In addition, the HRTEM image and its selected area electron diffraction pattern (inset) confirm that the crystal structure of the aminated nanosheets is the same as that of conventional MXene. Sardar et al. Modified MXene with PEG6 and observed the morphology by TEM. Figure 17c shows TEM images of pristine Ti3C2 MXene flakes and modified Ti3C2-PEG6, respectively. Any small nanoparticles formed in the form of TiO2 nanoparticles in the Ti3C2-PEG6 flakes were not observed. In addition, the length of Ti3C2-PEG6 flakes is in the order of microns. These results indicate that the chemical modification of the Ti3C2 MXene flakes does not change the size of the MXene flakes and does not lead to any oxidation of the MXene flakes. This is an important discovery for maintaining the structural integrity of the Ti3C2 MXene flakes. Yu et al. achieved surface amino functionalization of Nb2CTx MXene by hydrazine treatment. Figure 17d,e are TEM and HRTEM images of pristine P-Nb2CTx and T-Nb2CTx MXene with amino functionalization. All the Nb2CTx MXene exhibited a typical 2D morphology, and the size and thickness of Nb2CTx MXene did not change before and after treatment. HRTEM images of P-Nb2CTx and T-Nb2CTx MXene nanosheets indicate that the lattice parameters of P-Nb2CTx and T-Nb2CTx MXene nanosheets are 0.27 nm. In addition, the corresponding selected area electron diffraction (SAED) patterns of P-Nb2CTx and T-Nb2CTx MXene nanosheets are shown in the inset. This means that the P-Nb2CTx and T-Nb2CTx MXene nanosheets have hexagonal structures.
Karlsson et al. used STEM-EELS to study the chemical and structural properties of individual Ti3C2 MXene sheets with associated surface groups at the atomic level, as shown in Figure 18. MXene show inhomogeneous coverage of surface groups that locally affect chemical properties. According to the EELS measurement results, the average components of Ti, C, O, N and F were 47, 33, 20, 3, and 1 at.%, respectively, ignoring the edge effect of a single MXene sheet. Double layer MXene sheets corresponding to 46, 29, 24, 2, and 1 at.%. For multilayer MXene sheets, 40, 24, 31, 2, and 3 at.%, respectively. This indicates that C decreases while F and O groups strongly increase upon a larger thickness of MXene, and the presence of N is also revealed by EELS. As well, while EELS can recognize O as an element and spatially link O to surface groups, it cannot distinguish between molecules [63,64]. In contrast, XPS was unable to distinguish the number of sheets but identified the presence of Ti-bound Ox, (OH)x and absorbed H2O as well as a small fraction of TiOx.

3.5. Raman and FTIR

Raman spectroscopy is an analytical method based on the Raman scattering effect discovered by the Indian scientist C.V. Raman, which analyzes the scattering spectrum with different frequencies from the incident light to obtain information on molecular vibration and rotation, and is applied to the study of molecular structure. Raman spectrometer features: (1) No contact or damage to the sample, and the sample does not need to be prepared. (2) Rapid analysis and identification of the characteristics and structure of various materials. (3) Suitable for black water and water samples, accurate measurement in high, low temperature, and high-pressure conditions. Application directions of Raman spectroscopy: Raman spectroscopy is a molecular structure characterization technique based on the Raman effect, and its signal comes from the correlation between molecular vibration and rotation. The analysis directions of Raman spectroscopy are as follows:
  • Qualitative analysis: Different substances have different characteristic spectra, so qualitative analysis can be carried out by spectra.
  • Structure analysis: The analysis of spectral bands is the basis of material structure analysis.
  • Quantitative analysis: According to the absorbance characteristics of the spectrum of substances, one can have a good ability to analyze the amount of substances.
MXene spectrum mainly has four characteristic regions: formant, thin section, Tx area, and carbon area. MXene can be characterized according to the change of peak displacement, intensity, width, and other characteristics in the characteristic region. Different surface terminations, intercalated, and adsorbed species may affect the lattice vibrations. Unit cell distortion and distribution of surface groups result in peak shifting and broadening. Both HF-HCl and LiF-HCl etched Ti3C2Tx spectra show strong -OH surface groups and similar Alg (C) peak positions. However, the -O component dominates the HF etched multilayer Ti3C2Tx spectrum, and the peak position of A1g shifts to 711 cm−1, mainly because the addition of HCl reduces the pH value and leads to surface protonation. Talapin et al. measured the Raman spectra of S-terminated MXene using 632 nm laser excitation, and the peak positions of the out-of-plane (A1g) and in-plane (Eg) vibrational modes corresponding to the surface groups are particularly sensitive to MXene surface chemistry. Raman spectra of Ti3C2Cl2, Ti3C2S (Figure 19a), and Ti3C2NH (Figure 19b) MXene show the characteristic A1g and Eg vibrational modes. The Raman spectrum is different from that of Ti3C2O MXene (Figure 19c). In particular, for Ti3C2Cl2 MXene, the position of the A1g mode remains at about 170 cm−1; for the position of the Ti3C2S MXene, A1g mode is kept at 180 cm−1; for the position of the Ti3C2NH MXene, A1g mode remains at about 210 cm−1. The A1g pattern is completely different from that of Ti3C2O MXene at 222 cm−1. The position of the A1g mode is mainly determined by the atomic mass of the surface group, and Br and Se lead to lower frequencies than Cl and S. The Eg mode corresponds to the in-plane vibration of the surface group and the external Ti atom has a similar tendency [29]. The 153, 235, and 277 cm−1 characteristic peaks of the Ti3C2 vibrational Eg, A1g, and Eu modes appear in the bare Ti3C2 MXene Raman spectra. The characteristic peaks of Ti3C2 vibration Eg and A1g modes in Ti3C2-NH, Ti3C2-O, and Ti3C2Tx are blue-shifted. As well, the Eg mode of the in-plane vibration of the surface group does not exist in the region of 300–400 cm−1 where a characteristic peak appears corresponding to the Eg mode of the in-plane vibration of the surface group. The assignment of the bare Ti3C2 vibrational modes is based on the work of Hu et al. The appearance of the Raman-forbidden Eu mode at 277 cm−1 may be related to the disorder present in the stack of bare Ti3C2 MXene sheets.
Raman spectroscopic analysis of Ti3C2-PEG6 flakes is shown in Figure 19d. First, the peak at 123 cm−1 is the result of the incident laser interacting with the plasmon resonance of the material. In addition, the planar peaks of (Ti, C, O) are clearly visible near 203 cm−1 and at 725 cm−1. Second, the Raman peak position of A1g is slightly shifted to the left when PEG6 is esterified to form a link, and there is a similar change when the surface group is changed. The most important feature is the absence of Raman stretching due to possible oxidation products during the processing and analysis of the Ti3C2 MXene. If there is obvious oxidation in the PEG6-functionalized Ti3C2 MXene sample, it is expected that a very pronounced Raman stretching will be observed due to the formation of TiO2 [38]. Raman spectra show that there is no obvious oxidation of Ti3C2 MXene flakes during the synthesis process, which is prevented by the extremely mild reaction conditions.
In particular, Raman can not only clarify the molecular structure of materials, but also illustrate the reaction process and energy conversion mechanism. To prove the energy conversion mechanism of halogenated MXene, the structural and compositional changes of Ti3C2Br2 and Ti3C2I2 under different charge-discharge states were studied. As shown in Figure 19e, f, a broad Raman shift at 260–270 cm−1, corresponding to Brn (3 ≤ n ≤ 5), was detected in the Ti3C2Br2 electrode upon charging. Considering the reversible reaction, Br+ (n − 1) Br0↔ Brn, the formation of Brn indicates the generation of Br0 (Br2) due to the oxidation of Br. Upon discharge, after trapping electrons, the peak of Brn disappears because Br0 is restored to Br [28].
In addition, a recent study showed that the low sensitivity of Raman scattering was improved by using gold or silver tips, increasing Raman scattering by six orders of magnitude. Although Raman spectroscopy of monolayer MXene sheets can be studied, Raman imaging of single sheets has not been reported so far.
FTIR is an infrared spectrometer based on the principle of the Fourier transform of infrared light after interference. It can be used for qualitative and quantitative analysis of samples, not only to measure the absorption and reflection spectra of various gas, solid and liquid samples, but also to measure chemical reactions in a short time.
In order to determine the amino end groups on the surface of Ti3C2, FTIR spectroscopy was performed. As shown in Figure 20a, MXene-NH2 also has -F, -O, -OH groups similar to traditional MXene. However, the appearance of two characteristic peaks can be observed. Generally, the characteristic peak at 3313 cm−1 corresponds to the N-H stretching vibration, and the peak at 1617 cm−1 is caused by the N-H bending vibration. Therefore, it can be preliminarily proved that amino groups are successfully introduced on the surface of MXene. In addition, the peak of the 1650 cm−1 CMC/MXene-NH2 film was assigned to the amide group (-NH-CO-) stretching vibration, which was attributed to the thermal reaction between the carboxyl group in the CMC and the amino group on the MXene-NH2 film. Thus, it further confirms the amination termination on the surface of MXene-NH2. This also makes the MXene-NH2 nanosheets in the CMC/MXene-NH2 composite film more compactly stacked. These interfacial covalent bonds will also impart enhanced mechanical properties to the CMC/MXene-NH2 film. Figure 20b shows the FTIR spectra of pure Ti3C2 MXene flakes and surface modification with PEG6-COOH. The green box indicates the presence of -OH in the sheet structure, and the green line represents a typical surface group bound to a Ti atom, such as the vibrational mode of Ti-O at ~550 cm−1. The characteristic low-energy vibrational mode (green dashed line) remains unchanged even after the attachment of PEG6 (see blue spectrum). The O-H stretch at ~3400 cm−1 is absent, and it is thought that this may be due to the formation of -COO bonds by the esterification chemistry of all -OH. In addition, the carbonyl stretch of pristine Ti3C2 is present in the Ti3C2-PEG6 flakes, as well as an additional carbonyl stretch at 1704 cm−1, derived from the ester bond carbonyl. Therefore, the FTIR characterization strongly supports the covalent surface modification of pristine Ti3C2 MXene with PEG6-COOH through ester bond formation.

3.6. Others

Figure 21a shows Ti3C2 with Cl atoms as surface terminations. Ti3C2Cl2 MXene shows good ordering along the basal plane, which is consistent with the strong (000l) peak in the XRD pattern. The electronic band structure and phonon spectrum calculated by DFT for a single layer of Ti3C2Cl2, which is essentially a metallic layer with a finite density of states at the Fermi level, are shown in Figure 21b. The phonon spectra also show that all phonon frequencies are positive, i.e., Ti3C2Cl2 MXene is dynamically stable [27].
Figure 21c shows the HAADF-STEM image of Ti3C2Br2 and the corresponding EDS elemental map. Br element was uniformly distributed on the particles, and no element segregation was observed. Figure 21d, HR-TEM image of Ti3C2Br2 shows that the d value of the (0002) plane is ~11.77 Å, which is consistent with the XRD result [28]. Figure 21e,f are atomic resolution scanning transmission electron microscopy (STEM) images, and the corresponding EDS elemental maps show the -Br terminations of the Ti3C2 surface. The brightness of the atoms depends on the mass-dependent scattering conditions, demonstrating also the successful functionalization of the MXene surface by Br. The atomic and electronic structures of two single-halogen-terminated MXene, Ti3C2Br2, and Ti3C2I2, were investigated by density functional theory (DFT) calculations, as shown in Figure 21g. Atomic and electronic structures of Ti3C2Br2 and Ti3C2I2. -Br and -I preferentially occupy the FCC site, which is similar to -Cl in Ti3C2Cl2 and -O and -F in Ti3C2(OF). For the interplanar spacing of Ti3C2Br2 and Ti3C2I2, the halogen terminations and the outermost Ti layer are 1.86 and 2.0606 Å, respectively, corresponding to the strong bonding of Ti-Br with a bond length of 2.6464 Å and the weaker bonding of Ti-I with a bond length of 2.8383 Å [28]. The bonding strength is further confirmed by the formation energy (Ef) of the halogen atoms on the surface of MXene, which is directly related to the bonding strength between the end group and the Ti3C2 matrix. The calculated Ef for Ti3C2Br2 and Ti3C2I2 are −5.624 and −3.644 eV, respectively, which are less negative than the previously reported Ti3C2Cl2 value of −6.694 eV, Ti3C2F2 (−7.111 eV) and Ti3C2O2 (−9.589 eV). In addition, the AFM image and size distribution of the obtained MXene-NH2 nanosheets are shown in Figure 21h [49].
To establish an MXene characterization platform measuring the termination structure, elemental composition, and coordination relationship, and developing accurate quantitative test methods are the key route for MXene research.
However, quantitative analysis of MXene is still challenging due to the lack of accuracy. As well, most of the functional groups are light and difficult to detect and resolve. The development of new characterization methods is necessary for a further study of MXene surface terminations.

4. Properties Depended on Surface

As a new 2D layered material, MXene has the merits of other 2D materials. More importantly, surface functionalization renders them easily achieve improved properties. Meanwhile, the applications of MXene have been expanded, including biomedical, energy storage devices (battery, supercapacitor), sensors, catalysis, and electromagnetic interference shielding [65].

4.1. Electrical Conductivity

MXene has a higher electrical conductivity, with a theoretical electron mobility of up to 104 cm2/(V · s). In 2014, Gogotsi et al. obtained a flexible self-supported Ti3C2Tx MXene thin film material by vacuum filtration of the monolithic dispersions of Ti3C2Tx MXene, and its electrical conductivity was 1500 S/cm by the four-probe method. Mathis et al. used the MAX phase with more aluminum as the precursor to reduce the defects of the Ti layer on the MXene surface during etching, and obtained Ti3C2Tx MXene film with electrical conductivity up to 20,000 S/cm. By adjusting the surface chemical properties of MXene, the electronic conductivity of MXene can also be effectively adjusted. Some surface terminations (especially -F) can be changed or removed, so that the lamellar structure of MXene is more compact and the conductivity is significantly increased.
DFT calculation results show that MXene without surface terminations generally behaves as metal because the d electrons of transition metals have a very high density of states (DOS) near the Fermi level. However, MXene synthesized by experiment often has abundant surface terminations, and the number and type of surface terminations are difficult to accurately control. In addition, there is still a lack of advanced characterization methods for the conductivity test of MXene monolayer nanosheets, which poses a challenge to the study of the intrinsic electronic properties of MXene. According to the calculation results of Perdew-Burke-Ernzerhof (PBE) and HSE06, as shown in Figure 22, both Ti2X and MAX phases exhibit metal properties, while the density of states of Ti2XT2 with surface terminations at the Fermi level decreases significantly.
With the introduction of surface terminations, the band structure of MXene changes significantly. Due to the hybridization between the p orbital of the end group and the d orbital of the transition metal M, the peak density of states is significantly reduced, and it is possible to raise the d orbital above the Fermi level, resulting in a band gap. In addition, a larger atomic number of M usually produces a larger bandgap value [65]. Due to the generation of the band gap, the MAX phase and the MXene (Mn+1Xn) without termination groups, which originally exhibited metal properties, are transformed into semiconductors after the introduction of termination groups (Mn+1XnTx), as shown in Table 2.
The results show that MXene with a larger n value is more likely to exhibit metal properties, and the introduction of surface termination has little effect on its metal properties. For example, Ti3C2, Zr3C2, Ti4C3, Nb4C3 and Ta4C3 MXene exhibit metal properties no matter which is the surface termination. In the ordered bimetallic MXene structure, the density of states near the Fermi level largely depends on the D electrons of the surface metal atoms. Therefore, compared with the inner metal, the outer transition metals have a greater effect on the electronic properties. Compared to carbide MXene, nitride and carbonic nitride MXene generally have higher total state densities at Fermi levels due to stronger Ti-N bonds (compared to Ti-C bonds) and extra nitrogen electrons [66].
The effect of amino termination on the electronic behavior of MXene-NH2 was studied by UPS spectroscopy. The finite density of states of MXene and MXene-NH2 nanosheets shown in Figure 23a at the Fermi level (set at 0 eV) indicates that MXene-NH2 exhibits metallic conductive behavior similar to conventional MXene. The secondary electron cutoff regions of the UPS spectra of the two nanosheets are shown in Figure 23b. The results show that the work function of MXene-NH2 increases from 6.28 eV to 6.33 eV. It has also been shown that NH4+ is an electron acceptor during amino substitution [90]. Although NH4+ replacing the -OH reduce the electron carrier density of MXene-NH2, which has a higher electron transport. As shown in Figure 23c, the conductivity of pure MXene-NH2 film reaches 21,100 S/cm, which is still a considerable value, although it is 82.5% lower than that of conventional MXene film.
Since the surface terminations of MXene can provide wider layer spacing and more space for ion shuttling [91], its application in various energy storage can be improved by controlling the interconversion between different functional groups. For example, among various MXene, the surface-modified Ti3C2Tx dominates the current solar cell research due to its high conductivity, carrier mobility, and adjustable work function (WF) [91]. Transition metals tend to be less electronegative than functional groups and carbon atoms, with -OH and -F accepting one electron and -O accepting two electrons from the transition metal atom. In addition, Ti, Zr, and Hf are located in the same column of the periodic table and have the same shell electron configuration, therefore, after -O functional modification, a similar trend of metal-semiconductor characteristic transition is found in the corresponding M2C MXene [66]. This property can be used in capacitors to achieve energy storage. In addition, MXene can be used as a cocatalyst for the separation and transfer of charge in photocatalytic reactions due to its good conductivity [92]. After the modification of functional groups, it showed excellent performance and reduced the reaction cost.

4.2. Magnetic Properties

Most MAX phases reported so far are not magnetic, while some MAX phases containing Cr or Mn are magnetic. Theoretical calculation shows that most of the unterminated MXene obtained after removing the A atomic layer in the MAX phase are magnetic. To determine the magnetic moment of the unit cell of MXene, the factors to be considered mainly include the transition metal oxidation state, the coordination number of the transition metal, and the number of d electrons. Moreover, C4−, N3−, F, OH and O2− can also contribute a small amount of magnetic moment. The unbonded d orbitals in MXene are formed near the Fermi level between the bonded and anti-bonded states; therefore, the electrons occupying the unbonded D orbitals play a major role in magnetism [93].
According to the theoretical prediction, most of MXene without termination groups have magnetic properties, but when surface termination groups are introduced, the magnetic properties of MXene may change greatly, as shown in Table 3 [65]. In addition, the magnetic properties of MXene of different transition metals usually show different changes when they are combined with different types of surface end groups. For example, some magnetic MXene, such as Ti3C2Tx and Ti4C3Tx, change to non-magnetic after surface termination. When the surface terminations of Cr2CTx are -OH and -F, the ferromagnetism remains at room temperature, and the magnetism disappears when the surface termination groups of Cr2CTx are -O. However, Mn2NTx always remains ferromagnetic, independent of surface termination. Monolithic V2C is an antiferromagnetic metal, and after -F or -OH termination, V2C transforms into a small bandgap antiferromagnetic semiconductor. The gratuitous group Cr2C is a ferromagnetic semi-metal; however, with the end-group of -F, -Cl, -OH, or -H, Cr2CTx becomes an antiferromagnetic semiconductor. Cr2N has more valence electrons than Cr2C, and the ground state of Cr2N is antiferromagnetic, but it becomes semi-metallic with -F, -O, or -OH surfaces.
Some magnetic MXene exhibit nonmagnetic properties after binding to surface termination, which is mainly due to the formation of p-d bonds between surface transition metal atoms and end groups, leading to the disappearance of magnetism [94]. Orbital splitting theory can also be used to explain the phenomenon of magnetic properties change of MXene: For MXene with surface termination, each transition metal atom is surrounded by C/N atoms and surface termination, forming an octahedral cage around the transition metal. The resulting crystalline field near the octahedron divides the d orbital of the transition metal into t2g (dxy, dyz, dxz) and eg (dx2−y2, dz2) orbitals. Because of the different orbital shapes, the energy of the eg orbital manifold is higher than that of the t2g orbital manifold; Therefore, the electrons occupy the t2g orbital first. The valence electrons of transition metals are different, and the electron configurations of d orbitals are different, so the magnetic behavior of MXene is different.
MXene with asymmetric surface termination, has also been studied. Cr2CFCl, Cr2CHF, Cr2CFOH and V2CFOH are antiferromagnetic semiconductors [95]. In Sc2CTT ′(T, T′=H, O, OH, F, and Cl,) with different asymmetric combinations, these containing -O are ferromagnetic ordered semi-metal, while the others are non-magnetic semiconductors. The magnetic properties of MXene vary greatly due to different termination combinations. Therefore, MXene can afford various magnetic properties by adjusting termination, and the application of MXene will be expanded into the field of magnetism.

4.3. Optical Properties

In the visible range, the presence of -F and -OH decreases the absorption and reflectance of MXene, while in the UV region, the presence of surface terminations increases the reflectance compared to bare MXene. In the infrared and visible light ranges, -O has a higher ε2 (ω) than -F with -OH, and therefore has a stronger optical absorption performance, which is mainly attributed to the formation of oxidation states close to the Fermi level.
The surface modification of MXene using magnetic transition metals, such as Fe, CO, Ni, etc., can improve the optical properties of Ti3C2 MXene material and increase the optical absorption coefficient of Ti3C2 in the ultraviolet region, visible region, and near-infrared short-wave region [96]. The optical absorption coefficient of Ni/Ti3C2 composites increases by 50% in the visible and ultraviolet light range, and the optical absorption coefficient of Fe/Ti3C2 increases by more than 100% compared with that before modification, especially in the infrared region. This significant enhancement of light absorption may be related to the change of electronic structure after doping magnetic atoms. DFT calculations show that Ti3C2 MXene retains its metallic properties after Fe, Co, and Ni modification, and the d electron orbital plays a dominant role in the electronic states near the Fermi level. A strong peak in the range of −4~−1 eV appears in the d orbital, resulting in a spin-polarized state, leading to a great change in the optical properties.

4.4. Solubility and Dispersion

Ti3C2 powder with -Cl as termination prepared by the molten salt method is black, as shown in Figure 24a. These substances were well dispersed in absolute alcohol or water (0.17 mg mL−1) and exhibited a pronounced Dundar effect. In addition, the effect of amino closure on the dispersion of MXene-NH2 solution was also pointed out, and the synthesized Ti3C2 MXene-NH2 nanosheets were tested for Zeta potential. As shown in Figure 24b, compared with traditional MXene, the Zeta potential of MXene-NH2 ranges from −51.3 mV to −30.3 mV in water and from −6.68 mV to −5.80 mV in 75% ethanol solution. This is mainly due to the introduction of an amino terminus on the surface of MXene. The amination of MXene-NH2 reduces the electronegativity of the nanosheets. So, its Zeta potential goes down. However, MXene-NH2 nanosheets also showed excellent dispersion in aqueous suspensions. After standing for 2 weeks, only a small amount of MXene-NH2 precipitation was observed at the bottom of its aqueous suspension, as shown in Figure 24c. Good dispersion and the same Zeta potential are the keys to the preparation of homogeneous hybrid films by vacuum-assisted filtration. The excellent dispersibility of the solution provides a feasible preparation method for MXene-NH2 nanosheets [49].

4.5. Mechanical and Tribological Properties

Ti3C2Tx MXene clays synthesized by the LiF/HCl method show good compression properties and can be rolled into films with tensile strength up to 568 ± 24 MPa [97]. To study the effect of different surface terminations on the mechanical properties of MXene, the bond energy of M − Tx tends to be maximized in the early transition metal groups, which is interpreted as the tendency of the surface terminations (halides) to contribute electrons to most of the unfilled D orbitals in the early transition metal groups on the surface. This reduces the electronegativity difference between the surface termination and the transition metal in the late transition metal group [98]. In general, -O2 terminations resulted in higher 2D stiffness of MXene, whereas -F2, -(OH)2, and -Cl2 terminations resulted in a slight increase in stiffness (Figure 25). The effect of different surface terminations on mechanical properties can be attributed to the changed M-X structure [99,100,101]. In the Mn+1 Xn structure, the addition of strong mixed ion/covalent M-O bonds reduces the contribution of M-M interactions [98], which helps to increase the strength of the M-X bond. However, the major ionic bonds of M-F, M-Cl, or M-(OH), did not result in a reduction in M-M interactions compared to the M-O bond. An exception can be found at the −Cl2 end of the posterior transition metal, which is the result of covalent bonds, due to the length of the M-Cl bond being longer than that of the M-F bond and the increased ionization energy of the right transition metal in the periodic table. This could explain the similar mechanical properties of V2CCl2 and Cr2CCl2 compared to V2CF2 and Cr2CF2 [102].
The friction of the amino-functionalized Ti3C2Tx MXene-reinforced epoxy composite was reduced by about 35%, and the wear rate was reduced by 72%. This is attributed to the improved dispersion and compatibility of functionalized MXene in the organic matrix, resulting in increased storage modulus and interfacial strength, and reduced plastic deformation and microcracking [103]. The hydrophilicity of the surface terminations of MXene reduces the stability of its dispersions in polar solvents. Thus, we can use the existing functional groups (−O2, −(OH)2, −F2, and −Cl2) as anchors for hydrophobic molecules to progressively manipulate their hydrophobic/hydrophilic properties. This significantly improves the stability of dispersants in hydrophobic liquids and reduces agglomerating, thus improving the resulting frictional properties.

4.6. Water Purification

The ultrathin two-dimensional structure and high surface area of MXene lead to the exposure of more internal atoms, which leads more likely to the formation of defects and in turn provide more surface active sites, which will facilitate its application in the adsorption field [104]. Therefore, MXene materials have been used to remove inorganic pollutants and dye compounds. The adsorption mechanism can be interpreted as pure MXene adsorbs pollutants with positive and negative charges due to its surface charge. However, the surface of MXene is rich in functional groups that can interact with opposite charges of compounds in water. In addition, due to the large layer spacing of pure MXene, pollutants can be embedded in MXene and then removed [105].
According to density functional theory (DFT), the bond lengths of termination groups (such as -OH, -F, =O) on the surface of MXene are about 0.97 Å for Ti-OH, 2.1 Å for Ti-F and 1.9 Å for Ti-O. Compared with -OH, -H, -F, and -Cl, =O has the largest adsorption energy. This will affect the adsorption properties of MXene [105].

4.7. Biological Characteristics

The antimicrobial mechanism of MXene may be the direct physical interaction between the sharp edge of the nanosheet and the bacterial membrane, and the highly hydrophilic surface with a negative charge can enhance the contact between bacteria and the membrane surface, and contact killing leads to microbial inactivation. In addition, the antibacterial activity of MXene was related to particle size and action time [106].
Wu et al. immobilized the enzyme on the -OH surface termination to maintain the biological activity and stability of MXene on the media-free phenol biosensor platform [107]. It was found that the electrostatic adsorption of lysozyme [108] and collagen [109] on the surface of Ti2C and Ti3C2 resulted in changes in physicochemical and biological properties of MXene. Liu et al. obtained similar results by modifying the surface of Ti3C2 with hyaluronic acid, thereby achieving active tumor targeting and aggregation and enhancing the stability of the multifunctional nanoplatforms [110]. Notably, polyethylene glycol (PEG) can also be used to modify Ti2C to stabilize the photothermal ablation of cancer cells and minimize negative effects on nonmalignant cells [111]. Other studies have used soy phospholipid-stabilized Ta4C3 MXene [112] or polyvinylpyridol (PVP) -stabilized Nb2C [113] for dual-mode photoacoustic/CT imaging and photothermal treatment of cancer without significant in vitro and in vivo toxicity.

4.8. Catalytic Performance

The unique optical characteristics and heterojunction structure of MXene are conducive to improving the photocatalytic performance of MXene, which plays a crucial role in the application of MXene in the field of catalysis. Heterogeneous structures terminated by van der Waals (vdW) forces are important for photocatalysis [114]. The heterostructures are classified as type I, type II, and type III, which represent the crossover gap, interleaved gap, and fragmentation gap, respectively. MXene and transition metal dichalogenates can form a variety of heterostructures, such as MoS2/Ti2CF2 [115], graphene/Ti2C [116], and so on [117]. Scientists found that MXene heterostructure can enhance the photocatalytic performance of hydrogen evolution reaction (HER) [118,119], N2 reduction reaction, water splitting, and photocatalytic CO2 reduction. Cao et al. found that Ti3C2/Bi2WO6 produced nearly five times more CH4 and CH3OH than Bi2WO6 [120].
ROS plays a decisive role when considering the photocatalytic degradation process [121,122]. In this regard, titanium dioxide (TiO2), which is of course very much related to Ti-based MXene and MBenes, offers excellent photocatalytic properties, as well as low cost and low toxicity. When TiO2 is used in a photocatalytic process, the absorbed energy induces an electron-hole pair due to electron transfer from the valence guide band [123,124]. Their interaction with surrounding water or oxygen induces the formation of ROS. In addition to the important effects of TiO2 particle size, size, surface area, and crystallinity, the photocatalytic activity is also quite sensitive to the amount of ROS generated [125,126]. This has also confirmed that TiO2’s material properties, such as crystal phase and size, as well as defect density, significantly affect the formation of ROS, thus affecting photocatalytic activity [126,127].
The important effect of ROS on the photocatalytic performance of MXene nanosheets has also recently been demonstrated [128]. Rosales et al. verified by XPS that a small amount of active TiO2 sites on the surface of Ti3C2 nanosheets could form electron/hole pairs under UVA irradiation [129]. Electrons undergo a transfer from the TiO2 lead to the surface of the nanosheet, thereby helping to separate light-generated charge carriers. The existing surface terminations and the excellent electrical conductivity of MXene reduce the possibility of recombination, thus significantly facilitating the photocatalytic process. It can be hypothesized that complex redox reactions (-OH surface terminations and oxygen or water) are likely to induce ROS generation. Due to the increased TiO2 content on the surface of monolayer Ti3C2 nanosheets compared with multilayer nanosheets, the monolayer Ti3C2 nanosheets showed higher ROS generation capacity [130].
Gogotsi found that Ti3C2Tx degraded 81% of MB and 62% of AB80 in wastewater within 5 h of UV irradiation, while only four times MB was degraded in the dark. This study is the first to explore the application of MXene in photodegradation, but does not explain this complex process [131]. MXene is unstable in aquatic environments, and Ti3C2Tx will be oxidized to TiO2 to enhance photodegradation performance [132]. In 2020, scientists proposed that the cleavage process of HgClOH adsorbed on Ti3C2 would produce hydroxyl radical (·OH), which could promote the conversion of Ti3C2 to TiO2/C nanocomposites. The nanocomposites show good photocatalytic performance. This distinctive mechanism is applicable for mercury (II) removal [105].
Since the catalytic performance mainly depends on the instability of MXene, it is still a major challenge to improve the reversible and cyclic performance of MXene. In addition, further studies are needed to investigate the performance mechanism of MXene in the catalytic region. In practice, it is still inconvenient to use and recycle low-dimensional materials. How to obtain the macroscopic MXene materials and how to effectively separate MXene materials remain to be studied.

4.9. Others

In addition to the above properties, the influence of surface chemistry on the properties of MXene is also reflected in many other fields due to the unique structure of MXene which affords excellent physicochemical properties. For example, MXene is easily oxidized in air or water, which reduces its physical and chemical properties, thus greatly limiting its application scope [133]. Surface functionalization of Ti3C2 MXene sheets by various organic ligands has been performed in various ways, including catechol [134], silane [135,136], isocyanate [137,138], amino acid [139,140], and phosphonic acid [30]. Although these methods have been successful, there are still many limitations in the oxidization of MXene. For example, the attachment of amino acids requires an aqueous phase synthesis route, which may result in the oxidation of Ti3C2 sheets to TiO2 due to the presence of dissolved oxygen or water itself. The surface terminations of MXene can be tailored toward certain redox chemistries. Based on the aforementioned characteristics, modified MXene with high electronic/ionic transport is becoming a promising electrode material [30].
MXene-based materials are used to develop sensors for gas detection, strain, pressure, and humidity currently. The main working mechanism behind all of these MXene-based sensors is the change in the electrical conductivity of external stimuli. However, although a large number of studies have shown that MXene is a potential two-dimensional material, there are still many challenges to fully apply MXene in sensing, such as difficulty to achieving green preparation of materials, insufficient stability, and in complete control of surface functional groups [141,142]. However, some studies have pointed out that cetyl trimethyl ammonium bromide (CTAB) or octadecyl trimethyl ammonium bromide (STAB) can be used to change the surface terminations of Ti3C2Tx, and the interlayer spacing can be adjusted [30]. This adjustable interlayer spacing makes it more prominent in applications such as gas sensing and gas separation.
In addition, Khaledialidusti et al. compared the M-B, B-B, and M-M bond strengths of MBene and proves that the B-B is the strongest bond, while M-M tends to be the weakest one. This tendency also holds true for bare and terminated MXene with the order of B-B > M-O > M-F ≈ M-OH > M-B > M-M [46,62,143]. We speculate that the existence of surface terminations on the outer MBenes’ surface and between the layers is inevitable. Considering the available knowledge about MXenes’ surface terminations, we anticipate that similar surface terminations including -O, -OH, -F, and -Cl groups will be present for MBenes. Nair et al. summarized the electronic properties of various surface-terminated MBenes [50,144]. It indicates that the electronic properties of MBenes are affected by surface terminations. The same property may be applied to MXene.
The future applications of MXene are expanding and have the potential to change the world. To truly realize the potential of the entire MXene family, basic research on these necessary properties will be required in addition to clarifying the effects of structure and composition on MXene [145,146,147].

5. Conclusions and Perspective

This account reviews these investigations of the MXene surface terminations, and provides fundamental information on regulation methods, characterization techniques, and the relevance among terminations, properties, and applications.
Moreover, some issues were proposed as research emphasis and difficulties of MXene surface terminations.
(I)
How to observe these surface terminations of MXene at the atomic level? Employing spherical aberration-corrected electron microscopy combined with advanced spectral characterization techniques may offer new sights for MXene’s surface studies. Through cross-section sample preparation, the interlayer of MXene can be observed.
(II)
How to prepare the bare Mxene without surface terminations? It is required for understanding the detailed bonding nature between surface species of bare MXene, the regulation rule between surface metal species and the surface functional groups, and the interfacial interaction of MXene and hybrid phases.
(III)
And how to construct the unsymmetrical MXene surface terminations on both sides of the single MXene nanosheets? There are key issues to improve to study the physical properties of MXene and support an accurate model for theoretical calculation and simulation.
(IV)
Developing MXene with a new organo-functional group and analyzing the interface between MXene and other hybrids.
To address these issues, devoting to understanding the MXene surface terminations is the route to improving the MXene’s function to practical application.

Author Contributions

Conceptualization, M.L. and W.Z.; Methodology, S.X.; formal analysis, H.L. and W.H.; writing—original draft preparation, M.T., M.L. and H.L.; writing—review and editing, M.T., M.L. and W.Z.; supervision, W.Z., J.L. and Y.W. All authors have read and agreed to the published version of the manuscript.

Funding

The authors gratefully acknowledge the financial support provided by the Development Plan of Science and Technology of Jilin Province (YDZJ202201ZYTS305), the Science and Technology Research Project of the Education Department of Jilin Province (JJKH20210453KJ).

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Hu, M.; Zhang, H.; Hu, T.; Fan, B.; Wang, X.; Li, Z. Emerging 2D MXenes for Supercapacitors: Status, Challenges and Prospects. Chem. Soc. Rev. 2020, 49, 6666–6693. [Google Scholar] [CrossRef] [PubMed]
  2. Naguib, M.; Kurtoglu, M.; Presser, V.; Lu, J.; Niu, J.; Heon, M.; Hultman, L.; Gogotsi, Y.; Barsoum, M.W. Two-Dimensional Nanocrystals Produced by Exfoliation of Ti3AlC2. Adv. Mater. 2011, 23, 4248–4253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Mendoza-Sánchez, B.; Gogotsi, Y. Synthesis of Two-Dimensional Materials for Capacitive Energy Storage. Adv. Mater. 2016, 28, 6104–6135. [Google Scholar] [CrossRef]
  4. Wan, Y.J.; Rajavel, K.; Li, X.M.; Wang, X.Y.; Liao, S.Y.; Lin, Z.Q.; Zhu, P.L.; Sun, R.; Wong, C.P. Electromagnetic Interference Shielding of Ti3C2Tx MXene Modified by Ionic Liquid for High Chemical Stability and Excellent Mechanical Strength. Chem. Eng. J. 2021, 408, 127303. [Google Scholar] [CrossRef]
  5. Wu, X.; Cui, X.; Wu, W.; Wang, J.; Li, Y.; Jiang, Z. Elucidating Ultrafast Molecular Permeation through Well-Defined 2D Nanochannels of Lamellar Membranes. Angew. Chem. Int. Ed. 2019, 131, 18695–18700. [Google Scholar] [CrossRef]
  6. Chen, J.; Huang, Q.; Huang, H.; Mao, L.; Liu, M.; Zhang, X.; Wei, Y. Recent Progress and Advances in the Environmental Applications of MXene Related Materials. Nanoscale 2020, 12, 3574–3592. [Google Scholar] [CrossRef]
  7. Magnuson, M.; Halim, J.; Näslund, L. Chemical Bonding in Carbide MXene Nanosheets. J. Electron. Spectrosc. Relat. Phenom. 2018, 224, 27–32. [Google Scholar] [CrossRef] [Green Version]
  8. Han, X.; Huang, J.; Lin, H.; Wang, Z.; Li, P.; Chen, Y. 2D Ultrathin MXene-Based Drug-Delivery Nanoplatform for Synergistic Photothermal Ablation and Chemotherapy of Cancer. Adv. Healthc. Mater. 2018, 7, 1701394. [Google Scholar] [CrossRef]
  9. Ouyang, H.; Zhou, M.; Guo, Y.; He, M.; Huang, H.; Ye, X.; Feng, Y.; Zhou, X.; Yang, S. Metabolites Profiling of Pulsatilla Saponin D in Rat by Ultra Performance Liquid Chromatography—Quadrupole Time-of-Flight Mass Spectrometry (UPLC/Q-TOF-MS/MS). Fitoterapia 2014, 96, 152–158. [Google Scholar] [CrossRef]
  10. Zhang, X.; Li, J.; Xie, B.; Wu, B.; Lei, S.; Yao, Y.; He, M.; Ouyang, H.; Feng, Y.; Xu, W.; et al. Comparative Metabolomics Analysis of Cervicitis in Human Patients and a Phenol Mucilage-Induced Rat Model Using Liquid Chromatography Tandem Mass Spectrometry. Front. Pharmacol. 2018, 9, 00282. [Google Scholar] [CrossRef]
  11. Zhang, X.; He, M.; Lei, S.; Wu, B.; Tan, T.; Ouyang, H.; Xu, W.; Feng, Y. An Integrative Investigation of the Therapeutic Mechanism of Ainsliaea Fragrans Champ. in Cervicitis Using Liquid Chromatography Tandem Mass Spectrometry Based on a Rat Plasma Metabolomics Strategy. J. Pharm. Biomed. Anal. 2018, 156, 221–231. [Google Scholar] [CrossRef] [PubMed]
  12. Xie, Y.; Naguib, M.; Mochalin, V.N.; Barsoum, M.W.; Gogotsi, Y.; Yu, X.; Nam, K.W.; Yang, X.Q.; Kolesnikov, A.I.; Kent, P.R.C. Role of Surface Structure on Li-Ion Energy Storage Capacity of Two-Dimensional Transition-Metal Carbides. J. Am. Chem. Soc. 2014, 136, 6385–6394. [Google Scholar] [CrossRef] [PubMed]
  13. Wang, J.; Ma, X.; Zhou, J.; Du, F.; Teng, C. Bioinspired, High-Strength, and Flexible MXene/Aramid Fiber for Electromagnetic Interference Shielding Papers with Joule Heating Performance. ACS Nano 2022, 16, 6700–6711. [Google Scholar] [CrossRef] [PubMed]
  14. Parra-Muñoz, N.; Soler, M.; Rosenkranz, A. Covalent Functionalization of MXenes for Tribological Purposes—A Critical Review. Adv. Colloid Interface Sci. 2022, 309, 102792. [Google Scholar] [CrossRef]
  15. VahidMohammadi, A.; Rosen, J.; Gogotsi, Y. The World of Two-Dimensional Carbides and Nitrides (MXenes). Science 2021, 372, eabf1581. [Google Scholar] [CrossRef]
  16. Gogotsi, Y.; Anasori, B. The Rise of MXenes. ACS Nano 2019, 13, 8491–8494. [Google Scholar] [CrossRef] [Green Version]
  17. Tang, J.; Mathis, T.S.; Kurra, N.; Sarycheva, A.; Xiao, X.; Hedhili, M.N.; Jiang, Q.; Alshareef, H.N.; Xu, B.; Pan, F.; et al. Tuning the Electrochemical Performance of Titanium Carbide MXene by Controllable in Situ Anodic Oxidation. Angew. Chem. Int. Ed. 2019, 58, 17849–17855. [Google Scholar] [CrossRef] [Green Version]
  18. Li, K.; Liang, M.; Wang, H.; Wang, X.; Huang, Y.; Coelho, J.; Pinilla, S.; Zhang, Y.; Qi, F.; Nicolosi, V.; et al. 3D MXene Architectures for Efficient Energy Storage and Conversion. Adv. Funct. Mater. 2020, 30, 2000842. [Google Scholar] [CrossRef]
  19. Luo, Y.; Que, W.; Bin, X.; Xia, C.; Kong, B.; Gao, B.; Kong, L.B. Flexible MXene—Based Composite Films: Synthesis, Modification, and Applications as Electrodes of Supercapacitors. Small 2022, 18, 2201290. [Google Scholar] [CrossRef]
  20. Malaki, M.; Varma, R.S. Mechanotribological Aspects of MXene-Reinforced Nanocomposites. Adv. Mater. 2020, 32, 2003154. [Google Scholar] [CrossRef]
  21. Wu, Z.; Liu, X.; Shang, T.; Deng, Y.; Wang, N.; Dong, X.; Zhao, J.; Chen, D.; Tao, Y.; Yang, Q. Reassembly of MXene Hydrogels into Flexible Films towards Compact and Ultrafast Supercapacitors. Adv. Funct. Mater. 2021, 31, 2102874. [Google Scholar] [CrossRef]
  22. Yang, X.; Yao, Y.; Wang, Q.; Zhu, K.; Ye, K.; Wang, G.; Cao, D.; Yan, J. 3D Macroporous Oxidation-Resistant Ti3C2Tx MXene Hybrid Hydrogels for Enhanced Supercapacitive Performances with Ultralong Cycle Life. Adv. Funct. Mater. 2022, 32, 2109479. [Google Scholar]
  23. Zhang, X.; Lv, R.; Wang, A.; Guo, W.; Liu, X.; Luo, J. MXene Aerogel Scaffolds for High-Rate Lithium Metal Anodes. Angew. Chem. Int. Ed. 2018, 57, 15028–15033. [Google Scholar] [CrossRef] [PubMed]
  24. Wang, Y.; Feng, W.; Chen, Y. Chemistry of Two-Dimensional MXene Nanosheets in Theranostic Nanomedicine. Chin. Chem. Lett. 2020, 31, 937–946. [Google Scholar] [CrossRef]
  25. 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]
  26. Sang, X.; Xie, Y.; Lin, M.W.; Alhabeb, M.; van Aken, K.L.; Gogotsi, Y.; Kent, P.R.C.; Xiao, K.; Unocic, R.R. Atomic Defects in Monolayer Titanium Carbide (Ti3C2Tx) MXene. ACS Nano 2016, 10, 9193–9200. [Google Scholar] [CrossRef]
  27. Li, M.; Lu, J.; Luo, K.; Li, Y.; Chang, K.; Chen, K.; Zhou, J.; Rosen, J.; Hultman, L.; Eklund, P.; et al. Element Replacement Approach by Reaction with Lewis Acidic Molten Salts to Synthesize Nanolaminated MAX Phases and MXenes. J. Am. Chem. Soc. 2019, 141, 4730–4737. [Google Scholar] [CrossRef] [Green Version]
  28. Li, M.; Li, X.; Qin, G.; Luo, K.; Lu, J.; Li, Y.; Liang, G.; Huang, Z.; Zhou, J.; Hultman, L.; et al. Halogenated Ti3C2 MXenes with Electrochemically Active Terminals for High-Performance Zinc Ion Batteries. ACS Nano 2021, 15, 1077–1085. [Google Scholar] [CrossRef]
  29. Kamysbayev, V.; Filatov, A.S.; Hu, H.; Rui, X.; Lagunas, F.; Wang, D.; Klie, R.F.; Talapin, D.V. Covalent Surface Modifications and Superconductivity of Two-Dimensional Metal Carbide MXenes. Science 2020, 369, 979–983. [Google Scholar] [CrossRef]
  30. Kim, D.; Ko, T.Y.; Kim, H.; Lee, G.H.; Cho, S.; Koo, C.M. Nonpolar Organic Dispersion of 2D Ti3C2Tx MXene Flakes via Simultaneous Interfacial Chemical Grafting and Phase Transfer Method. ACS Nano 2019, 13, 13818–13828. [Google Scholar] [CrossRef]
  31. Zhang, S.; Zhuo, H.; Li, S.; Bao, Z.; Deng, S.; Zhuang, G.; Zhong, X.; Wei, Z.; Yao, Z.; Wang, J. Effects of Surface Functionalization of Mxene-Based Nanocatalysts on Hydrogen Evolution Reaction Performance. Catal. Today 2021, 368, 187–195. [Google Scholar] [CrossRef]
  32. Wang, S.; Liu, Y.; Lü, Q.F.; Zhuang, H. Facile Preparation of Biosurfactant-Functionalized Ti2CTx MXene Nanosheets with an Enhanced Adsorption Performance for Pb(II) Ions. J. Mol. Liq. 2020, 297, 111810. [Google Scholar] [CrossRef]
  33. Rajan, A.C.; Mishra, A.; Satsangi, S.; Vaish, R.; Mizuseki, H.; Lee, K.R.; Singh, A.K. Machine-Learning-Assisted Accurate Band Gap Predictions of Functionalized Mxene. Chem. Mater. 2018, 30, 4031–4038. [Google Scholar] [CrossRef]
  34. 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]
  35. Yang, J.; Naguib, M.; Ghidiu, M.; Pan, L.M.; Gu, J.; Nanda, J.; Halim, J.; Gogotsi, Y.; Barsoum, M.W. Two-Dimensional Nb-Based M4C3 Solid Solutions (MXenes). J. Am. Ceram. Soc. 2016, 99, 660–666. [Google Scholar] [CrossRef]
  36. 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]
  37. Liu, F.; Zhou, A.; Chen, J.; Jia, J.; Zhou, W.; Wang, L.; Hu, Q. Preparation of Ti3C2 and Ti2C MXenes by Fluoride Salts Etching and Methane Adsorptive Properties. Appl. Surf. Sci. 2017, 416, 781–789. [Google Scholar] [CrossRef]
  38. Lee, J.T.; Wyatt, B.C.; Davis, G.A.; Masterson, A.N.; Pagan, A.L.; Shah, A.; Anasori, B.; Sardar, R. Covalent Surface Modification of Ti3C2Tx MXene with Chemically Active Polymeric Ligands Producing Highly Conductive and Ordered Microstructure Films. ACS Nano 2021, 15, 19600–19612. [Google Scholar] [CrossRef]
  39. Guo, M.; Geng, W.C.; Liu, C.; Gu, J.; Zhang, Z.; Tang, Y. Ultrahigh Areal Capacitance of Flexible MXene Electrodes: Electrostatic and Steric Effects of Terminations. Chem. Mater. 2020, 32, 8257–8265. [Google Scholar] [CrossRef]
  40. Zou, X.; Liu, H.; Xu, H.; Wu, X.; Han, X.; Kang, J.; Reddy, K.M. A Simple Approach to Synthesis Cr2CTx MXene for Efficient Hydrogen Evolution Reaction. Mater. Today Energy 2021, 20, 100668. [Google Scholar] [CrossRef]
  41. Qiu, S.Y.; Wang, C.; Jiang, Z.X.; Zhang, L.S.; Gu, L.L.; Wang, K.X.; Gao, J.; Zhu, X.D.; Wu, G. Rational Design of MXene@TiO2 nanoarray Enabling Dual Lithium Polysulfide Chemisorption towards High-Performance Lithium-Sulfur Batteries. Nanoscale 2020, 12, 16678–16684. [Google Scholar] [CrossRef] [PubMed]
  42. Biggin, S.; Enderby, J.E. The Structure of Molten Zinc Chloride. J. Phys. C Solid State Phys. 1981, 14, 3129–3136. [Google Scholar] [CrossRef]
  43. Hefeng, L.; Kunquan, L.; Zhonghua, W.; Jun, D. EXAFS Studies of Molten EXAFS Studies of Molten ZnCl2, RbCl and Rb2ZnCl4. J. Phys. Condens. Matter 1994, 6, 3629. [Google Scholar] [CrossRef]
  44. Shen, M.; Jiang, W.; Liang, K.; Zhao, S.; Tang, R.; Zhang, L.; Wang, J. One-Pot Green Process to Synthesize MXene with Controllable Surface Terminations Using Molten Salts. Angew. Chem. Int. Ed. 2021, 60, 27013–27018. [Google Scholar] [CrossRef]
  45. Xiao, X.; Song, H.; Lin, S.; Zhou, Y.; Zhan, X.; Hu, Z.; Zhang, Q.; Sun, J.; Yang, B.; Li, T.; et al. Scalable Salt-Templated Synthesis of Two-Dimensional Transition Metal Oxides. Nat. Commun. 2016, 7, 11296. [Google Scholar] [CrossRef] [Green Version]
  46. Khaledialidusti, R.; Khazaei, M.V.; Wang, N.; Miao, C.; Si, J.; Wang, J.; Wang, J. Exploring structural, electronic, and mechanical properties of 2D hexagonal MBenes. J. Phys. Condens. Matter 2021, 33, 155503. [Google Scholar] [CrossRef]
  47. Weng, W.; Jiang, B.; Wang, Z.; Xiao, W. In Situ Electrochemical Conversion of CO2 in Molten Salts to Advanced Energy Materials with Reduced Carbon Emissions. Sci. Adv. 2020, 6, eaay9278. [Google Scholar] [CrossRef] [Green Version]
  48. Liang, X.; Xiao, J.; Weng, W.; Xiao, W. Electrochemical Reduction of Carbon Dioxide and Iron Oxide in Molten Salts to Fe/Fe3C Modified Carbon for Electrocatalytic Oxygen Evolution. Angew. Chem. Int. Ed. 2021, 60, 2120–2124. [Google Scholar] [CrossRef]
  49. Peng, M.; Dong, M.; Wei, W.; Xu, H.; Liu, C.; Shen, C. The Introduction of Amino Termination on Ti3C2 MXene Surface for Its Flexible Film with Excellent Property. Carbon 2021, 179, 400–407. [Google Scholar] [CrossRef]
  50. Nair, V.G.; Birowska, M.; Bury, D.; Jakubczak, M.; Rosenkranz, A.; Jastrzębska, A.M. 2D MBenes: A Novel Member in the Flatland. Adv. Mater. 2022, 34, 2108840. [Google Scholar] [CrossRef]
  51. Lu, C.; Yang, L.; Yan, B.; Sun, L.; Zhang, P.; Zhang, W.; Sun, Z.M. Nitrogen-Doped Ti3C2 MXene: Mechanism Investigation and Electrochemical Analysis. Adv. Funct. Mater. 2020, 30, 2000852. [Google Scholar] [CrossRef]
  52. Michałowski, P.P.; Anayee, M.; Mathis, T.S.; Kozdra, S.; Wójcik, A.; Hantanasirisakul, K.; Jóźwik, I.; Piątkowska, A.; Możdżonek, M.; Malinowska, A.; et al. Oxycarbide MXenes and MAX Phases Identification Using Monoatomic Layer-by-Layer Analysis with Ultralow-Energy Secondary-Ion Mass Spectrometry. Nat. Nanotechnol. 2022. [Google Scholar] [CrossRef]
  53. Zhang, Y.; Chen, K.; Li, Y.; Lan, J.; Yan, B.; Shi, L.; Ran, R. High-Strength, Self-Healable, Temperature-Sensitive, MXene-Containing Composite Hydrogel as a Smart Compression Sensor. ACS Appl. Mater. Inter. 2019, 11, 47350–47357. [Google Scholar] [CrossRef]
  54. Chao, M.; He, L.; Gong, M.; Li, N.; Li, X.; Peng, L.; Shi, F.; Zhang, L.; Wan, P. Breathable Ti3C2Tx MXene/Protein Nanocomposites for Ultrasensitive Medical Pressure Sensor with Degradability in Solvents. ACS Nano 2021, 15, 9746–9758. [Google Scholar] [CrossRef] [PubMed]
  55. Wu, Q.; Li, N.; Wang, Y.; Xu, Y.; Wu, J.; Jia, G.; Ji, F.; Fang, X.; Chen, F.; Cui, X. Ultrasensitive and Selective Determination of Carcinoembryonic Antigen Using Multifunctional Ultrathin Amino-Functionalized Ti3C2-MXene Nanosheets. Anal. Chem. 2020, 92, 3354–3360. [Google Scholar] [CrossRef] [PubMed]
  56. Halim, J.; Cook, K.M.; Naguib, M.; Eklund, P.; Gogotsi, Y.; Rosen, J.; Barsoum, M.W. X-Ray Photoelectron Spectroscopy of Select Multi-Layered Transition Metal Carbides (MXenes). Appl. Surf. Sci. 2016, 362, 406–417. [Google Scholar] [CrossRef] [Green Version]
  57. Schier, V.; Michel, H.-J.; Halbritter, J. ARXPS-Analysis of Sputtered TiC, SiC and Ti0.5Si0.5C Layers. Fresenius J. Anal. Chem. 1993, 346, 227–232. [Google Scholar] [CrossRef]
  58. Magni, E.; Somorjai, G. Preparation of a Model Ziegler-Natta Catalyst: Electron Irradiation Induced Titanium Chloride Deposition on Magnesium Chloride Thin Films Grown on Gold. Surf. Sci. 1996, 345, 1–16. [Google Scholar] [CrossRef]
  59. Mousty-Desbuquoit, C.; Riga, J.; Verbist, J.J. Solid State Effects in the Electronic Structure of TiCl4 Studied by XPS. J. Chem. Phys. 1983, 79, 26–32. [Google Scholar] [CrossRef]
  60. Saha, N.C.; Tompkins, H.G. Titanium Nitride Oxidation Chemistry: An x-ray Photoelectron Spectroscopy Study. J. Appl. Phys. 1992, 72, 3072–3079. [Google Scholar] [CrossRef]
  61. Tong, X.; Liu, S.; Zhao, Y.; Huang, L.; Crittenden, J.; Chen, Y. MXene Composite Membranes with Enhanced Ion Transport and Regulated Ion Selectivity. Environ. Sci. Technol. 2022, 56, 8964–8974. [Google Scholar] [CrossRef] [PubMed]
  62. Zhang, J.; Huang, C.; Sun, Y.; Yu, H. Amino-Functionalized Niobium-Carbide MXene Serving as Electron Transport Layer and Perovskite Additive for the Preparation of High-Performance and Stable Methylammonium-Free Perovskite Solar Cells. Adv. Funct. Mater. 2022, 32, 2113367. [Google Scholar] [CrossRef]
  63. Miao, N.; Wang, J.; Gong, Y.; Wu, J.; Niu, H.; Wang, S.; Li, K.; Oganov, A.R.; Tada, T.; Hosono, H. Computational Prediction of Boron-Based MAX Phases and MXene Derivatives. Chem. Mater. 2020, 32, 6947. [Google Scholar] [CrossRef]
  64. Karlsson, L.H.; Birch, J.; Halim, J.; Barsoum, M.W.; Persson, P.O.A. Atomically Resolved Structural and Chemical Investigation of Single MXene Sheets. Nano Lett. 2015, 15, 4955–4960. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. Xie, Y.; Kent, P.R.C. Hybrid Density Functional Study of Structural and Electronic Properties of Functionalized Tin+1Xn (X=C, N) Monolayers. Phys. Rev. B Matter Mater. Phys. 2013, 87, 235441. [Google Scholar] [CrossRef] [Green Version]
  66. Kim, H.; Wang, Z.; Alshareef, H.N. MXetronics: Electronic and Photonic Applications of MXenes. Nano Energy 2019, 60, 179–197. [Google Scholar] [CrossRef]
  67. Khazaei, M.; Arai, M.; Sasaki, T.; Chung, C.Y.; Venkataramanan, N.S.; Estili, M.; Sakka, Y.; Kawazoe, Y. Novel Electronic and Magnetic Properties of Two-Dimensional Transition Metal Carbides and Nitrides. Adv. Funct. Mater. 2012, 23, 2185–2192. [Google Scholar] [CrossRef]
  68. Kumar, S.; Schwingenschlögl, U. Thermoelectric Performance of Functionalized Sc2C MXenes. Phys. Rev. B 2016, 94, 035405. [Google Scholar] [CrossRef] [Green Version]
  69. Lee, Y.; Hwang, Y.; Chung, Y.C. Achieving Type-I, II, and III Heterojunction Using Functionalized MXene. ACS Appl. Mater. Inter. 2015, 7, 7163–7169. [Google Scholar] [CrossRef]
  70. Liu, J.H.; Kan, X.; Amin, B.; Gan, L.Y.; Zhao, Y. Theoretical Exploration of the Potential Applications of Sc-Based MXenes. Phys. Chem. Chem. Phys. 2017, 19, 32253–32261. [Google Scholar] [CrossRef]
  71. Chandrasekaran, A.; Mishra, A.; Singh, A.K. Ferroelectricity, Antiferroelectricity, and Ultrathin 2D Electron/Hole Gas in Multifunctional Monolayer MXene. Nano Lett. 2017, 17, 3290–3296. [Google Scholar] [CrossRef] [PubMed]
  72. Khazaei, M.; Arai, M.; Sasaki, T.; Estili, M.; Sakka, Y. Two-Dimensional Molybdenum Carbides: Potential Thermoelectric Materials of the MXene Family. Phys. Chem. Chem. Phys. 2014, 16, 7841–7849. [Google Scholar] [CrossRef] [PubMed]
  73. Ma, Z.; Hu, Z.; Zhao, X.; Tang, Q.; Wu, D.; Zhou, Z.; Zhang, L. Tunable Band Structures of Heterostructured Bilayers with Transition-Metal Dichalcogenide and MXene Monolayer. J. Phys. Chem. C 2014, 118, 5593–5599. [Google Scholar] [CrossRef]
  74. Guo, J.; Sun, Y.; Liu, B.; Zhang, Q.; Peng, Q. Two-Dimensional Scandium-Based Carbides (MXene): Band Gap Modulation and Optical Properties. J. Alloys Compd. 2017, 712, 752–759. [Google Scholar] [CrossRef]
  75. Gandi, A.N.; Alshareef, H.N.; Schwingenschlögl, U. Thermoelectric Performance of the MXenes M2CO2 (M=Ti, Zr, or Hf). Chem. Mater. 2016, 28, 1647–1652. [Google Scholar] [CrossRef] [Green Version]
  76. Gao, G.; O’Mullane, A.P.; Du, A. 2D MXenes: A New Family of Promising Catalysts for the Hydrogen Evolution Reaction. ACS Catal. 2017, 7, 494–500. [Google Scholar] [CrossRef] [Green Version]
  77. Zhang, H.; Yang, G.; Zuo, X.; Tang, H.; Yang, Q.; Li, G. Computational Studies on the Structural, Electronic and Optical Properties of Graphene-like MXenes (M2CT2, M=Ti, Zr, Hf; T=O, F, OH) and Their Potential Applications as Visible-Light Driven Photocatalysts. J. Mater. Chem. A 2016, 4, 12913–12920. [Google Scholar] [CrossRef]
  78. Zha, X.H.; Huang, Q.; He, J.; He, H.; Zhai, J.; Francisco, J.S.; Du, S. The Thermal and Electrical Properties of the Promising Semiconductor MXene Hf2CO2. Sci. Rep. 2016, 6, 27971. [Google Scholar] [CrossRef] [Green Version]
  79. Yang, J.; Zhou, X.; Luo, X.; Zhang, S.; Chen, L. Tunable Electronic and Magnetic Properties of Cr2M′C2T2 (M’=Ti or V.; T=O, OH or F). Appl. Phys. Lett. 2016, 109, 203109. [Google Scholar] [CrossRef]
  80. Hu, J.; Xu, B.; Ouyang, C.; Yang, S.A.; Yao, Y. Investigations on V2C and V2CX2 (X=F, OH) Monolayer as a Promising Anode Material for Li Ion Batteries from First-Principles Calculations. J. Phys. Chem. C 2014, 118, 24274–24281. [Google Scholar] [CrossRef]
  81. Si, C.; Zhou, J.; Sun, Z. Half-Metallic Ferromagnetism and Surface Functionalization-Induced Metal-Insulator Transition in Graphene-like Two-Dimensional Cr2C Crystals. ACS Appl. Mater. Inter. 2015, 7, 17510–17515. [Google Scholar] [CrossRef] [PubMed]
  82. Weng, H.; Ranjbar, A.; Liang, Y.; Song, Z.; Khazaei, M.; Yunoki, S.; Arai, M.; Kawazoe, Y.; Fang, Z.; Dai, X. Large-Gap Two-Dimensional Topological Insulator in Oxygen Functionalized MXene. Phys. Rev. B 2015, 92, 075436. [Google Scholar] [CrossRef]
  83. Khazaei, M.; Wang, V.; Sevik, C.; Ranjbar, A.; Arai, M.; Yunoki, S. Electronic Structures of IMAX Phases and Their Two-Dimensional Derivatives: A Family of Piezoelectric Materials. Phys. Rev. Mater. 2018, 2, 074002. [Google Scholar] [CrossRef] [Green Version]
  84. Lind, H.; Halim, J.; Simak, S.I.; Rosen, J. Investigation of Vacancy-Ordered Mo1.33C MXene from First Principles and X-ray Photoelectron Spectroscopy. Phys. Rev. Mater. 2017, 1, 044002. [Google Scholar] [CrossRef]
  85. Zha, X.H.; Zhou, J.; Luo, K.; Lang, J.; Huang, Q.; Zhou, X.; Francisco, J.S.; He, J.; Du, S. Controllable Magnitude and Anisotropy of the Electrical Conductivity of Hf3C2O2 MXene. J. Phys. Condens. Matter 2017, 29, 165701. [Google Scholar] [CrossRef]
  86. Dong, L.; Kumar, H.; Anasori, B.; Gogotsi, Y.; Shenoy, V.B. Rational Design of Two-Dimensional Metallic and Semiconducting Spintronic Materials Based on Ordered Double-Transition-Metal MXenes. J. Phys. Chem. Lett. 2017, 8, 422–428. [Google Scholar] [CrossRef]
  87. Zhang, J.J.; Lin, L.; Zhang, Y.; Wu, M.; Yakobson, B.I.; Dong, S. Type-II Multiferroic Hf2VC2F2 MXene Monolayer with High Transition Temperature. J. Am. Chem. Soc. 2018, 140, 9768–9773. [Google Scholar] [CrossRef] [Green Version]
  88. Khazaei, M.; Ranjbar, A.; Arai, M.; Yunoki, S. Topological Insulators in the Ordered Double Transition Metals M2′M″C2 MXenes (M’= Mo, W.; M”=Ti, Zr, Hf). Phys. Rev. B 2016, 94, 125152. [Google Scholar] [CrossRef] [Green Version]
  89. Si, C.; Jin, K.H.; Zhou, J.; Sun, Z.; Liu, F. Large-Gap Quantum Spin Hall State in MXenes: D-Band Topological Order in a Triangular Lattice. Nano Lett. 2016, 16, 6584–6591. [Google Scholar] [CrossRef]
  90. Wang, H.; Yuan, H.; Yu, J.; Zhang, C.; Li, K.; You, M.; Li, W.; Shao, J.; Wei, J.; Zhang, X.; et al. Boosting the Efficiency of NiOx—Based Perovskite Light-Emitting Diodes by Interface Engineering. ACS Appl. Mater. Inter. 2020, 12, 53528–53536. [Google Scholar] [CrossRef]
  91. Yin, L.; Li, Y.; Yao, X.; Wang, Y.; Jia, L.; Liu, Q.; Li, J.; Li, Y.; He, D. MXenes for Solar Cells. Nano-Micro Lett. 2021, 13, 78. [Google Scholar] [CrossRef] [PubMed]
  92. Sun, Y.; Meng, X.; Dall’Agnese, Y.; Dall’Agnese, C.; Duan, S.; Gao, Y.; Chen, G.; Wang, X.F. 2D MXenes as Co-Catalysts in Photocatalysis: Synthetic Methods. Nano-Micro Lett. 2019, 11, 79. [Google Scholar] [CrossRef] [PubMed]
  93. Khazaei, M.; Mishra, A.; Venkataramanan, N.S.; Singh, A.K.; Yunoki, S. Recent Advances in MXenes: From Fundamentals to Applications. Curr. Opin. Solid St. M. 2019, 23, 164–178. [Google Scholar] [CrossRef] [Green Version]
  94. Naguib, M.; Mochalin, V.N.; Barsoum, M.W.; Gogotsi, Y. 25th Anniversary Article: MXenes: A New Family of Two-Dimensional Materials. Adv. Mater. 2014, 26, 992–1005. [Google Scholar] [CrossRef] [PubMed]
  95. Khazaei, M.; Ranjbar, A.; Arai, M.; Sasaki, T.; Yunoki, S. Electronic Properties and Applications of MXenes: A Theoretical Review. J. Mater. Chem. C 2017, 5, 2488–2503. [Google Scholar] [CrossRef] [Green Version]
  96. Wang, X.; Huang, S.; Deng, L.; Luo, H.; Li, C.; Xu, Y.; Yan, Y.; Tang, Z. Enhanced Optical Absorption of Fe-, Co- and Ni- Decorated Ti3C2 MXene: A First-Principles Investigation. Phys. E 2021, 127, 114565. [Google Scholar] [CrossRef]
  97. Zhang, J.; Kong, N.; Uzun, S.; Levitt, A.; Seyedin, S.; Lynch, P.A.; Qin, S.; Han, M.; Yang, W.; Liu, J.; et al. Scalable Manufacturing of Free-Standing, Strong Ti3C2Tx MXene Films with Outstanding Conductivity. Adv. Mater. 2020, 32, 2001093. [Google Scholar]
  98. Moltved, K.A.; Kepp, K.P. The Chemical Bond between Transition Metals and Oxygen: Electronegativity, d-Orbital Effects, and Oxophilicity as Descriptors of Metal-Oxygen Interactions. J. Phys. Chem. C 2019, 123, 18432–18444. [Google Scholar] [CrossRef]
  99. Kazemi, S.A.; Wang, Y. Super Strong 2D Titanium Carbide MXene-Based Materials: A Theoretical Prediction. J. Phys.-Condens. Mat. 2019, 32, 11LT01. [Google Scholar] [CrossRef]
  100. Fu, Z.; Zhang, H.; Si, C.; Legut, D.; Germann, T.C.; Zhang, Q.; Du, S.; Francisco, J.S.; Zhang, R. Mechanistic Quantification of Thermodynamic Stability and Mechanical Strength for Two-Dimensional Transition-Metal Carbides. J. Phys. Chem. C 2018, 122, 4710–4722. [Google Scholar] [CrossRef]
  101. Hu, T.; Yang, J.; Li, W.; Wang, X.; Li, C.M. Quantifying the Rigidity of 2D Carbides (MXenes). Phys. Chem. Chem. Phys. 2020, 22, 2115–2121. [Google Scholar] [CrossRef] [PubMed]
  102. Wyatt, B.C.; Rosenkranz, A.; Anasori, B. 2D MXenes: Tunable Mechanical and Tribological Properties. Adv. Mater. 2021, 33, 2007973. [Google Scholar] [CrossRef] [PubMed]
  103. Yan, H.; Cai, M.; Li, W.; Fan, X.; Zhu, M. Amino-Functionalized Ti3C2Tx with Anti-Corrosive/Wear Function for Waterborne Epoxy Coating. J. Mater. Sci. Technol. 2020, 54, 144–159. [Google Scholar] [CrossRef]
  104. Karahan, H.E.; Goh, K.; Zhang, C.; Yang, E.; Yıldırım, C.; Chuah, C.Y.; Ahunbay, M.G.; Lee, J.; Tantekin-Ersolmaz, B.; Chen, Y.; et al. MXene Materials for Designing Advanced Separation Membranes. Adv. Mater. 2020, 32, 1906697. [Google Scholar] [CrossRef]
  105. Yu, F.; Zhang, X.; Yang, Z.; Yang, P.; Ma, J. Environmental Applications of Two-Dimensional Transition Metal Carbides and Nitrides for Water Purification: A Review. Environ. Chem. Lett. 2022, 20, 633–660. [Google Scholar] [CrossRef]
  106. Jakubczak, M.; Szuplewska, A.; Rozmysłowska-Wojciechowska, A.; Rosenkranz, A.; Jastrzębska, A.M. Novel 2D MBenes—Synthesis, Structure, and Biotechnological Potential. Adv. Funct. Mater. 2021, 31, 2103048. [Google Scholar] [CrossRef]
  107. Wu, L.; Lu, X.; Dhanjai; Wu, Z.S.; Dong, Y.; Wang, X.; Zheng, S.; Chen, J. 2D Transition Metal Carbide MXene as a Robust Biosensing Platform for Enzyme Immobilization and Ultrasensitive Detection of Phenol. Biosens. Bioelectron. 2018, 107, 69–75. [Google Scholar] [CrossRef]
  108. Rozmysłowska-Wojciechowska, A.; Wojciechowski, T.; Ziemkowska, W.; Chlubny, L.; Olszyna, A.; Jastrzębska, A.M. Surface Interactions between 2D Ti3C2/Ti2C MXenes and Lysozyme. Appl. Surf. Sci. 2019, 473, 409–418. [Google Scholar] [CrossRef]
  109. Rozmysłowska-Wojciechowska, A.; Szuplewska, A.; Wojciechowski, T.; Poźniak, S.; Mitrzak, J.; Chudy, M.; Ziemkowska, W.; Chlubny, L.; Olszyna, A.; Jastrzębska, A. A Simple, Low-Cost and Green Method for Controlling the Cytotoxicity of MXenes. Mater. Sci. Eng. C 2020, 111, 110790. [Google Scholar] [CrossRef]
  110. Liu, G.; Zou, J.; Tang, Q.; Yang, X.; Zhang, Y.; Zhang, Q.; Huang, W.; Chen, P.; Shao, J.; Dong, X. Surface Modified Ti3C2 MXene Nanosheets for Tumor Targeting Photothermal/Photodynamic/Chemo Synergistic Therapy. ACS Appl. Mater. Interfaces 2017, 9, 40077–40086. [Google Scholar] [CrossRef]
  111. Szuplewska, A.; Kulpińska, D.; Dybko, A.; Jastrzębska, A.M.; Wojciechowski, T.; Rozmysłowska, A.; Chudy, M.; Grabowska-Jadach, I.; Ziemkowska, W.; Brzózka, Z.; et al. 2D Ti2C (MXene) as a Novel Highly Efficient and Selective Agent for Photothermal Therapy. Mater. Sci. Eng. C 2019, 98, 874–886. [Google Scholar] [CrossRef] [PubMed]
  112. Lin, H.; Wang, Y.; Gao, S.; Chen, Y.; Shi, J. Theranostic 2D Tantalum Carbide (MXene). Adv. Mater. 2017, 30, 1703284. [Google Scholar]
  113. Lin, H.; Gao, S.; Dai, C.; Chen, Y.; Shi, J. A Two-Dimensional Biodegradable Niobium Carbide (MXene) for Photothermal Tumor Eradication in NIR-I and NIR-II Biowindows. J. Am. Chem. Soc. 2017, 139, 16235–16247. [Google Scholar] [CrossRef]
  114. Geim, A.K.; Grigorieva, I.V. Van Der Waals Heterostructures. Nature 2013, 499, 419–425. [Google Scholar] [CrossRef]
  115. Gan, L.Y.; Zhao, Y.J.; Huang, D.; Schwingenschlögl, U. First-Principles Analysis of MoS2/Ti2C and MoS2/Ti2CY2 (Y=F and OH) All-2D Semiconductor/Metal Contacts. Phys. Rev. B 2013, 87, 245307. [Google Scholar] [CrossRef] [Green Version]
  116. Paul, J.T.; Singh, A.; Dong, Z.; Zhuang, H.; Revard, B.C.; Rijal, B.; Ashton, M.; Linscheid, A.; Blonsky, M.; Gluhovic, D.; et al. Computational Methods for 2D Materials: Discovery, Property Characterization, and Application Design. J. Phys. Condens. Matter 2017, 29, 473001. [Google Scholar] [CrossRef] [PubMed]
  117. Jeon, J.; Park, Y.; Choi, S.; Lee, J.; Lim, S.S.; Lee, B.H.; Song, Y.J.; Cho, J.H.; Jang, Y.H.; Lee, S. Epitaxial Synthesis of Molybdenum Carbide and Formation of a Mo2C/MoS2 Hybrid Structure via Chemical Conversion of Molybdenum Disulfide. ACS Nano 2018, 12, 338–346. [Google Scholar] [CrossRef]
  118. Sun, Y.; Jin, D.; Sun, Y.; Meng, X.; Gao, Y.; Dall’Agnese, Y.; Chen, G.; Wang, X.F. G-C3N4/Ti3C2Tx (MXenes) Composite with Oxidized Surface Groups for Efficient Photocatalytic Hydrogen Evolution. J. Mater. Chem. A Mater. 2018, 6, 9124–9131. [Google Scholar] [CrossRef]
  119. Zhou, S.; Yang, X.; Pei, W.; Liu, N.; Zhao, J. Heterostructures of MXenes and N-Doped Graphene as Highly Active Bifunctional Electrocatalysts. Nanoscale 2018, 10, 10876–10883. [Google Scholar] [CrossRef] [Green Version]
  120. Cao, S.; Shen, B.; Tong, T.; Fu, J.; Yu, J. 2D/2D Heterojunction of Ultrathin MXene/Bi2WO6 Nanosheets for Improved Photocatalytic CO2 Reduction. Adv. Funct. Mater. 2018, 28, 1800136. [Google Scholar] [CrossRef]
  121. Nosaka, Y.; Nosaka, A.Y. Generation and Detection of Reactive Oxygen Species in Photocatalysis. Chem. Rev. 2017, 117, 11302–11336. [Google Scholar] [CrossRef] [PubMed]
  122. Ramírez-Sánchez, I.M.; Tuberty, S.; Hambourger, M.; Bandala, E.R. Resource Efficiency Analysis for Photocatalytic Degradation and Mineralization of Estriol Using TiO2 Nanoparticles. Chemosphere 2017, 184, 1270–1285. [Google Scholar] [CrossRef] [PubMed]
  123. Ribao, P.; Corredor, J.; Rivero, M.; Ortiz, I. Role of Reactive Oxygen Species on the Activity of Noble Metal-Doped TiO2 Photocatalysts. J. Hazard Mater. 2019, 372, 45–51. [Google Scholar] [CrossRef] [PubMed]
  124. Jiang, H.; Wang, Z.; Yang, Q.; Tan, L.; Dong, L.; Dong, M. Ultrathin Ti3C2Tx (MXene) Nanosheet-Wrapped NiSe2 Octahedral Crystal for Enhanced Supercapacitor Performance and Synergetic Electrocatalytic Water Splitting. Nanomicro Lett. 2019, 11, 31. [Google Scholar] [CrossRef] [PubMed]
  125. Rosales, M.; Zoltan, T.; Yadarola, C.; Mosquera, E.; Gracia, F.; García, A. The Influence of the Morphology of 1D TiO2 Nanostructures on Photogeneration of Reactive Oxygen Species and Enhanced Photocatalytic Activity. J. Mol. Liq. 2019, 281, 59–69. [Google Scholar] [CrossRef]
  126. Zoltan, T.; Rosales, M.; Yadarola, C. Reactive Oxygen Species Quantification and Their Correlation with the Photocatalytic Activity of TiO2 (Anatase and Rutile) Sensitized with Asymmetric Porphyrins. J. Environ. Chem. Eng. 2016, 4, 3967–3980. [Google Scholar] [CrossRef]
  127. Zhang, J.; Nosaka, Y. Mechanism of the OH Radical Generation in Photocatalysis with TiO2 of Different Crystalline Types. J. Phys. Chem. C 2014, 118, 10824–10832. [Google Scholar] [CrossRef]
  128. Tariq, A.; Ali, S.I.; Akinwande, D.; Rizwan, S. Efficient Visible-Light Photocatalysis of 2D-MXene Nanohybrids with Gd3+- and Sn4+-Codoped Bismuth Ferrite. ACS Omega 2018, 3, 13828–13836. [Google Scholar] [CrossRef] [Green Version]
  129. Ye, S.; Yan, M.; Tan, X.; Liang, J.; Zeng, G.; Wu, H.; Song, B.; Zhou, C.; Yang, Y.; Wang, H. Facile Assembled Biochar-Based Nanocomposite with Improved Graphitization for Efficient Photocatalytic Activity Driven by Visible Light. Appl. Catal. B Environ. 2019, 250, 78–88. [Google Scholar] [CrossRef]
  130. Rosales, M.; Garcia, A.; Fuenzalida, V.M.; Espinoza-González, R.; Song, G.; Wang, B.; Yu, J.; Gracia, F.; Rosenkranz, A. Unprecedented Arsenic Photo-Oxidation Behavior of Few- and Multi-Layer Ti3C2Tx Nano-Sheets. Appl. Mater. Today 2020, 20, 100769. [Google Scholar] [CrossRef]
  131. Mashtalir, O.; Cook, K.M.; Mochalin, V.N.; Crowe, M.; Barsoum, M.W.; Gogotsi, Y. Dye Adsorption and Decomposition on Two-Dimensional Titanium Carbide in Aqueous Media. J. Mater. Chem. A 2014, 2, 14334–14338. [Google Scholar] [CrossRef]
  132. Gao, Y.; Wang, L.; Zhou, A.; Li, Z.; Chen, J.; Bala, H.; Hu, Q.; Cao, X. Hydrothermal Synthesis of TiO2/Ti3C2 Nanocomposites with Enhanced Photocatalytic Activity. Mater Lett. 2015, 150, 62–64. [Google Scholar] [CrossRef]
  133. 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] [CrossRef] [PubMed]
  134. Pang, J.; Mendes, R.G.; Bachmatiuk, A.; Zhao, L.; Ta, H.Q.; Gemming, T.; Liu, H.; Liu, Z.; Rummeli, M.H. Applications of 2D MXenes in Energy Conversion and Storage Systems. Chem. Soc. Rev. 2019, 48, 72–133. [Google Scholar] [CrossRef]
  135. Kumar, S.; Lei, Y.; Alshareef, N.H.; Quevedo-Lopez, M.A.; Salama, K.N. Biofunctionalized Two-Dimensional Ti3C2 MXenes for Ultrasensitive Detection of Cancer Biomarker. Biosens. Bioelectron. 2018, 121, 243–249. [Google Scholar] [CrossRef] [Green Version]
  136. Lim, S.; Park, H.; Yang, J.; Kwak, C.; Lee, J. Stable Colloidal Dispersion of Octylated Ti3C2-MXenes in a Nonpolar Solvent. Colloids. Surf. A Physicochem. Eng. Asp. 2019, 579, 123648. [Google Scholar] [CrossRef]
  137. McDaniel, R.M.; Carey, M.S.; Wilson, O.R.; Barsoum, M.W.; Magenau, A.J.D. Well-Dispersed Nanocomposites Using Covalently Modified, Multilayer, 2D Titanium Carbide (MXene) and In-Situ “Click” Polymerization. Chem. Mater. 2021, 33, 1648–1656. [Google Scholar] [CrossRef]
  138. Bian, R.; Xiang, S.; Cai, D. Fast Treatment of MXene Films with Isocyanate to Give Enhanced Stability. ChemNanoMat 2020, 6, 64–67. [Google Scholar] [CrossRef]
  139. Guo, Q.; Zhang, X.; Zhao, F.; Song, Q.; Su, G.; Tan, Y.; Tao, Q.; Zhou, T.; Yu, Y.; Zhou, Z.; et al. Protein-Inspired Self-Healable Ti3C2 MXenes/Rubber-Based Supramolecular Elastomer for Intelligent Sensing. ACS Nano 2020, 14, 2788–2797. [Google Scholar] [CrossRef]
  140. Mayorga-Burrezo, P.; Muñoz, J.; Zaoralová, D.; Otyepka, M.; Pumera, M. Multiresponsive 2D Ti3C2Tx MXene via Implanting Molecular Properties. ACS Nano 2021, 15, 10067–10075. [Google Scholar] [CrossRef]
  141. Riazi, H.; Taghizadeh, G.; Soroush, M. MXene-Based Nanocomposite Sensors. ACS Omega 2021, 6, 11103–11112. [Google Scholar] [CrossRef] [PubMed]
  142. Liu, Y.; Zhang, X.; Zhang, W.; Ge, X.; Wang, Y.; Zou, X.; Zhou, X.; Zheng, W. MXene-based Quantum Dots Optimize Hydrogen Production via Spontaneous Evolution of Cl- to O-terminated Surface Groups. Energy Environ. Mater. 2022. [Google Scholar] [CrossRef]
  143. Liu, Y.; Zhang, W.; Zheng, W. Quantum Dots Compete at the Acme of MXene Family for the Optimal Catalysis. Nano-Micro Lett. 2022, 14, 158. [Google Scholar] [CrossRef] [PubMed]
  144. Wang, Y.; Liu, M.; Wang, Z.; Gu, Q.; Liu, B.; Zhao, C.; Zhang, J.; Xu, S.; Lu, M.; Li, H.; et al. Interlayer Environment Engineered MXene: Pre-Intercalated Zn2+ Ions as Intercalants Renders the Modulated Li Storage. J. Energy Chem. 2022, 68, 306–313. [Google Scholar] [CrossRef]
  145. Liu, M.; Qu, X.; Wang, Y.; Zhang, Y.; Zhao, C.; Liu, B.; Xue, X.; Zhang, J.; Wang, Z.; Li, J.; et al. Inspired by Philosophizing of “Point to Area”: Ion Pre-Intercalation Induces the Reconstitution of Interlayer Environment of MXene for Lithium Storage. Chem. Eng. J. 2023, 451, 139015. [Google Scholar] [CrossRef]
  146. Tan, S.; Zhao, Q.; Geng, Y.; Yin, J.; Zhou, C.; Zhang, P.; Chu, X.; Xu, S.; Lu, M.; Wang, L.; et al. Enhanced Lithium Storage Capacitance of Layered CoFe2O4&V2CTx Hybrid Anode Material Synthesized by In-Situ Hydrothermal Method. J. Alloys Compd. 2022, 918, 165778. [Google Scholar]
  147. Li, J.; Zhang, W.; Ge, X.; Lu, M.; Xue, X.; Wang, Z.; Yue, N.; Zhang, J.; Lang, X.; Jiang, Q.; et al. Etching-Courtesy NH4+ Pre-Intercalation Enables Highly-Efficient Li+ Storage of MXenes via the Renaissance of Interlayer Redox. J. Energy Chem. 2022, 72, 26–32. [Google Scholar] [CrossRef]
Figure 1. The emerging MXene family and its structure. Schematic diagram of (a) the elements related to the reported MXene [14]; (b) theoretically predicted and experimentally synthesized MXene [15]; (c) the crystal structure of MXene [16].
Figure 1. The emerging MXene family and its structure. Schematic diagram of (a) the elements related to the reported MXene [14]; (b) theoretically predicted and experimentally synthesized MXene [15]; (c) the crystal structure of MXene [16].
Symmetry 14 02232 g001
Figure 2. The surface termination investigations of MXene. (The preparation method is marked in red, the surface terminal is marked in yellow, and the application is marked in blue).
Figure 2. The surface termination investigations of MXene. (The preparation method is marked in red, the surface terminal is marked in yellow, and the application is marked in blue).
Symmetry 14 02232 g002
Figure 3. The most common process for preparing MXene at present.
Figure 3. The most common process for preparing MXene at present.
Symmetry 14 02232 g003
Figure 5. Schematic diagram of the synthesis of Ti3C2Tx MXene by fluoride etching [36].
Figure 5. Schematic diagram of the synthesis of Ti3C2Tx MXene by fluoride etching [36].
Symmetry 14 02232 g005
Figure 6. Preparation of Ti3C2Cl2 MXene by molten salt etching [27].
Figure 6. Preparation of Ti3C2Cl2 MXene by molten salt etching [27].
Symmetry 14 02232 g006
Figure 7. Reaction coordinate of MS-E-Etching Ti3C2Tx MXene [44]. (After electrolytic voltage was applied, adsorbed active sites of Ti electron holes were generated, causing Cl anions in the molten salt to attach to these active sites to form Slab-Ti*Clx).
Figure 7. Reaction coordinate of MS-E-Etching Ti3C2Tx MXene [44]. (After electrolytic voltage was applied, adsorbed active sites of Ti electron holes were generated, causing Cl anions in the molten salt to attach to these active sites to form Slab-Ti*Clx).
Symmetry 14 02232 g007
Figure 8. Synthesis of MXene from the MAX phase by MS-E-etching and the in-situ modification of surface terminations [45].
Figure 8. Synthesis of MXene from the MAX phase by MS-E-etching and the in-situ modification of surface terminations [45].
Symmetry 14 02232 g008
Figure 9. Surface reactions of MXene in molten inorganic salts [29].
Figure 9. Surface reactions of MXene in molten inorganic salts [29].
Symmetry 14 02232 g009
Figure 10. Preparation of Ti3C2 MXene with amino as the surface termination [49].
Figure 10. Preparation of Ti3C2 MXene with amino as the surface termination [49].
Symmetry 14 02232 g010
Figure 11. Schematic representation of a covalent surface modification of pristine Ti3C2Tx (Tx: -F, -OH, and/or -O) MXene via esterification reaction where the hydroxy terminations are modified with ester-bonded ω-PEG6-COOH, forming Ti3C2-PEG6 [38].
Figure 11. Schematic representation of a covalent surface modification of pristine Ti3C2Tx (Tx: -F, -OH, and/or -O) MXene via esterification reaction where the hydroxy terminations are modified with ester-bonded ω-PEG6-COOH, forming Ti3C2-PEG6 [38].
Symmetry 14 02232 g011
Figure 12. (a) Schematic illustration of simultaneous interfacial chemical grafting reaction and phase transfer for preparing a stable nonpolar Ti3C2Tx MXene dispersion [30]. (b) Photographs of Ti3C2Tx MXene dispersion in immiscible water and chloroform phases before and after the reaction [30].
Figure 12. (a) Schematic illustration of simultaneous interfacial chemical grafting reaction and phase transfer for preparing a stable nonpolar Ti3C2Tx MXene dispersion [30]. (b) Photographs of Ti3C2Tx MXene dispersion in immiscible water and chloroform phases before and after the reaction [30].
Symmetry 14 02232 g012
Figure 13. (a) Comparison of XRD patterns between Ti3AlC2 and intermediate Ti3ZnC2 [27]. (b) XRD patterns of reaction products and HCl rinsed products [2]. (c) XRD patterns of Ti3C2-O and Ti3C2-S [44]. (d) The XRD patterns of halogen termination MXene were compared [28]. (e) XRD patterns of MXene-NH2 [55].
Figure 13. (a) Comparison of XRD patterns between Ti3AlC2 and intermediate Ti3ZnC2 [27]. (b) XRD patterns of reaction products and HCl rinsed products [2]. (c) XRD patterns of Ti3C2-O and Ti3C2-S [44]. (d) The XRD patterns of halogen termination MXene were compared [28]. (e) XRD patterns of MXene-NH2 [55].
Symmetry 14 02232 g013
Figure 14. (a) SEM image of Ti3AlC2 powder [27]. (b) Ti3ZnC2 powder [27]. (c) Ti3C2Cl2. (d) The corresponding EDS [27]. (e) SEM image of -O termination MXene. (f) -S termination MXene [44]. (gm) SEM and EDS analysis of halogen termination MXene [28].
Figure 14. (a) SEM image of Ti3AlC2 powder [27]. (b) Ti3ZnC2 powder [27]. (c) Ti3C2Cl2. (d) The corresponding EDS [27]. (e) SEM image of -O termination MXene. (f) -S termination MXene [44]. (gm) SEM and EDS analysis of halogen termination MXene [28].
Symmetry 14 02232 g014
Figure 15. (a) XPS spectra of Ti 2p and Cl 2p of Ti3C2Cl2 [27]. (b) High-resolution Ti 2p spectra of -O and -S terminations Ti3C2Tx powders [44]. (c) XPS spectra of MXene-NH2 [55]. (d) XPS spectra of Ti3C2Br2 and Ti3C2Te MXene [29]. (e) The left is the high-resolution XPS spectrum of Ti3C2Br2 MXene, and the right is the high-resolution XPS spectrum of Ti3C2Te MXene [29]. (f) XPS spectra of Ti3C2(NH)MXene [29]. (g) The high-resolution XPS spectrum of Ti3C22 MXene [29]. (h) XPS spectra of F 1s and S 2p of Ti3C2Tx MXene. (i) XPS spectra of F1s and S2p of Ti3C2Tx-PSS [60].
Figure 15. (a) XPS spectra of Ti 2p and Cl 2p of Ti3C2Cl2 [27]. (b) High-resolution Ti 2p spectra of -O and -S terminations Ti3C2Tx powders [44]. (c) XPS spectra of MXene-NH2 [55]. (d) XPS spectra of Ti3C2Br2 and Ti3C2Te MXene [29]. (e) The left is the high-resolution XPS spectrum of Ti3C2Br2 MXene, and the right is the high-resolution XPS spectrum of Ti3C2Te MXene [29]. (f) XPS spectra of Ti3C2(NH)MXene [29]. (g) The high-resolution XPS spectrum of Ti3C22 MXene [29]. (h) XPS spectra of F 1s and S 2p of Ti3C2Tx MXene. (i) XPS spectra of F1s and S2p of Ti3C2Tx-PSS [60].
Symmetry 14 02232 g015
Figure 16. (a) XPS spectra of F 1s and S 2p of Ti3C2 MXene [60]. (b) XPS spectra of F 1s and S 2p of Ti3C2Tx-PSS [60].
Figure 16. (a) XPS spectra of F 1s and S 2p of Ti3C2 MXene [60]. (b) XPS spectra of F 1s and S 2p of Ti3C2Tx-PSS [60].
Symmetry 14 02232 g016
Figure 17. TEM images of (a) Ti3C2Cl2 [27], (b) MXene-NH2 [49], (c) Ti3C2-PEG6 [38], (d) P-Nb2CTx MXene [62], (e) T-Nb2CTx MXene [62].
Figure 17. TEM images of (a) Ti3C2Cl2 [27], (b) MXene-NH2 [49], (c) Ti3C2-PEG6 [38], (d) P-Nb2CTx MXene [62], (e) T-Nb2CTx MXene [62].
Symmetry 14 02232 g017
Figure 18. (a) HAADF STEM images of Ti3C2Tx sheets [64]. (b) EELS averaging over Ti3C2Tx sheets [64].
Figure 18. (a) HAADF STEM images of Ti3C2Tx sheets [64]. (b) EELS averaging over Ti3C2Tx sheets [64].
Symmetry 14 02232 g018
Figure 19. (a) Raman spectra of Ti3C2 MXene at surface terminations of Cl, S [29]. (b) Ti3C2 MXene at NH surface terminations [29]. (c) NH, O terminations, and naked MXene [29]. (d) Ti3C2-PEG6 sheets [38]. (e) Structure and composition change of the Ti3C2Br2 electrode after cycling [28]. (f) For Ti3C2Br2, the voltage of each point is A: 0.3 V, B: 1.3 V, C: 2.2 V, D: 1.2 V, E: 0.3 V [28].
Figure 19. (a) Raman spectra of Ti3C2 MXene at surface terminations of Cl, S [29]. (b) Ti3C2 MXene at NH surface terminations [29]. (c) NH, O terminations, and naked MXene [29]. (d) Ti3C2-PEG6 sheets [38]. (e) Structure and composition change of the Ti3C2Br2 electrode after cycling [28]. (f) For Ti3C2Br2, the voltage of each point is A: 0.3 V, B: 1.3 V, C: 2.2 V, D: 1.2 V, E: 0.3 V [28].
Symmetry 14 02232 g019
Figure 20. (a) FTIR spectroscopy analysis of MXene-NH2 [49]. (b) pure Ti3C2 MXene sheets and PEG6-COOH surface modification [38].
Figure 20. (a) FTIR spectroscopy analysis of MXene-NH2 [49]. (b) pure Ti3C2 MXene sheets and PEG6-COOH surface modification [38].
Symmetry 14 02232 g020
Figure 21. (a) The STEM/EDS of Ti3C2Cl2 MXene [27]. (b) Electron band structure and phonon spectrum of monolayer Ti3C2Cl2 calculated by DFT [27]. (c) The HAADF-STEM image of Ti3C2Br2 and corresponding EDS element map [28]. (d) The HR-TEM image of Ti3C2Br2 [28]. (e) The STEM image of Ti3C2Br2 [28]. (f) EDS element diagram corresponding to Figure 21e [28]. (g) Atomic and electronic structure of Ti3C2Br2, and Ti3C2I2 [28]. (h) AFM image and size distribution of MXene-NH2 nanosheets [49].
Figure 21. (a) The STEM/EDS of Ti3C2Cl2 MXene [27]. (b) Electron band structure and phonon spectrum of monolayer Ti3C2Cl2 calculated by DFT [27]. (c) The HAADF-STEM image of Ti3C2Br2 and corresponding EDS element map [28]. (d) The HR-TEM image of Ti3C2Br2 [28]. (e) The STEM image of Ti3C2Br2 [28]. (f) EDS element diagram corresponding to Figure 21e [28]. (g) Atomic and electronic structure of Ti3C2Br2, and Ti3C2I2 [28]. (h) AFM image and size distribution of MXene-NH2 nanosheets [49].
Symmetry 14 02232 g021
Figure 22. Band structures of (a) Ti2CT2 and (b) Ti2NT2 (including precursor MAX phase and MXene with different end groups) [65]. (The abscissa is the high symmetry point of the Brillouin region of the crystal).
Figure 22. Band structures of (a) Ti2CT2 and (b) Ti2NT2 (including precursor MAX phase and MXene with different end groups) [65]. (The abscissa is the high symmetry point of the Brillouin region of the crystal).
Symmetry 14 02232 g022
Figure 23. (a) UPS spectrum of MXene and MXene-NH2 valence band [49]. (b) Secondary electron cutoff regions of UPS spectra of MXene and MXene-NH2 nanosheets [49]. (c) Electrical conductivity of MXene and MXene-NH2 films [49].
Figure 23. (a) UPS spectrum of MXene and MXene-NH2 valence band [49]. (b) Secondary electron cutoff regions of UPS spectra of MXene and MXene-NH2 nanosheets [49]. (c) Electrical conductivity of MXene and MXene-NH2 films [49].
Symmetry 14 02232 g023
Figure 24. (a) Optical image of Ti3C2-Cl prepared by the molten salt method in water [49]. (b) Zeta potentials of MXene and MXene-NH2 in different solutions [49]. (c) Ultrasound images of MXene and MXene-NH2 in water (top) and 75% ethanol solution (bottom) for 1 week and 2 weeks, respectively [49].
Figure 24. (a) Optical image of Ti3C2-Cl prepared by the molten salt method in water [49]. (b) Zeta potentials of MXene and MXene-NH2 in different solutions [49]. (c) Ultrasound images of MXene and MXene-NH2 in water (top) and 75% ethanol solution (bottom) for 1 week and 2 weeks, respectively [49].
Symmetry 14 02232 g024
Figure 25. Predictions on the mechanical 2D stiffness for terminated M2XTx MXene for different surface terminations including −O2, −F2, −(OH)2, and −Cl2 [102].
Figure 25. Predictions on the mechanical 2D stiffness for terminated M2XTx MXene for different surface terminations including −O2, −F2, −(OH)2, and −Cl2 [102].
Symmetry 14 02232 g025
Table 1. Composition of the starting materials for preparing the halogenated MXene [28].
Table 1. Composition of the starting materials for preparing the halogenated MXene [28].
SamplesMolar Ratio
Ti3C2Cl2Ti3AlC2/CuCl2 = 1/4
Ti3C2Br2Ti3AlC2/CuBr2 = 1/4
Ti3C2I2Ti3AlC2/CuI = 1/6
Ti3C2(ClBr)Ti3AlC2/CuCl2/CuBr2 = 1/1/4
Ti3C2(ClI)Ti3AlC2/CuCl2/CuI = 1/1/6
Ti3C2(BrI)Ti3AlC2/CuBr2/CuI = 1/1/6
Ti3C2(ClBrI)Ti3AlC2/CuCl2//CuBr2/CuI = 1/1/1/6
Table 2. Termination semiconductor MXene and its bandgap calculated based on PBE and HSE06 [66].
Table 2. Termination semiconductor MXene and its bandgap calculated based on PBE and HSE06 [66].
MXeneTerminationPBE (eV)HSE06 (eV)
Sc2CO1.8 [67], 1.84 [68], 1.86 [69]2.90 [70], 2.92 [71], 3.01 [69]
F1.0 [69,72], 1.03 [67,68], 1.05 [73]1.64 [73], 1.84 [70], 1.88 [69]
OH0.34 [69], 0.44 [68], 0.45 [67], 0.71 [72]0.71 [69], 0.74 [70]
Cl0.88 [70]1.64 [70]
Ti2CO0.17 [72], 0.24 [74], 0.33 [75] 0.78 [76], 0.88 [74], 0.92 [77]
Zr2CO0.66 [72], 0.88 [67], 0.95 [75]1.54 [77]
Hf2CO0.8 [72], 1.00 [67,75]1.657 [78,79], 1.75 [77]
V2CF0.56 [80]
OH0.44 [80]
Cr2CO
F0.22 [72]3.15 [79], 3.49 [81]
OH0.03 [72]1.39 [79], 1.76 [81]
Cl0.15 [72]2.56 [81]
Mo2CO
F0.25 [72]
OH0.1 [72]
Cl0.15 [72]
W2CO0.194 [82]0.472 [82]
(Mo2/3Sc1/3)2CO0.04 [83] 0.58 [83]
(Mo2/3Y1/3)2CO0.45 [83]1.23 [83]
(W2/3Sc1/3)2CO0.675 [83]1.3 [83]
(W2/3Y1/3)2CO0.625 [83]1.3 [83]
Mo1.33CO2/3F1/30.5 [84]
Hf3C2O 0.155 [85]
Hf2MnC2O0.238 [86]
F1.027 [86]
Hf2VC2F0.4 [87]0.9 [87]
Mo2TiC2O0.041 [88], 0.052 [89]0.119 [88], 0.125 [89]
Mo2ZrC2O0.069 [88], 0.087 [89] 0.125 [88], 0.147 [89]
Mo2HfC2O0.153 [88], 0.213 [89] 0.238 [88], 0.301 [89]
W2TiC2O0.136 [88]0.290 [88]
W2ZrC2O1.170 [88]0.280 [88]
W2HfC2O0.285 [88]0.409 [88]
Cr2TiC2F 1.35 [79]
OH0.85 [79]
Table 3. Magnetic moment of MXene predicted by theoretical calculation [65]. (Unit: μB/unit cell).
Table 3. Magnetic moment of MXene predicted by theoretical calculation [65]. (Unit: μB/unit cell).
MXeneMagnetic Moments
(Pristine) (μB)
Magnetic Moments (Termination Group) (μB)
-O-F-OH
Ti3C21.8~1.93NonmagneticNonmagnetic
Ti2N1.0~1.1
Ti2C1.9~1.91
Ti3N20.34/Ti atom
V2C0.16
V2NNonmagnetic
Fe2C3.95
Zr2C1.90
Zr3C21.73
Mn2N7.09.08.8
Cr2C0.54/Cr atomNonmagnetic2.71/Cr atom2.24
Cr2N5.6/Cr atom3.23/Cr atom3.01/Cr atom
Sc2N1.00
Mn2N7.09.08.8
(Ti2Mn)C22.974.243.90
(Hf2Mn)C23.005.004.84
(Hf2V)C21.001.271.33
Ti4N37.000.370.88Nonmagnetic
(TiMn2)C216.34.0
(TiCr2)C23.41.83.33.0
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Tang, M.; Li, J.; Wang, Y.; Han, W.; Xu, S.; Lu, M.; Zhang, W.; Li, H. Surface Terminations of MXene: Synthesis, Characterization, and Properties. Symmetry 2022, 14, 2232. https://doi.org/10.3390/sym14112232

AMA Style

Tang M, Li J, Wang Y, Han W, Xu S, Lu M, Zhang W, Li H. Surface Terminations of MXene: Synthesis, Characterization, and Properties. Symmetry. 2022; 14(11):2232. https://doi.org/10.3390/sym14112232

Chicago/Turabian Style

Tang, Mengrao, Jiaming Li, Yu Wang, Wenjuan Han, Shichong Xu, Ming Lu, Wei Zhang, and Haibo Li. 2022. "Surface Terminations of MXene: Synthesis, Characterization, and Properties" Symmetry 14, no. 11: 2232. https://doi.org/10.3390/sym14112232

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