Skip to main content
ACS Omega logoLink to ACS Omega
. 2023 Jul 24;8(33):29859–29909. doi: 10.1021/acsomega.3c01182

Can MXene be the Effective Nanomaterial Family for the Membrane and Adsorption Technologies to Reach a Sustainable Green World?

Şirin Massoumılari , Sadiye Velioǧlu †,‡,*
PMCID: PMC10448662  PMID: 37636908

Abstract

graphic file with name ao3c01182_0010.jpg

Environmental pollution has intensified and accelerated due to a steady increase in the number of industries, and exploring methods to remove hazardous contaminants, which can be typically divided into inorganic and organic compounds, have become inevitable. Therefore, the development of efficacious technology for the separation processes is of paramount importance to ensure the environmental remediation. Membrane and adsorption technologies garnered attention, especially with the use of novel and high performing nanomaterials, which provide a target-specific solution. Specifically, widespread use of MXene nanomaterials in membrane and adsorption technologies has emerged due to their intriguing characteristics, combined with outstanding separation performance. In this review, we demonstrated the intrinsic properties of the MXene family for several separation applications, namely, gas separation, solvent dehydration, dye removal, separation of oil-in-water emulsions, heavy metal ion removal, removal of radionuclides, desalination, and other prominent separation applications. We highlighted the recent advancements used to tune separation potential of the MXene family such as the manipulation of surface chemistry, delamination or intercalation methods, and fabrication of composite or nanocomposite materials. Moreover, we focused on the aspects of stability, fouling, regenerability, and swelling, which deserve special attention when the MXene family is implemented in membrane and adsorption-based separation applications.

1. Introduction

Environmental pollution originating from population growth, increased affluence, and the industrial revolution has increased quickly.1 As a result of the ruthless consumption of the resources, the ecological balance has been deteriorated.2 The main resources used dominantly are the nonrenewable fossil fuels such as coal, natural gas, and oil, formed over thousands of years.3 The amount of anthropogenic CO2 released to the atmosphere by the use of these resources between the years of 1750 and 2011 is approximately 2040 Gt, and 50% of this amount has been released only in the last 40 years.4 Unfortunately, the release of greenhouse gases causes the local temperature to increase. For instance, a 3.5 °C increase in an average world temperature corresponds to a 7 °C increase in the polar regions. The temperature rise leads to the ice melting at the poles and mountains, and hence, the sea level increases, which is expected to increase between 9 and 88 cm in 2100.4 Regrettably, besides the air pollution, soil contamination has reached alarming levels due to the release of heavy metals, solvents, pharmaceuticals, pesticides, fertilizers, radioactive wastes, and plastics.5 Similarly, the pollution risk also exists for water sources such as rivers, lakes, oceans, etc. It was expected that by 2025, half of the world’s population would face water scarcity due to the contamination of drinking-water sources with feces and the merger of clean water with sea water as a result of the rising sea level.6 Hereby, people all over the world will be forced to become a climate migrant. Moreover, biodiversity and the marine ecosystem are also under serious threat due to the rapid climate change and the acidification of the ocean.7 The direct impact of pollution on human health is recognized as the occurrence of 9 million premature deaths in 2019, of which 6.7 million were from air pollution and 1.4 million from water pollution. Premature deaths were distressingly recorded in low and middle-income countries dominantly, which was explained by the “climate justice” concept.8 Collectively, air, soil, and water pollution are interconnected like links in a chain, and human health is influenced from them, either directly or indirectly. To offer a remedy for the environmental pollution, a number of international meetings and conferences have been held since the 1970s.4 In addition, the foundation of the Intergovernmental Panel on Climate Change (IPCC) was established in 1988 by the decision of the United Nations General Assembly.4 The most important international agreement is the Kyoto protocol in which strict obligations for the reduction and limitation of greenhouse gases were introduced. The experiences gained from the provided agreements paved the way for the content of the Paris Agreement in which several strict targets were set.4 Its long-term goal is limiting the global warming to about 1.5 °C, which was assured by the every five-year meetings to evaluate what has been done so far.

More than 30,000 different chemicals are used in our daily lives. The number of pollutants originated from the combination of these chemicals is definitely countless nowadays.9 The major pollutants produced via reaction of these chemicals and dominantly encountered in the water are heavy metals, dyes, oils, plastics, herbicides, and pesticides.10 In order to provide water safety, physical, chemical, and biological methods are being used for water remediation. Physical methods such as sedimentation, degasification, and membrane are the simple ones. They are energy efficient methods, whereas their capacities are limited due to the clogging problem. Although they reveal a solid performance, generally chemical methods are preferred in industry. Highly used chemical methods are flocculation and coagulation, ozonation, chemical precipitation, and ion exchange. Although these methods are also simple and have low operational costs, their main disadvantage is the high sludge production and disposal.10 Besides water treatment processes, the development of gas separation processes are also essential in order to suppress environmental pollution. To keep environmental sustainability and restrict carbon emissions, commonly used gas separation technologies are absorption/amine scrubbing, cryogenic distillation, adsorption, and membrane-based separation.11 Though amine-based scrubbing and cryogenic distillation are conventional and effective methods, a high energy requirement is the major challenge faced during CO2 separation.11,12 As an alternative, adsorption- and membrane-based gas separation applications were provided. They have become pervasive in applications related to energy, food, water, biotechnology, etc. Adsorption based separation is an attractive method where it promotes the use of various adsorbent materials, applicability in multiple operation modes, and high adsorption capacity.11 Likewise, the reasons for the attention of membrane-based separation processes are that they have low energy consumption, facile operation, and are scalable and environmentally friendly. Considering the application of membrane technology in both water treatment and gas separation, it has been proven in the literature that its efficiency in separation was close to the order of other methods, and its energy consumption was very low compared to the others, making it stand out from others.1315 Accordingly, membrane market size was supposed as approximately 7 billion dollars in 2021 globally.16 As environmental pollution continues to increase, the membrane market will become a huge and competitive platform and membrane technology will record a big leap.17 Similarly, adsorbent market size is also flourishing with the estimation of 3.9 billion dollars in 2020.16 The adsorption market is expected to thrive more in the future with the development of novel adsorbents having high recyclability and reusability, improvement in the process design, and applied advanced characterization methods.12

To date, various membrane and adsorbent materials were examined for the target separation applications. Mostly, polymeric membranes have been widely used due to their scalability, processability, and cost effectiveness. However, the newly developed nanoporous inorganic materials offer advantages in their use as membrane and adsorbent materials due to their uniform pores and narrow pore size distribution. Very recently, after the discovery of graphene in 2004, two-dimensional (2D) nanomaterials started to bring much attention to membrane and adsorption technologies. Especially in the membrane area, it deserves this intense interest due to leading an ultrathin membrane fabrication, which lowers the mass transfer resistance and improves the production. Moreover, in the adsorption process, ultranarrow and functionalized interlayer channels distinguish the solute particles according to their size and chemical properties, resulting in an increase in the particle removal rate. There are different 2D material families such as graphene oxide (GO), zeolite, MXene, transition metal dichalcogenide (TMD), metal organic framework (MOF), covalent organic framework (COF), etc. which offer great potential in both membrane and adsorption technologies. 2D materials used in adsorption or membrane-based separation applications are prepared in the form of either nanoporous or nanolaminates. To enhance the separation properties of 2D membranes or adsorbents, several strategies were developed such as adjusting interlayer channels, modifying surface properties via functionalization, intercalating various molecules (cations, solvents, organic molecules), tailoring the porosity, etc.18 These strategies are beneficial to handle with the trade-off between permeability and selectivity existing in membrane processes and between working capacity and adsorption selectivity existing in adsorption processes.12

Discovery of transition metal carbides and nitrides named MXene in 2011 had a deep impact on the world of materials science. MXene nanomaterials are synthesized from MAX phases formulated Mn+1AXn (n = 1, 2, 3, 4) that comprise M, A, and X, representing early transition metals (Sc, Y, Ti, Zr, Hf, V, Nb, Ta, Cr, Mo, W, and Mn), IIIA or IVA group elements, and C and/or N, respectively.19 MAX phase which is the laminar hexagonal structure has the form of stacked transition metal carbide or nitride sheets in which X atoms are glued within A layers.20 Initially, the MXene (e.g., Ti3C2Tx) fabrication method was the immersion of the MAX phase (e.g., Ti3AlC2) into the hydrofluoric acid (HF) at room temperature, which is named as the etching process. Thereby, the bonds between transition metals and A elements deteriorated, and A elements were removed.21 Within the early years of its discovery, MXene members of Ti2CTx, (Ti,Nb)2CTx, (V,Cr)3C2Tx, Ti3CNTx, and Ta4C3Tx are included in the MXene family via the fabrication of the HF etching method.22,23 In 2013, the isolation of MXene as a single sheet by intercalating the organic molecules, which is known as the delamination of layers, was one of the most important turning points in the history of MXene nanomaterials. Instead of using a dangerous HF solution in etching, the use of acid mixtures with fluoride salts emerged in 2014 and single sheets of Ti3C2Tx with a small lateral size was obtained.23 After realizing that MXene nanomaterials have synthesis method dependent properties, various etching approaches were examined such as molten salt,24 Lewis acids,25 and electrochemical etching.26 With the help of evolutionary steps in MXene nanomaterials, their different forms such as MXene with in-plane and out-of-plane ordering of the metal atoms, in-plane vacancies of metal atoms, solid solution of metal or carbon/nitrogen atoms, and various surface terminations were synthesized and members of the MXene family increased rapidly.23,27 In 2011, there were only 70 MAX phases, whereas this number has reached 150, leading to the observation of a theoretically unlimited number of MXene structures.20

Attempts to adjust etching-delamination protocols or use alternative synthesis methods provided researchers with a way to obtain different MXene members in the above-mentioned forms, which enabled scientists to diversify the physical properties of the MXene family. For instance, the mechanical strength and stiffness of Ti3C2Tx nanosheets were examined by in situ transmission electron microscopy (TEM) probing and atomic force microscopy (AFM) nanomechanical mapping28 and displayed its superiority both experimentally and theoretically.29,30 The highest adhesion energy was identified between the nanosheets of Ti3C2Tx regardless of the layer number compared to the other 2D nanomaterials such as graphene, MoSe2, and SiO2.31 It was realized that depending on the each element within its structure, its electronic conductivity could be altered. MXenes having transition metals of chromium, molybdenum, and tungsten were defined as topological insulators,21,3237 while Ti3C2Tx was accepted as the greatest conductor.38 However, Ti3C2Tx terminated with the groups of −F2 and −(OH)2 displayed the attitude of semiconductors with narrow band gaps, contrary to the one terminated with −O2 groups.39 Promisingly, Ti3C2Tx revealed high temperature stability up to 500 °C under argon atmosphere, with a formation of TiO2 crystals onto the edges.40 However, Zr3C2Tx kept its structure up to 1000 °C under vacuum.41 Considering the superb properties of already existing MXene members such as high mechanical strength, adhesion, electrical conductivity, and thermal stability, depending on their application of interest, further improvements on novel MXene members can be obtained by only manipulating their precursor, composition, and synthesis methods.

The tuneability of MXene properties has been studied from different perspectives. It was proved both experimentally and theoretically that optical and electronic properties of MXene nanomaterials having a M2CTx (M = Ti, Nb, and V) structure can be altered depending on the M-site as well as its composition.42 For instance, Halim et al.43 reported that Ti2CTx and Nb2CTx displayed different optical properties due to their inevitable discrepancy in electronic configuration and bonding between the carbon and transition metal. It was also suggested for a large number of structures and corresponding compositions as M2XTx, M3X2Tx, and M4X3Tx that both their electrical conductivity and electromagnetic interference shielding (EMI) performance could be adjusted in a broad range.44 Additionally, it was displayed on 62 different MXene structures formulated in the form of (M′2/3M″1/3)2X that magnetic properties can be easily varied because of their geometries.45 Beyond mechanical, optical, electrical, and magnetic properties, MXene nanomaterials have exceptional antibacterial activities. Antibacterial performance of Nb2CTx, Nb4C3Tx, Ti3C2Tx, and Ti2CTx were compared by the group of Mahmoud.46 Nb2CTx displayed lower antibacterial efficiency than Nb4C3Tx against the Gram-positive (S. aureus) (96%) and negative bacteria (E. coli) (94%), which were slightly lower than Ti3C2Tx nanosheets (98%).46 Tunable structure–property relation of MXene nanomaterials provides many advantages for various applications of interest.

Besides the outstanding properties of MXene nanomaterials, there are some challenges associated with their structures, which restrict their use in separation applications. The main limitation is their restricted stabilities under different conditions. It was revealed that Ti3C2Tx heated at different temperatures in air lost its structure stability very rapidly and TiO2 nanocrystals were observed in thin sheets of disordered graphitic carbon.47 The corresponding oxidation rate of MXene under different conditions was often explored in the literature.48 The strongest MXene oxidation rate was observed in a H2O2 environment, while the weakest rate was recorded in dry air.48 To tackle with the fast oxidation rate of MXene, improvement in either the synthesis procedure of MXene in order to eliminate the defects or its storage conditions is required. Since the defective sites on or at the edge of the MXene nanosheets become susceptible to the oxidative degradation, novel synthesis approaches were aimed to be developed.49 Mathis et al.50 just modified the MAX phase of Ti3C2Tx and increased its structural stability in closed vials up to the 10 months at room temperature. Barsoum et al.51 deactivated the possible interaction between dissolved oxygen in water and the edge of MXenes (Ti3C2Tx and V2CTx) via polyanionic salts and prevented oxidation. Alternatively, storage conditions such as temperature, pH, and concentration of colloidal dispersion of MXene were examined. Athavale et al.52 reported a comprehensive survey of their group’s progress in mitigating the degradation of MXenes via low-temperature storage and the use of antioxidants. Zhang et al.53 proposed storing conditions for the colloidal dispersion of Ti3C2Tx as Ar-filled bottles at 5 °C where chemical stability was improved from days to months. Zhao et al.54 replaced water with a suitable solvent as a storage medium in order to protect nanosheets from oxidation. 1-(3-aminopropyl)-3-methylimidazolium bromide [type of ionic liquid (IL)] exfoliated Ti3C2Tx and revealed a stability of up to 80 days even after exposure to water.54 Not only the stability but also the restacking problem of nanosheets were prevented by the use of ILs as solvents for the storage and dispersion of MXene nanosheets. Other limitations of its use in separation applications is the restacking of MXene nanosheets. To keep the stacking distance (nanoconfined space) between nanosheets constant, MXene structures were delaminated via various intercalants such as solvents, ions, organic molecules, other layered materials, and polymers.55 By altering the size of intercalant molecules, it was targeted to tailor the interlayer distance of MXene. Alternatively, it was utilized from the interaction capability of intercalants with the MXene surface in order to create a synergy between them.55 Munir et al.56 identified the effect of etchant concentration, solvent type, and sonication time on the adjustment of interlayer distance of MXene nanomaterials. They demonstrated that 30% HF etchant, dimethyl sulfoxide (DMSO) as a solvent, and 135 min of sonication time were the best parameters to synthesize MXene with the highest interlayer spacing.56 Since the implementation of the MXene nanomaterial family has increased for a variety of applications, its toxicity becomes another constraint. Therefore, its biotoxicity was aimed to be identified for several normal and cancer cell lines. It was reported that cell viability of the normal cell line (HaCaT) was over 80% in the presence of Ti3C2Tx, even with the highest concentration of 500 mg/L.57 Nb2C modified with polyvinylpyrrolidone (PVP) was inserted into a mouse model, and no adverse effects were recorded, supported by the biochemistry and hematological tests. Similarly, Ti3C2Tx inserted into the zebrafish embryo model with a concentration of up to 100 mg/L did not lead to any adverse effects.58 The influence of not only the MXene concentration but also its exposure time, lateral size, and functional groups on cytotoxicity should be investigated. Additionally, even though toxicity studies are increasing day by day, more studies are required for different types of MXene in vivo.

To adopt the MXene family in membrane or adsorption technologies, other criteria that need to be figured out are the adjustment of surface termination groups, sheet size, surface area, and defect ratio. Therefore, novel etching or delamination routes were proposed. Since 2011, dozens of MXene nanomaterials have been prepared via the main MXene synthesis approach of the HF-etching route. Even it is the effective way to fabricate high quality of MXene nanosheets, due to the safety and environmental impact, different etching, and delamination processes to fabricate MXene from the MAX phase performed to date. Alternative to the HF-etching method, it was reported that the LiF + HCl route also provided high quality MXene nanosheets.59 Differently, Ti4N3Tx by fluoride molten salt,24 fluorine free MXene by hydrothermal alkali etching,60 Ti3C2Tx from non-Al MAX of Ti3SiC2 by HF-H2O2 as etchant,61 and hardly obtained MXenes by Lewis acidic molten salt25 were fabricated in a high quality form. On the other hand, covalent modification of surface terminations of MXene,62 delamination of titanium- and niobium-based MXenes with urea and amine,63,64 delamination of Ti3CNTx with tetrabutylammonium hydroxide (TBAOH),65 use of lithium salts,66 in-plane ordered vacancy (Mo1.33CTx) MXene synthesis with the improved etching and delamination methods,67 and etching-delamination strategies of UV-induced68 and chemical-combined etching69 were applied successfully without leading any structural deformations. And it seems that new methods will continue to be developed to meet the criteria for the target of applications.

Especially, for membrane applications, MXene nanosheets are mostly casted in a film form by processing the MXene solution.70 One of the ways for processing the MXene solutions is the printing methods such as inkjet, screen, and extrusion, which provide a large scale and low-cost film fabrication, and controlled MXene deposition.70 However, there are some challenges due to the use of low viscosity of water-based solvents in printing methods as the clogging of nozzles of the printer. Recently, with the help of synthetic structural proteins as binder molecules and DMSO as solvent, novel inks were develop and printed onto the various substrates containing polyethylene terephthalate (PET), poly(methyl methacrylate) (PMMA), glass, and cellulose paper.71 Surprisingly, the MXene solution containing multilayered Ti3C2Tx and access to the MAX phase after the delamination step was used as ink for screen printing.72 Although the thickness of the MXene film was lower than that obtained via inkjet printing, using MXene sediment-based ink named “trashed” emerged as a green and sustainable fabrication method for the fabrication of MXene films.72 Hybrid inks consisting of specific cellulose nanofibrils and Ti3C2Tx were applied via 3D extrusion printing on a flexible smart textile.73 Comparing these printing methods, fabrication yield of screen printing is mainly higher than inkjet and extrusion printing methods.70 The other way for processing the MXene solutions to form a film is the coating techniques such as dip, spin, blade, and spray coatings. Dip and spin coatings enabled a film obtained with the lowest thickness and good conductivity for the MXene dispersion having the viscosity range between 1 and 100 mPa × s.74 Importantly, good alignment of nanosheets was provided via spin coating as a result of the applied centrifugal force.75 However, its production yield was lower than the other coating techniques. As the viscosity of the MXene dispersion surpassed 1000 mPa × s, blade coating became suitable to create the MXene film with a large film area.76 However, thick film thickness was observed compared to the films fabricated via the dip and spin coating methods. Fabrication throughputs of coating techniques were reported as follows, spray > blade > dip-spin coatings.70 The highly preferred way in laboratory conditions for processing the MXene solutions to form a film is the filtration methods such as vacuum- and pressure-assisted filtration methods. In vacuum-assisted filtration methods, the MXene dispersion is filtered through a porous filter substrate under vacuum to form a free-standing MXene layer.77 Film thickness can be adjusted by varying the concentration of MXene in dispersion and filtration volume. Moreover, film formation can be influenced from the vacuum speed, the choice of solvent and substrate membrane, viscosity of dispersion, etc.70 The highest specific surface area (SSA) was reported as 575 cm2 for Ti3C2Tx, prepared by an electrophoretic deposition method,78 not vacuum filtration. Although a vacuum-assisted filtration method is easy to control, it has a large scalability problem in order to be applied in membrane industry.

The MXene family with their tremendous properties such as high surface area, hydrophilic manner, high metallic conductivity, active surface sides with tailorable functional groups, tunable interlayer spacing, and surface chemistry has been garnering increasing attention from several industries. MXene nanomaterials were practiced in widespread use in different applications such as energy storage,79,80 supercapacitors,81 hydrogen storage,82 catalysis,83 ultrafast photonics,84 membrane-based separation technologies,8594 biomedical engineering, biosensors,95,96 and memristive and tactile sensory systems.97 Unlocking the full potential of MXene has revealed the need for mass production. Promisingly, Gogotsi and colleagues98 synthesized Ti3C2Tx, for small (1 g) and large (50 g) batch sizes in a reactor (1 L) they designed. To synthesize a large batch of MXene, the protocol used to synthesize the small batch size was scaled up easily without any adjustment in conditions. Interestingly, very recently, Chen et al.99 used a supercritical etching method assisted by supercritical carbon dioxide for the mass production of Ti3C2Tx, Nb2CTx, Ti2CTx, Mo2CTx, and Ti3CNTx with the yield of ∼1 kg within ∼2–5 h. These are inspiring studies that pave the way for the scalability of MXene synthesis in order to observe industrial quantities.98,99 However, more studies are required to reach above the kilogram scale.

Since the MXene nanomaterial family provides outstanding physicochemical and structural properties, it attracted considerable attention in different applications. Therefore, several critical reviews have been published dominantly focusing on its separation applications.86,90,91,94,100124 However, most of them could not cover such a large number of separation applications and aimed to concentrate on specifically either the membrane90,100102 or adsorption94,103108 based separation processes. However, there are comprehensive review studies covering the potential of MXene membranes and adsorbents for a single separation application such as pervaporation,109 gas separation,110112 desalination,86,113,114 dye removal,91,115 oil-in-water separation,116 removal of heavy metal ions, and radionuclides.115,117120 On the other hand, few studies108,121,122 aimed to highlight the effect of several strategies such as various surface modification approaches of MXene, intercalation methods to enhance interlayer distance of MXene nanochannels, etc. on the improvement of separation properties of MXene-based membranes and adsorbents for several practical applications. Additionally, there are also few studies that only discussed the specific modification approaches applied for the MXene family (intercalation chemistry123 and surface functionalization124) to improve its potential in different fields.

Addressing the lack of adequate and safe water and impact of climate change is among the most important challenges of our time. Herein, we focused on the evaluation of performance of the most cost-effective, high performing, easy implementable membrane- and adsorption-based separation processes to suppress these challenges. As a being potential candidate nanomaterial family, successful implementation of MXene was revealed in membrane- and adsorption-based applications, namely, gas separation, solvent dehydration, dye removal, separation of oil-in-water emulsions, heavy metal ion and radionuclides removal, desalination, and some other prominent applications. In each section, the inspiring article in which the MXene nanomaterial was applied to that application was given initially, and then it was followed by compiling the considerable progress that has been made in each application. Information of inspiring articles for each application is depicted in Figure 1(a) along with a timeline manner. Additionally, it was also aimed, on the one hand, to prove the applicability and the effectiveness of MXene nanomaterials to several separation applications by tabulating their performances and, on the other hand, to unroll the use of a limited number of MXene members, instead of increasing the number of studies evaluating the MXene nanomaterials for different areas. Up to now, intensive efforts have been devoted to the use of the Ti3C2Tx member from the MXene family for the provided diversified separation applications, as summarized in Figure 1(b), although there are an abundant number of MXene members identified. However, as illustrated in Figure 1(c), a tremendous increase was observed in the applications of MXene nanomaterials for various fields since its discovery in 2011. Additionally, in this review, we provide an up-to-date overview of the major achievements on bio/fouling, swelling, regenerability, and long-term stability, which offer great promise to enable the proliferation of the MXene nanomaterial in separation applications. Collectively, we have highlighted not only the separation performance of the MXene family but also its possible applications to industry by discussing them in terms of different perspectives. Considering that only a few of the MXene have been revealed to date, researchers have a long way to go in the near future to thoroughly test them all for a variety of separation applications.

Figure 1.

Figure 1

Literature survey of MXene nanomaterials. (a) Milestones of each separation application,125143 (b) distribution of MXene types that were investigated for the separation application (other MXene types were listed on the left side of the figure), and (c) growth of the MXene-based studies in all fields on the Web of Science database reported on December of 2022 (reviews and patents were excluded). Inset figure of panel (c) represents the growth of the MXene literature in only separation applications, along with its percentage over all fields in each year.

2. Gas Separation

With regard to the high demand for technologies in the context of sustainability and resource recovery, membrane and adsorption technologies attracted eminent attention as the major paradigm for gas separation in the past few decades. For gas separation application, the first MXene-based membrane was fabricated very recently by Ding et al. in 2018.129 Ding et al.129 synthesized MXene (Ti3C2Tx) nanosheets with the uniform subnanometer channels and used these channels as blocks to create 2D laminated membranes with different thicknesses for the investigation of H2 separation. They reported that while single H2 permeability of the MXene membrane was on the order of 2400 Barrer, its binary H2 permeabilities for H2/CO2, H2/N2, H2/CH4, H2/C3H6, and H2/C3H8 mixtures were 2227, 1976, 1931, 2312, and 2357 Barrer, respectively.129 In addition to its superior single and binary H2 permeation coefficients, they also revealed that it had an ideal selectivity of 238 and a separation factor of 167 for H2/CO2.129 To further estimate the reason behind this inspiring H2 separation performance of MXene membrane, they used molecular simulation techniques and ascribed to the channel orderliness. As a result of this study, among all 2D and three-dimensional (3D) inorganic materials synthesized up to now, MXene with the thickness of 2 μm has emerged as the best performing material with H2 permeability of 2267 Barrer and H2/CO2 selectivity of 167 for the binary gas mixture, as illustrated in Figure 2(a).129 Its heart’s beating location in Robeson’s trade-off diagram for H2/CO2 separation accelerated the research about gas separation performance of MXene within five years. After the first encouraging and exciting results came from Ding et al.129 about H2/CO2 separation of MXene membranes, other studies aroused targeting different gas pairs like the study of Fan et al.144 focusing on H2/N2 separation. Fan et al.144 similarly fabricated 2D lamellar MXene (Ti3C2Tx) membrane for the investigation of H2/N2 separation. While single H2 and N2 permeabilities were reported as ∼4.05 × 10–7 (968 Barrer) and ∼1.84 × 10–8 (44 Barrer) mol/m2 × s × Pa, those for binary mixtures were measured as ∼3.99 × 10–7 (954 Barrer) and ∼0.37 × 10–8 (8.76 Barrer) mol/m2 × s × Pa, respectively. Ideal and binary H2/N2 selectivities of the MXene membrane were measured as ∼22 and 41, respectively.144 Compared with the single H2 permeability and ideal selectivity (129) of MXene membrane fabricated by Ding et al.129 for H2/N2 separation, the one provided by Fan et al.144 revealed an order of magnitude greater permeability, but an order of magnitude lower H2/N2 selectivity, as given in Figure 2(a). Ding et al.129 fabricated 2D lamellar MXene by etching with hydrogen chloride (HCl) + lithium fluoride (LiF) solutions, which provide high exfoliation yields. However, Fan et al.144 synthesized 2D lamellar MXene via HF etching method and used DMSO as an intercalating substance, leading to the enhanced interlayer distance between MXene laminates. To reveal the gas separation performance of MXene membranes, we compiled the data of free-standing MXene membranes in the early published studies and plotted them on the Robeson diagram as illustrated in Figure 2(a). Concerning the H2/CO2 and H2/N2 separation performances of MXene membranes,129,144147 they revealed H2/CO2 separation well above the upper bound and H2/N2 separation in the range of upper bound line exceeding the other 2D membranes. Mainly, this performance was attributed to its subnanometer interlayer spacing between the well-aligned and regular MXene nanosheets, which behave as a molecular sieving membrane.

Figure 2.

Figure 2

Gas separation performance of MXenes. (a) Comparison of H2/CO2 and H2/N2 separation performances of MXene membranes proposed by Ding et al.,129 Fan et al.,144 Wang et al.,145 Qu et al.,146 and Fan et al.147 with the other nanomaterial membranes. The red and blue lines denote the 2008 upper bound of the polymeric membrane for H2/CO2 and H2/N2, respectively, assuming membrane thickness is 0.1 μm. (b) H2 permeability and H2/CO2 selectivity of MXene membranes as a function of membrane thickness, which were measured by Ding et al.,129 Qu et al.,146 and Shen et al.153 (c) Analysis of the interlayer spaces of pristine and functionalized MXene nanofilms using XRD spectra. MB and MBP refer to the MXene-borate and MXene-borate-PEI composites, respectively. Adapted with permission from ref (153). Copyright 2018 Wiley Online Library. (d) Comparison of gas adsorption on pristine and functionalized MXenes at 25 °C. Adapted with permission from ref (153). Copyright 2018 Wiley Online Library.

In addition to the membrane-based separation process, MXene nanomaterials were also used as adsorbents for the capture of different targeting gases such as CH4,148,149 CO2,150,151 formaldehyde (HCHO),152 etc. Zhang et al.152 demonstrated the stability of Ti3C2O2 up to 177 °C and its high HCHO adsorption capacity (6.9 mmol/g) in their computational study where density functional theory (DFT)-based ab initio molecular dynamics simulation was carried out. They proposed that interfacial van der Waals (vdW), H–O, and C–O interactions were responsible for gas capture in MXene adsorbents. These promising studies have enlightened the importance of exploring the gas separation performance of MXene nanomaterials for the membrane- and adsorption-based technologies. Since they were very recently discovered for gas separation applications in membrane and adsorption technologies, the mechanism underlying their outstanding separation performance needs to be investigated. Therefore, several issues were considered in the literature such as rationally tailoring its 2D channels by either tuning thickness or intercalating specific species between its nanochannels to optimize the separation properties. The studies identifying the MXene nanomaterials as membranes and adsorbents tried to answer the question of “Can better gas separation performance be achieved by the help of MXene nanomaterials?”

2.1. Membrane-Based Separation

Gas transport mechanism within the membrane can be affected by various parameters especially physicochemical properties of the membrane material. Membrane structure and thickness, nonselective defects/voids, and galleries between nanosheets can change the transport rate of gas molecules. Interlayer spacing was suggested as the most important factor affecting the gas transport in lamellar membranes due to the tunable channel width. To investigate the degree of effect of interlayer spacing on gas separation performance of MXene membranes, several studies were carried out in the literature.129,144,153155 Ding et al.129 varied the thickness of the MXene membrane with 0.2, 0.5, 1.1, 2, 3.2, and 5.1 μm to show the relationship between membrane thickness and H2 permeability and H2/CO2 selectivity. H2 permeability was decreased with the membrane thickness from 2961 Barrer of 0.2 μm-thick membrane to 1302.7 Barrer of a 5.1 μm-thick one, while H2/CO2 selectivity was increased from ∼4.5 to 200 as given in Figure 2(b). Likewise, Shen et al.153 revealed the effect of membrane thickness on H2 permeance and H2/CO2 selectivity using ultrathin MXene (Ti3C2Tx) lamellar membranes having the thickness of 0.005, 0.01, 0.02, 0.03, 0.05, and 0.08 μm (change in interlayer spacing from 15 to 11.6 Å). Similar gas separation performance alteration of the thick MXene membranes fabricated by Ding et al.129 was revealed by the ultrathin membranes synthesized by Shen et al.153 where H2 permeances and H2/CO2 selectivities of MXene membranes were changed from ∼36,000 to 350 GPU and from ∼5 to 31, respectively, as given in Figure 2(b). The most severe drop in H2 permeances (from 6000 to 3 GPU) and similar increase in H2/CO2 selectivities (from 2.32 to 30.3) were reported by Qu et al.146 accompanying with the increase in thickness from 0.05 to 0.22 μm (decrease in interlayer spacing from 4.0 to 3.1 Å) for the self-cross-linked MXene (Ti3C2Tx) membranes by heat treatment at 140 °C for 10 h. The reverse relation of membrane thickness with interlayer spacing was suggested, and therefore, the importance of interlayer distance on gas separation was underlined by these studies. To clarify the proposed claim and explain the gas separation mechanism, Ding et al.129 carried out molecular dynamics (MD) simulations for MXene nanosheets having two different interlayer spacing and found that small disturbances of nanochannels regularity affected the interlayer spacing leading to the deteriorated gas selectivity. While H2/CO2 selectivity of the MXene membrane with the interlayer distance of 4.5 Å was computed as only ∼70, that of MXene membrane with interlayer distance of 3.5 Å was >200.129 In accordance with these studies, the decrease in gas permeability and increase in selectivity with the membrane thickness was related to the prolonged diffusion pathway due to the shrinkage of interlayer spacing. Interlayer spacing was influenced not only by the membrane thickness but also by the membrane film treatment conditions, such as temperature, which had a dominant influence on the change of interlayer spacing. Fan et al.144 identified the effect of high temperature treatment on H2/N2 separation of MXene membrane. For mixed gas permeation measurements, H2 permeance and H2/N2 separation factor of a MXene membrane treated at 320 °C were reported as 612 GPU and 41, respectively, whereas those treated at 20 °C were 690 GPU and 12.144 Since its interlayer spacing decreased from 13.4 to 12.8 Å with the temperature increase from 20 to 300 °C, an enhanced molecular sieving was observed yielding the improvement in the H2/N2 separation factor.144 The most tremendous decrease in the interlayer spacing was observed from 3.45 to 0.24 Å by the study of Emerenciano et al.155 with treatment of the Ti3C2Tx membrane at 500 °C under an Ar/H2 atmosphere, compared to the one treated at 80 °C under vacuum.

On the other hand, since the best way to identify the underlying mechanism behind the interlayer spacing is to carry out molecular simulations, Li et al.154 examined in depth the diffusion of several gas molecules (He, H2, CO2, N2, and CH4) for anhydrous and hydrous MXene (Ti3C2O2) nanosheets by varying the interlayer spacing. Their results supported the above-mentioned experimental studies revealing the significant effect of interlayer spacing between MXene nanosheets on gas transport. They proposed that for small interlayer spacing, gas molecules diffused close to the MXene walls, whereas for large interlayer spacing, diffusion was affected from the mass of the gas molecule instead of its kinetic diameter.154 Li et al.154 suggested that hydrous MXene ranging between d-spacing of 6 and 8 Å and anhydrous MXene having d-spacing around 5 Å displayed the optimum gas separation performances based on the criteria of H2/CO2 diffusion selectivity >70 and H2 diffusion >7 × 10–5 cm2/s. Jin et al.156 investigated the thickness of MXene on H2/N2 separation by using two different nanochannel models, flatwise nanochannels (>20 Å) and corrugated nanochannels (<20 Å), which were formed with nonregularly distributed MXene nanolayers. Similarly, as the thickness of MXene reduced from 1 to 0.02 μm, H2 and N2 permeances increased from 477.6 to 4090 GPU and from 24.2 to 201.5 GPU, respectively. However, interestingly, H2/N2 selectivity kept stable as 13.5. It was ascribed to the stability of the fractional contribution of Knudsen diffusion and the molecular sieving mechanism as 18% and 82%, respectively. Therefore, according to their model, they suggested that the thickness of the MXene membrane did not have any effect on the fractional diffusion of gases. While studies on the effect of interlayer spacing on gas separation of MXene membranes continue, on the other hand, new approaches have been sought to enhance selectivity without sacrificing from permeability. Considering the adsorption process, in our recent atomistic scale simulation study,157,158 we have targeted to quantitatively identify the alteration of the CO2 working capacity and CO2/H2 adsorption selectivity of 730 MXene adsorbents after the enhancement in the interlayer distance. Therefore, adsorption simulation of the proposed MXene database was carried out for structures having two different interlayer distances (3.5 and 6.5 Å). We recorded a considerable decrease in both CO2/H2 adsorption selectivity and CO2 working capacity of 76% of the total MXene structures in numbers.157,158 As a result of electrostatic and vdW interactions, CO2 can be easily kept within the MXene nanochannels at the short interlayer distance, and the effective gas adsorption can be observed for MXene adsorbents.

One of the approaches that was intensely focused is the intercalation of MXene nanochannels via several types of molecules. This approach has an importance of either enhancing the gas permeability due to the increase in interlayer spacing or altering the perm-selectivity of a membrane due to the preferences of intercalated species. Shen et al.153 modified MXene by intercalating different molecules such as borate and amine to synthesize CO2-selective MXene membranes. The interlayer spacing of MXene was decreased from 15 to 12.9 Å with the temperature increase from 55 to 75 °C, which is related to the enhancement in cross-linking of a MXene surface with the functional groups, and hence CO2/CH4 selectivity of MXene was manipulated. The reason for this was speculated that with the addition of borate molecules, CO2 uptake was increased as 13%, and with the addition of amine, this improvement was raised to 43% (CO2 permeance, 350 GPU, and CO2/CH4 selectivity, 15.3) due to the zipped in-plane slit-like pores [see Figure 2(c–d)]. However, adsorption of H2 and CH4 showed no change. Additionally, Shen et al.153 suggested that pristine MXene membranes were diffusion-selective membranes, which had low H2 solubility and high H2 diffusivity. However, intercalated MXene membranes with borate and amine were proposed as solution-controlled membranes due to the greater adsorption selectivity (28.3) than diffusion selectivity (0.05). In addition to the intercalation of bulky molecules, ions were also preferred to stem from the strong interaction with the negatively charged MXene nanosheets. Fan et al.147 reported superior improvement in binary H2/CO2 selectivity of MXene intercalated with Ni2+ ions (615) compared to the pristine one (215), yielding in almost 3-fold increase. MXene intercalated with Pd2+ ions also displayed satisfactory binary H2/CO2 selectivity as 242,145 although not so high as that of MXene intercalated with Ni2+ ions.147 It was proposed that cationic intercalation is a vital approach to tune interlayer spacing by weakening repulsive electrostatic interactions and promoting a strong interaction between MXene nanosheets, which leads to regular channel size. Lin et al.159 aimed to adapt the similar mechanism via intercalating MXene nanosheets with deep eutectic solvent (DES). Rather than observing H2-selective membranes as in cation-intercalated membranes, they ended up with a CO2-selective membrane with a CO2/H2 selectivity of 12.4 by the DES intercalation. Collectively, although a limited number and type of intercalants was tested to tune the interlayer distance of MXenes and alter their gas separation properties, the identified ones evidently revealed their efficiency in selectivity due to their size-sieving effect.

To combine the superior permeability and selectivity performances of MXenes with the ease of manufacture of polymers, mixed matrix membranes (MMMs) were addressed where MXenes were used as fillers in a polymeric continuous phase. However, the main issue that was considered is the compatibility between MXene and polymer. Liu et al.160 fabricated MMMs consisting of Ti3C2Tx and commercial polyether-polyamide block copolymer (Pebax) with different MXene loadings of 0.05, 0.1, 0.15, 0.2, and 0.3 wt %. When MXene loading increased to 0.15 wt %, MMM displayed a uniform structure without aggregation. However, further increase in MXene loading led to the agglomeration of MXene nanosheets and hence the hindrance in the molecular transport through the membrane. Therefore, the greatest CO2 permeance and CO2/N2 selectivity were reported for MMM with MXene loading of 0.15 wt % as ∼22.23 GPU and ∼69.2, respectively. Compared to the Pebax membrane, MMM with MXene loading of 0.15 wt % exhibited around an 81 and 73.4% increase in CO2 permeance and CO2/N2 selectivity, respectively. Similarly, to investigate the interaction between MXene and Pebax and reveal how their assembly affects the separation performance, Shamsabadi et al.161 fabricated MMMs with MXene loadings of 0.05, 0.075, 0.1, 0.2, and 0.5 wt %. Considerable enhancement in both CO2 permeance and CO2/N2 selectivity was observed from 987 to 1820 GPU and from 32.1 to 42.0 with the incorporation of MXene loadings of 0.1 wt % into the Pebax polymer, respectively.161 They also compared the gas separation properties of MXene- and GO-based MMMs, and MXene-based MMMs were suggested as the optimum membranes for CO2 capture due to having enhancement in both separation properties while GO-based MMMs revealed an increase in only selectivity. Additionally, MD simulations were carried out to gain a depth of understanding of the interactions between MXene and Pebax. MD simulations revealed that strong interactions between hard segments of the Pebax and surface functional groups of MXene arose from the hydrogen bonding. This led to the increase in compatibility and homogeneous dispersion of MXene nanosheets in the polymer matrix and, hence, the improvement of CO2/N2 separation.161 Compared to low MXene loadings, Guan et al.162 increased it up to 1 wt % and Shi et al.163 even further enhanced the MXene content in a Pebax continuous phase up to 20 wt %. Similarly, they demonstrated the use of a certain optimal MXene loading where the permeability reaches a peak value. It was 0.5 wt % for the MMM synthesized by Guan et al.162 with CO2 permeability of 70.2 Barrer and CO2/N2 selectivity of 93.2 at 4 bar and 25 °C. Although high MXene loading was handled by Shi et al.,163 optimal MXene loading was identified as 1 wt % with CO2 permeability of 148 Barrer and CO2/N2 selectivity of 63 at 2 bar and 30 °C. However, for the humidified MMM, this weight content of MXene in a MMM increased to 10 wt % with a superior performance (584 Barrer CO2 permeability and 59 CO2/N2 selectivity) compared to that of dry MMM.163 Considering this limitation in MXene loading, to improve compatibility or in other words to remove the nonselective voids between MXene and polymer phase, the transformation of MXene into nanoscale ionic materials and then fabrication of MMM were proposed. With the surface functionalization of MXene which led to the enhancement in electrostatic interaction with the polymer, Wang et al.164 dispersed a great amount of MXene in a polymer phase and ended up with high-performing MMM displaying 176% and 29% increases in CO2 and CO2/N2 selectivity compared to the Pebax membrane, respectively. MXene nanosheets were also dispersed within the poly(ethylene glycol) (PEG) phase165 rather than the frequently preferred Pebax one. They revealed greater CO2 permeability as 1912 GPU at 1 bar and 25 °C and lower binary CO2/N2 selectivity as 31.2 than did the Pebax-based MMMs.165 As it was experienced from the reported studies including other inorganic fillers in MMMs, there are several approaches to handle the voids between filler and polymer. Therefore, they should be tested for the MXene-based MMMs, rather than just changing its content in polymer phase. Alternatively, there are high technology polymers that can be used as a continuous phase. Since the above-mentioned studies evidently prove the capability of MXene-based MMMs in gas separation, with a proper harmony between the high-performing polymers and MXene nanomaterials, novel MMMs can be designed for the target gas separation application.

In addition to MMMs, nanocomposite membranes were also fabricated with the combination of MXene nanomaterial with other inorganic materials such as MOF. Hong et al.166 synthesized a Ti3C2Tx (MXene)/ZIF-8(MOF) dual-layered nanocomposite membrane within 21 min with an active membrane layer of 525 cm2 using a novel approach. They initially used electrophoretic deposition (EPD) to accumulate MXene nanosheets (thickness of approximately 0.8 μm) onto the copper disc for 1 min and then synthesized a ZIF-8 layer (thickness of approximately 0.45 μm) onto the MXene layer within 20 min using a fast current-driven synthesis (FCDS) approach. Gas permeances of H2 and CO2 were 379 and 8.4 GPU for pristine MXene membrane, whereas those were 178.2 and 2.3 GPU for the Ti3C2Tx/ZIF-8 nanocomposite membrane, respectively.166 However, the average ideal selectivity of H2/CO2 for the nanocomposite membrane increased from 44.4 to 77.4 compared to the pristine MXene membrane.166 In addition to the alteration of the attractive environment for gas molecules in the nanocomposite membrane compared to the pristine MXene membrane, this performance was attributed to the size change of pores existing for gas transport. Because the interlayer distance of MXene was 3.7 Å and the pore aperture of ZIF-8 was 3.1 Å. In situ synthesis of MOF-801 (pore aperture of 4.7 Å) on the MXene nanosheet was also followed by Li et al.,167 leading to the formation of MOF crystals within MXene channels contrary to Hong et al.,166 where MOF crystals were grown on the MXene surface. Then, its superiority in H2/CO2 separation was verified by comparing with a pristine MXene membrane and Ti3C2Tx/MOF-801 nanocomposite membrane fabricated with a well-known physical intercalation method. H2 permeance was improved in both in situ (202%) and intercalated (136%) Ti3C2Tx/MOF-801 nanocomposite membrane compared to the pristine MXene membrane (773 GPU).13 However, due to the local agglomeration of MOF crystals within MXene nanosheets in the nanocomposite membrane fabricated by the intercalation, a severe drop of H2/CO2 selectivity (5) was observed, contrary to the selectivity improvement (29.4) in the in situ synthesized nanocomposite membrane. Building on this contradictory trend in H2 permeance of MXene/MOF nanocomposite membranes fabricated in two different research groups, this might be aroused from the growth of MOF crystals either within MXene channels or on the surface of MXene layer. Further elaboration of this issue is required by testing different MOFs.

Not only are outstanding gas separation properties of membranes essential but also is their long-term stability required for use in industry. Therefore, long-term stability of MXene membranes in gas separation performance was tested by several studies in the order of hours, days, and weeks.129,144147,153,156,161,165,166 Ding et al.129 reported that the MXene membrane displayed stable performances with H2 and CO2 permeability of ∼2210 and ∼15.9 Barrer, respectively, and H2/CO2 selectivity of ∼147 up to 70 h with shifting between a dry and wet (3 vol % steam) equimolar H2/CO2 mixture at 25 °C and 1 bar. More importantly, to confirm high temperature durability of MXenes, Fan et al.144 applied long-term permeation measurements for H2/N2 separation at 320 °C and reported stable H2 permeance of 612 GPU and the H2/N2 mixture selectivity of 41 during 200 h. Shen et al.153 suggested that as-prepared MXene with a thickness of 0.02 μm and functionalized MXene with borate and poly(ether imide) (PEI) preserved their separation performances of H2/CO2 and CO2/CH4 within 100 h at 1.5 bar and 25 °C, respectively. Likewise, MXene/Pebax MMMs with MXene loading of 0.15 wt % were examined for long-term operation in a mixed-gas system at 25 °C and 1 bar by Liu et al.160 Effectiveness of MXene nanosheets into the Pebax polymer and the stability of gas separation properties of MMMs were revealed by the stable CO2 permeance at 21.6 GPU and CO2/N2 selectivity at 72.5 up to around 120 h.160 Shamsabadi et al.161 tested the stability of MXene/Pebax MMMs with a MXene loading of 0.05 wt % at 25 °C and 4 bar for 2 weeks. MMM maintained its CO2/N2 selectivity of ∼40 up to 15 days, whereas its CO2 permeance decreased from ∼1800 GPU to ∼1400 GPU due to the rapid physical aging of highly permeable poly[1-(trimethylsilyl)-1-propyne] (PTMSP) used as a gutter layer. Interestingly, Wang et al.164 fabricated the MXene (Ti3C2Tx) nanoscale ionic materials (NIMs) with a novel synthesis approach benefitted from the electrical interaction, which displayed excellent antioxidation stability after 570 days (stored in air with a tight cap at room temperature), whereas pristine MXene was oxidized within 1 week. Collectively, the stability of MXene membranes under aggressive conditions, high temperature measurement, and long operation times were introduced by several studies.

The other requirement for membranes to be used in industry is to be resistant to the humidity. However, most of the inorganic nanomaterials such as MOFs lose their intensity when they are exposed to the humidity. Therefore, it is a serious challenge in inorganic membranes. Its effect on MXenes’ performance or stability have not been identified properly in the literature, except for a few studies. Li et al.154 studied hydrous and anhydrous MXenes to reveal the effect of water intercalation on gas diffusion via molecular simulation methods. Significant alteration in gas diffusion was observed for gas molecules having a large kinetic diameter at a short interlayer distance. For instance, for the interlayer spacing ranging between 6 and 14 Å, the ratios of diffusion coefficients of anhydrous MXene over hydrous MXene varied between 15 and ∼1.5 for H2, and 650 and ∼1.7 for CH4. This was attributed to the blockage of water molecules for the transport of gases due to dividing transport channels into the small parts resulting in narrow paths for large gas molecules. Furthermore, they displayed that diffusion coefficients of H2 and CO2 were decreased with a water content increase from 0 to 2.4 wt %. Finally, Li et al.154 highlighted that water molecules intercalated within nanosheets enhanced the sieving effect of the MXene membrane, yielding an almost one order of magnitude greater diffusion selectivity than anhydrous MXene. Shen et al.153 experimentally studied the humidity effect on gas permeation for ultrathin MXene nanosheets with a 0.02 μm thickness. They reported that H2 permeance decreased from 1030 to 618 GPU and CO2 permeance increased from 52.3 to 64.3 GPU with a relative humidity increase from 40 to 90%, leading to a decrease in mixed H2/CO2 selectivity from 19.7 to 9.8. These results were consistent with the simulation data computed by Li et al.154 displaying the severe hindrance in H2 permeation with the rise in humidity of MXene. However, there is still need for detailed elaboration of the effect of humidity on MXene gas separation performance.

2.2. Adsorption-Based Separation

From the discovery of first MXene member, more than 46 different MXene structures168 were synthesized for different purposes and almost 730 structures157 were theoretically investigated. However, only one of them was dominantly examined for membrane-based gas separation. Unlike this common MXene type, Ti3C2Tx, there were few studies which investigated the other MXene types149151 experimentally for the adsorption-based gas separation. Liu et al.149 fabricated two different MXene types (Ti3C2Tx and Ti2CTx) by etching with different salts such as LiF, sodium fluoride (NaF), potassium fluoride (KF), and ammonium fluoride (NH4F) in 4 mol/L salt solutions and then intercalated with DMSO, NH3·H2O (ammonium hydroxide), and urea to investigate the CH4 adsorption capacity of MXenes. Highest CH4 adsorption capacities were reported as 8.5 cm3/g for Ti3C2Tx etched with LiF and 11.6 cm3/g for Ti2CTx etched with KF. Since the Li+ ion was smaller than the other ions used for etching, it was easily adsorbed and kept on the surface of the MXene, leading to higher CH4 adsorption and lower CH4 desorption due to the strong interaction between CH4 and Li+. On the contrary, MXenes exfoliated by NaF and KF released greater amounts of adsorbed methane in the desorption process compared to MXenes exfoliated by LiF, due to the fact that few Na+ and K+ ions were located on the surface of the MXenes. Liu et al.148 examined both theoretically and experimentally the effect of intercalation of NH3·H2O on the CH4 adsorption capacity of MXene (Ti2C) under different temperatures. CH4 adsorption capacities were measured as 11.58, 18.18, and 52.76 cm3 (STP)/g at 25, 40, and 50 °C and 50 bar for as-synthesized MXene, respectively, indicating the positive effect of temperature on adsorption. On the other hand, CH4 adsorption improved with the intercalation of NH3·H2O to 16.81 cm3 (STP)/g at 25 °C and 50 bar due to the increase in interlayer spacing. Moreover, theoretical (22.9 wt %) and experimental (28.8 wt %) CH4 adsorption capacities agreed well with each other. In addition to CH4 adsorption performance, CO2 capture of MXene was also identified. Wang et al.150 experimentally studied the CO2 adsorption performance of Ti3C2Tx and V2CTx, which were produced using NaF and HCl as etching solutions at different temperatures. Then, as-synthesized MXenes (as-Ti3C2Tx and as-V2CTx) were intercalated with DMSO. This led to the increase in SSA from 21 to 66 m2/g for the intercalated Ti3C2Tx (int-Ti3C2Tx) and from 9 to 19 m2/g for the intercalated V2CTx (int-V2CTx). CO2 adsorption capacities of Ti3C2Tx and V2CTx were increased from 1.33 to 5.79 mol/kg and from 0.52 to 0.77 mol/kg at 40 bar, respectively, after the intercalation. They emphasized the very exiting conclusion that if Ti3C2Tx with theoretical SSA (%100 exfoliated) can be made in the future, the theoretical capacity would be 44.2 mol/kg. On the other hand, they introduced difficulty in the synthesis of V2CTx and proposed the fabrication of a V2CTx adsorbent with multilayers rather than a single layer in order to increase its adsorption capacity. However, in a different study, without intercalation, CO2 adsorption capacities of Ti3C2Tx were measured as 0.603 mol/kg at 10 bar and 25 °C.169 Although Petukhov and his co-workers could not reach a high CO2 adsorption capacity for Ti3C2Tx, its superior NH3 adsorption performance was revealed as ∼8.2 mol/kg at the same conditions and was related to the occupancy of −(OH)2 functional groups on the MXene surface.169 By the transformation of Ti3C2Tx into nanoscale ionic materials, its CO2 adsorption capacities were further improved to 1.80 mol/kg at the same conditions.164 Surprisingly, even at severe conditions (0.01 bar and 125 °C) its performance in terms of CO2 capture reached to 12 mol/kg. To observe this superior performance, the novel approach was followed to modify the surface termination of MXene as initially treating at high temperature to desorb F atoms, subsequently exposing to H2 to remove the persistent O atoms from the surfaces, and then exposing to CO2 to synthesize the termination depleted MXene.170 Morales-García et al.151 theoretically studied MXenes formulated as M2C (M = Ti, Zr, Hf, V, Nb, Ta, Cr, Mo, and W) for CO2 adsorption. CO2 capture increased along the series of MXene following the order of Ti > V > Zr > Nb > Mo > Hf > Ta > W, ranging from 2.34 to 8.25 mol/kg at 25 °C, in accordance with the adsorption bonding energies between CO2 and MXene. MXenes exhibited a high CO2 adsorption capacity compared to common adsorbents such as zeolites (Ca-X, 13X), GO, and metal oxides. By this exciting study, CO2 adsorption capacities of several MXene types have been studied using DFT calculations and their real performances have been revealed for the first time. On the other hand, awesome performance of incompletely etched Ti2CTx for H2 storage was revealed by Liu et al.,171 where more than twice the storage performance of previously reported adsorbents under ∼50–60 bar was measured as 8 wt %. Additionally, a positive impact of the narrow interlayer distance and −F2 functional groups were introduced on reversible H2 storage in Ti2CTx.171 It is expected that these pioneer studies will open a new perspective and hopefully accelerate the investigation of synthesis of other MXene types whose gas adsorption performances have not been revealed yet.

In the light of all studies mentioned above, MXene nanomaterials with easily scalable physicochemical properties, unprecedented gas separation capacity, long-term stability, and compatibility with polymeric materials provide an excellent opportunity to develop superior membranes or adsorbents for the gas separation process. For gas separation application, all studies were concentrated on tailoring the interlayer distance of MXene lamellar nanomaterials to obtain zipped in-plane slit-like pores. Especially for the membrane process, to tune the interlayer distance, several strategies were followed with success such as tuning the membrane thickness,129,146,153155 treatment at high temperature,144,155 intercalation with ions,147 modification with CO2 soluble materials like amine153 and DES,159 and combination with polymeric materials.160165 As a result of such efforts, considerable improvements were achieved, and designed 2D MXene nanomaterials offer new avenues especially for membrane development. However, the most critical issue is the verification of long-term performance of MXene membranes. Instead of testing the pristine MXene membrane, measurements were carried out considering MMMs with a limited period up to a maximum of 8 days. We believe that a longer test period is essential to validate the stability and durability of gas separation MXene membranes. In gas separation applications, the humidity has severe adverse effects on the separation performance of nanomaterials like MOFs. CO2 separation improvement with the humidity was revealed with the limited number of studies in MXene membranes.153,154 However, further investigation is required for the verification of its effect on performance. Concerning the adsorption process, unlike the general tendency in the membrane process, different types of MXene nanomaterials rather than Ti3C2Tx were identified.149151 It is an encouraging manner to experience different members of MXene for gas separation. Collectively, to display the gas separation capacity of the MXene family, Table 1 and Table S1 are tabulated for MXene membranes and adsorbents, which are discussed in this section, respectively. Although the MXene family is still in its initial stage in membrane- and adsorption-based gas separation applications, more efforts should be performed in order to reveal the separation performances of all MXene family via theoretical and experimental studies.

Table 1. Survey of Gas Separation Performance of MXene-Based Membranesa.

membranes operating conditions H2 CO2 N2 CH4 unit H2/CO2 H2/N2 H2/CH4 CO2/CH4 CO2/N2 ref
Ti3C2Tx on AAO (0.005 μm) 1.5 bar, 25 °C 36000 7200     GPU 5         (153)
Ti3C2Tx on AAO (0.02 μm) 1.5 bar, 25 °C 1587 51.18 326.8 279.5 GPU 29.19 4.86 5.68 0.183 0.157 (153)
Ti3C2Tx on AAO (0.08 μm) 1.5 bar, 25 °C 350 11.29     GPU 31         (153)
Ti3C2Tx-borate @ 55 °C (0.02 μm) 1.5 bar, 25 °C   436.6   102.3 GPU       4.27   (153)
Ti3C2Tx-borate @ 75 °C (0.02 μm) 1.5 bar, 25 °C 291.4 322.4 52.9 42.4 GPU 0.9 5.5 6.9 6.1 7.6 (153)
Ti3C2Tx-borate @ 95 °C (0.02 μm) 1.5 bar, 25 °C   91.19   17.88 GPU       5.10   (153)
Ti3C2Tx-borate-PEI @ 75 °C (0.02 μm) 1.5 bar, 25 °C 246.7 349.5 28.9 22.8 GPU 0.7 8.5 10.8 15.3 12.1 (153)
Ti3C2Tx on AAO (0.8 μm) 1 bar, 25 °C 1209   54.9 110.4 GPU   22 11     (144)
Ti3C2Tx on AAO (0.8 μm) 1 bar, 320 °C 890   21.8 149.3 GPU   41 5.96     (144)
Ti3C2Tx on AAO (0.2 μm) 1 bar, 25 °C 2176.43 850.57 883.25 766.51 Barrer 2.56 2.46 2.84     (129)
Ti3C2Tx on AAO (2 μm) 1 bar, 25 °C 2402.3 10.1 18.6 3.08 Barrer 238 129 780     (129)
Ti3C2Tx on AAO (5.1 μm) 1 bar, 25 °C 1796.43 7.12 8.35 2.22 Barrer 252.30 215.14 809.20     (129)
Ti3C2Tx on AAO (0.02 μm) 1 bar, 22 °C 4090   201.5   GPU   20.3       (156)
Ti3C2Tx on AAO (0.8 μm) 1 bar, 22 °C 698.5   51.9   GPU   13.4       (156)
Ti3C2Tx on AAO (1 μm) 1 bar, 22 °C 477.6   24.2   GPU   19.8       (156)
Ti3C2Tx (15 wt %)/Pebax on PAN (2 μm) 2 bar, 25 °C   22.23 0.321   GPU         69.2 (160)
Ti3C2Tx (0.05 wt %)/Pebax on PVDF (60 μm) 4 bar, 25 °C 325.6 1986.5 47.5 134.2 GPU       14.8 41.8 (161)
Ti3C2Tx (0.075 wt %)/Pebax on PVDF (60 μm) 4 bar, 25 °C 313.9 1915.0 46.7 129.4 GPU       14.8 41.0 (161)
Ti3C2Tx (0.1 wt %)/Pebax on PVDF (60 μm) 4 bar, 25 °C 312.1 1810.3 43.2 120.7 GPU       15.0 42.0 (161)
Ti3C2Tx-Ni2+ inter. on Al2O3 hollow fiber (2.7 μm) 1 bar, 25 °C 321 0.53 0.72 1.55 GPU 750 430 240     (147)
Ti3C2Tx-Pb2+ inter. on AAO (0.78 μm) 1 bar, 25 °C 794       GPU 242         (145)
Ti3C2Tx (0.5 wt %)/Pebax (50 μm) 4 bar, 25 °C 5 70.24 0.92 2.5 Barrer       28 93.18 (162)
Ti3C2Tx on PC (2 μm) 0.6 bar, 25 °C 894.86 223.81 303.46 422.94 GPU       0.1058 0.083 (159)
Ti3C2Tx-DES inter. on PC (3 μm) 0.6 bar, 25 °C 2.128 26.35 0.083 0.106 GPU       249.01 319.15 (159)
Ti3C2Tx (1 wt %)/Pebax (dry) (0.001–0.002 μm) 2 bar, 30 °C   148 2.35   Barrer         63 (163)
Ti3C2Tx (10 wt %)/Pebax (humidified) (0.001–0.002 μm) 2 bar, 30 °C   584 9.9   Barrer         59 (163)
Ti3C2Tx @ 140 °C on YSZ hollow fiber (0.05 μm) 1 bar, 25 °C 6193 2670     GPU 2.32         (146)
Ti3C2Tx @ 140 °C on YSZ hollow fiber (0.22 μm) 1 bar, 25 °C 70.6 2.33     GPU 30.3         (146)
Ti3C2Tx (25 wt %)/PEG (400) (0.5 μm) 1 bar, 25 °C   1543     GPU       25.39 30.9 (165)
Ti3C2Tx (25 wt %)/PEG (600) (0.5 μm) 1 bar, 25 °C   1627 51 58.38 GPU       27.87 32.18 (165)
Ti3C2Tx-NIM (60 wt %)/Pebax (80–130 μm) 1 bar, 25 °C   91.9 1.58   Barrer         58.2 (164)
Ti3C2Tx (1 wt %)/CTA (∼50 μm) 1.5 bar, 25 °C   7   0.21 Barrer       34   (192)
Ti3C2Tx (3 wt %)/CTA (∼50 μm) 1.5 bar, 25 °C   16   0.28 Barrer       57.14   (192)
Ti3C2Tx on PES (0.67 μm) 1 bar, 25 °C 379 8.4     GPU 44.4         (166)
Ti3C2Tx/ZIF-8 on PES (0.45 μm) 1 bar, 25 °C 178.2 2.3     GPU 77.4         (166)
Ti3C2Tx on PTFE 1 bar, 25 °C 773 45.5     GPU 17         (167)
Ti3C2Tx/MOF-801 (intercalated) on PTFE 1 bar, 25 °C 1824 358     GPU 5.1         (167)
Ti3C2Tx/MOF-801 (in situ growth) on PTFE 1 bar, 25 °C 2334 79.4     GPU 29.4         (167)
a

Membrane thickness is given in parentheses. AAO: anodic aluminum oxide, DES: deep eutectic solvent, NIM: nanoscale ionic materials, PAN: polyacrylonitrile, PC: polycarbonate, Pebax: polyether-polyamide block copolymer, PEG: poly(ethylene glycol), PES: polyether sulfone, PTFE: poly(tetrafluoroethylene), PVDF: polyvinylidene fluoride, and YSZ: yttria-stabilized zirconia.

Rather than gas separation and gas capture research, a great number of studies are also concentrated on gas sensing performance of MXenes. Selective gas and vapor sensing of MXenes were studied either experimentally or theoretically in the literature where some examples are as follows: Ti3C2Tx172 for NH3, SO2, H2S, NO, and CO2; V2CTx173 for C2H5OH, C3H6O, NH3, CH4, H2, and H2S; thin film Ti3C2Tx174 for N2, CO2, and C2H5OH; Ti3C2Tx175,176 and Ti3C2Tx/polyaniline177 for C2H5OH, CH3OH, C3H6O, and NH3; Ti2C,178,179 V2C,179 Nb2C,179 and Mo2C179 for NH3, H2, CH4, CO, CO2, N2, NO2, and O2; V3C2180 and Nb3C2180 for N2; cation-intercalated Ti3C2Tx181 for NH3; Sc2CO2182 for SO2; tyrosinase-Ti3C2/chitosan183 for C6H6O; Ti3C2Tx with Au electrodes184 and CTAB-delaminated Nb2CTx185 for C2H5OH, C3H6O, NH3, C3H8O, NO2, SO2, and CO2; 3D Ti3C2Tx framework186 for volatile organic compounds (VOCs); Ti3C2Tx/W18O49187 for ppb-level detection of C3H6O; and Ti3C2Tx/Fe2(MoO4)3188 for low working temperature detection. Remarkable sensing performances of MXene nanomaterials for several gas types were ascribed to its high surface area, unique structure, and good conductivity. It is not the topic of this review; for detailed information, please refer to these comprehensive reviews.189191

3. Solvent Dehydration

In many different industries, organic synthesis is carried out in a solvent environment to produce high value-added products. Therefore, recovery and recycle of high amounts of organic solvents remaining at the end of the chemical reaction are required for the economic development and environmental protection. Both the recovery of a high-value solute from a dilute solution (solute enrichment) and the solvent by removing an impurity dissolved in it (solvent recovery) are achieved successfully by organic solvent nanofiltration (OSN) membranes. In this section, only solvent recovery performance of MXene OSN membranes will be discussed in terms of solvent permeance. On the other hand, water-solvent mixtures (aqueous solutions) were used in some chemical reactions to reduce the cost and save the environment rather than using pure solvents. To separate this mixture, pervaporation membranes were fabricated with a purpose of selectively removing water in a vapor phase.

Nowadays, OSN membranes are mostly dominated by organic polymer materials. As an alternative, inorganic membranes were proposed due to their large surface area, high stability in many conditions, and great hydrophilic properties. Recently, MXenes are counted in these inorganic membranes due to their superior hydrophilicity and ultrafast molecular transport. Unique solvent separation performance of MXene membranes was first provided by Wang et al.128 They produced Ti3C2Tx and GO membranes with varying membrane thicknesses to demonstrate the usage of rigid, regularly stacked nanosheets for the solvent separation. Water permeances of MXene with 0.23 μm and GO with 0.21 μm in solvated (rehydrated) state were reported as 2302 (1703) and 257 (129) L/m2 × h × bar, respectively.128 In addition to these exceptional water permeances of MXene membranes, unique solvent permeances, which were much greater than the highest reported values in the literature, were also observed for MXenes with two different membrane thicknesses. Solvent permeances of MXene were in the order of acetonitrile-acetone (5022.4) > methanol (3563) > ethanol (1916) > dimethylformamide (1616) > 2-propanol (983 L/m2 × h × bar) as given in Figure 3(a).128 Similarly, Kang et al.193 reported that the permeances of MXene with the thickness of 0.09 μm were 25, 6.62, 3.17, 2.14, and 0.79 L/m2 × h × bar at 5 bar for water, hexane, toluene, c-hexane, and isopropanol (IPA), respectively, indicating superior selectivity of MXene for water-IPA mixture due to the existing hydrogen bonding between IPA and Ti3C2Tx. Although the solvents have different sizes, these two preliminary studies revealed the importance of the interaction between MXene and solvents in solvent separation. Since it has been demonstrated that MXene membranes display superior solvent permeances compared to the well-known 2D inorganic membranes such as GO and graphene, Wei et al.194 targeted to fabricate a composite lamellar membrane of Ti3C2Tx and GO with various MXene loadings of 0, 50, 60, 70, 80, and 100 wt % in order to optimize the solvent separation performance of inorganic membranes. The solvent flux of the nanocomposite membrane increased with MXene loadings. Nanocomposite membranes with MXene loading of 70 wt % displayed nearly ten times higher acetone flux as 48.32 L/m2 × h than GO membranes, followed by the methanol, ethanol, and IPA fluxes of 25.03, 10.76, and 6.18 L/m2 × h, respectively. Qin et al.195 evaluated the separation performances of MXene, hexagonal boron nitride (hBN), and GO for the water–ethanol mixture by identifying the microstructure of solvent confinement within the nanochannels via MD simulations. They suggested that while ethanol molecules were observed as an accumulated layer near the solid surfaces of all 2D membranes, water molecules exhibited balanced distribution within the slits, as represented in Figure 3(b). Diffusion coefficients of water molecules for Ti3C2(OH)2, hBN, and GO in the subcontact layer were reported as 0.85, 0.81, and 0.60 × 10–9 m2/s, respectively.195 Since the average numbers of hydrogen bond (HB) per water molecules between the interfacial and sub contact layers for Ti3C2(OH)2 (∼2.5) was greater than both hBN and GO (<0.8), Ti3C2(OH)2 was suggested as the best membrane within the investigated membranes for solvent separation.195 Hereby, it has been evidently demonstrated that MXene membranes have a promising potential for solvent separation (see Tables S2 and S3).

Figure 3.

Figure 3

Solvent dehydration performance of MXene membranes. (a) Solvent permeances against the combined solvent property for MXene membranes with two different thicknesses. Adapted with permission from ref (128). Copyright 2018 Wiley Online Library. (b) (Left) Density profiles of methyl (blue line) and hydroxyl groups (yellow line) in ethanol molecules, and oxygen atoms (red line) in water molecules within the hBN, GO-0.2, GO-0.4, and Ti3C2(OH)2. Adapted with permission from ref (195). Copyright 2020 Elsevier. The light blue and orange regions denote the interfacial contact and the subcontact layers, respectively. (Right) The snapshots show the front view of the slits. The solvent fluxes of (c) PEI-based and (d) PDMS-based membranes under 10 bar. Adapted with permission from ref (204). Copyright 2017 Elsevier. (e) Detailed comparison of MXene/SA MMMs with various reported 2D nanomaterial-based membranes for ethanol dehydration performance. Adapted with permission from ref (200). Copyright 2020 Elsevier.

To enhance the solvent-specific separation performance, different strategies were developed for the MXene membranes used in pervaporation and OSN processes. Liu et al.196 embedded macromolecules into MXene nanosheets to investigate its solvent dehydration performance via the pervaporation process. They introduced hyper-branched polyethyleneimine (HPEI) into the Ti2CTx nanosheets and then carried out interfacial polymerization. Its water fluxes for methanol, ethanol, and IPA dehydration systems were reported around 2240, 1430, and 1020 g/m2 × h with water contents of 12.6, 82.4, and 99.2 wt %, respectively.196 The comparison of separation properties between pristine and intercalated Ti2CTx membranes for IPA dehydration provided that while the water content of the intercalated Ti2CTx membrane was significantly greater than that of the pristine one (11 wt %), the water flux of the pristine Ti2CTx membrane (6378 g/m2 × h) surpassed that of intercalated Ti2CTx, proving the insertion of HPEI molecules resulted in more stacked, regular membranes without defects.196 Liu et al.197 in their further study examined the effect of modification of Ti2CTx with different loadings of positively charged polyelectrolyte, polydiallyl dimethylammonium chloride (PDDA), on dehydration of IPA. Their aim in modification of MXene was to build electrostatic interaction between MXene nanosheets and macromolecules and consequently improve the water affinity of membranes by manipulating the water flux. With the increasing PDDA concentration and MXene loading amounts, the total flux decreased because of the enhanced membrane thickness, and the separation factor first increased and then decreased because of the aggregation of PDDA and improper distribution of MXene.197 Optimum separation performance with the total flux of 1237 g/m2 × h and separation factor of 1932 belonged to the MXene/PDDA membrane with the PDDA loading of 0.2 mg/mL and MXene amount of 288 mg/m2.197 Accordingly, while embedment of macromolecules within the nanochannels increased the flux, the separation performance of the 2D inorganic membrane was hindered. Therefore, the studies on gaining improvement in separation performances were escalated. Wu et al.198 studied the pristine MXene membrane having various thicknesses from 0.5 to 2.0 μm to separate an ethanol–water mixture via a pervaporation process. The separation factor increased from ∼20 to ∼83, and the total flux decreased from ∼592 to ∼221 g/m2 × h for 90% ethanol aqueous solution with membrane thickness, indicating the prolonged mass transfer path with increased mass transfer resistance.

MXene-based MMMs were synthesized with several organic polymers to investigate the improved solvent dehydration performance in the pervaporation process. Xu et al.199 fabricated Ti3C2Tx/chitosan (CS) MMMs with different MXene loadings of 0, 1, 3, and 5 wt % and evaluated the separation of three typical azeotropic mixtures: water-ethanol, water-ethyl acetate, or water-dimethyl carbonate. The total flux and ethanol separation factor of MMMs increased with MXene loading up to 3 wt % at 50 °C from ∼1150 to 1424 g/m2 × h and from 407 to 1421, respectively. For 98 wt % water-ethyl acetate and water-dimethyl carbonate mixtures, solvent separation factors of MMM with a MXene amount of 3 wt % were totally different, reaching to 4898 and 906 at 50 °C, respectively. Li et al.200 investigated the pervaporation performance of Ti3C2Tx/sodium alginate (SA) MMMs with MXene amounts of 0, 0.06, 0.12, 0.18, and 0.24 wt % for ethanol dehydration. Lower optimum MXene loading was observed for the system of Ti3C2Tx/SA than Ti3C2Tx/CS as 0.12 wt % with a separation factor of 9946 and total flux of 514 g/m2 × h.200 Since the hydrophilicity of membrane increased with MXene content, the interaction of water molecules with a membrane matrix enhanced through strong hydrogen bonds, leading to an increase in the adsorption of water and confinement of ethanol molecules. Compared to the other modified SA membranes for pervaporation performance, MXene/SA MMMs exhibited the highest separation factor [see Figure 3(e)] but comparable water flux, probably due to the uniform distribution of MXene sheets within the matrix and excellent compatibility as a result of cross-linking. Similarly, Cai et al.201 also studied the effect of MXene amounts of 0, 0.5, 1, 2, 3, and 4 wt % for ethanol dehydration in Ti3C2Tx/poly(vinyl alcohol) (PVA) MMMs. As the amount of MXene increased from 0 to 3.0 wt %, total flux decreased from 97 to 75 g/m2 × h and the separation factor increased from 144 to 2585, again suggesting an optimum MXene content in MMMs which was related to the improved cross-linking density. A greater separation factor was reported by Yang et al.202 for Ti3C2Tx/PVA MMM cross-linked with sulfosuccinic acid (SFA). MMM including 20 wt % SFA and 2 wt % MXene revealed water content in the permeate as 97.6, 99.5, 99.7, and 99.9 wt % with separation factors of 968, 4738, 7913, and 23,786 for methanol, ethanol, isopropanol, and tert-butanol aqueous solutions, respectively. However, compared to the uncross-linked one,201 ethanol dehydration performance was diminished, displaying a water flux of 1489 g/m2 × h and a separation factor of 4738.202 These studies paved a new way toward fabricating MMMs in order to combine favorable properties of MXene and the polymeric phase, thereby enhancing flux without sacrificing separation factor for solvent dehydration. Although the solvent separation performance of MXene is promising for pervaporation membranes, more studies are required for the complete understanding.

Similarly, in the OSN process, to improve the separation performance of MXene membranes, either surface functionalization to design the interlayer spacing of MXene or fabrication of composite membranes by incorporating MXene nanomaterials into the polymeric phase were performed. The experts from Zhengzhou University published successive papers related to these strategies. Initially, thin film nanocomposite (TFN) membranes were fabricated using hydrophilic PEI or hydrophobic polydimethylsiloxane (PDMS) polymers and Ti3C2Tx having abundant −(OH)2 groups at various weight ratios. Then, their solvent fluxes [ethanol, isopropanol, butanone (polar) and ethyl acetate, toluene, n-heptane (nonpolar)] were examined.203 With the increase of MXene content in both hydrophilic and hydrophobic composite membranes, flux for nonpolar solvents was reduced, whereas it was initially increased and then either decreased or kept stable for polar solvents. For instance, the greatest isopropanol flux for PEI-based TFN membrane was observed at 2 wt % MXene loading as 33.5 L/m2 × h at 10 bar.203 In their following study, they fabricated TFN membranes using the same type of polymers and functionalized Ti3C2Tx with −NH2, −COOR, −C6H6, and −C12H26 groups.204 For both PEI- and PDMS-based TFN membranes, −NH2 and −COOR functionalization led to improvement in polar and decrease in nonpolar solvents compared to unfunctionalized TFN membranes [Figure 3(c–d)]. Then, the same group in another study205 proposed a novel approach for the membrane fabrication and investigated separation of solvent mixtures. Heterostructured membranes were fabricated using hydrophilic pristine Ti3C2Tx and PEI as well as hydrophobic Ti3C2Tx functionalized with a −C6H5 group and PDMS via initial vacuum filtration of MXene onto a support and then coating its surface with polymer. They utilized from this novel membrane formation by selectively capturing polar solvents from the mixture via PEI and then introducing them into hydrophilic MXene nanochannels. This leads to the fast transport of polar solvents and hindered movement of nonpolar solvents, resulting in selective separation of the solvent mixture. Specifically, they reported a toluene separation factor of ∼4.46, 3.31, 2.34, and 2.0 for Ti3C2Tx/PEI membranes from acetonitrile, methanol, acetone, and ethyl acetate solutions, respectively.205 However, those for Ti3C2Tx/PDMS membrane functionalized with the −C6H5 group were 3.41, 4.68, 2.39, and 1.12, respectively.205 Finally, they investigated free-standing MXene membranes functionalized with −NH2, −C6H5, and −C12H25 groups to identify the effect of only functionalization on solvent flux.206 They reported that functionalization with either hydrophilic or hydrophobic groups did not influence the transport of nonpolar solvents, whereas hydrophobic functionalization resulted in the considerable decrease in the flux of polar solvents compared to pristine and hydrophilic functionalized MXene. Using molecular simulation approaches, they proposed that nonpolar solvents interacted weakly with the MXene surface compared to polar solvents and, hence, randomly orientated within the MXene channels, leading to almost similar permeation performance. However, polar solvents displayed ordered molecular alignment especially within hydrophilic MXene nanochannels, resulting in fast transport. Collectively, these studies evidently prove that functional groups attached to the surface of MXene nanosheets are capable of solvent-specific separation with a high solvent flux.

Besides, to overcome trade-off between flux and separation factor in solvent dehydration or solvent mixture separation, the other most important point is their durability/stability/swelling for long-term operation even under harsh conditions. Wang et al.128 proved the excellent permeance of pristine MXene with the thickness of 0.23 μm for water and isopropanol, which were decreased only 4 and 11%, respectively, for 25 h. Likewise, Wu et al.198 examined the pristine MXene membrane with a thickness of 2 μm for ethanol dehydration at room temperature and observed a stable total flux during 48 h. Accordingly, it can be deduced that the thickness of MXene does not affect the membrane stability for solvent dehydration. However, composite membranes with different polymers led to the changes in flux. For instance, for the HPEI intercalated Ti2CTx membrane196 in IPA dehydration, flux decrement was reported around 18% for the 120 h of continuous operation at 50 °C. However, for the PPDA-intercalated membrane197 the flux decline was around 8% within 120 h at 50 °C. On the other hand, solvent flux along 12 h under 10 bar for Ti3C2Tx–NH2/PEI (IPA) and Ti3C2Tx–C12H26/PDMS (n-heptane) composite membranes resulted in 29.3 and 31.0% drops, respectively.204 This situation is more promising in MMMs where stable total flux of Ti3C2Tx/CS199 for ethyl acetate dehydration was observed within 30 h and Ti3C2Tx/PVA201 preserved its flux and separation factor for ethanol dehydration about 7 days. Because of their superior flux stability for long-term operations, MXene membranes emerged as promising inorganic materials in solvent separation processes.

Concerning the OSN and pervaporation MXene membranes applied for solvent dehydration, it successfully benefited from the regularly stacked form of MXene nanosheets. For this separation application, primarily importance was claimed to be the interactions between MXene and the solvents. Therefore, to tailor these interactions, two main attitudes were followed such as intercalation of bulky and charged molecules within the MXene nanochannels and production of composite or mixed matrix membranes via combining MXene with different polymers. The former enables us to benefit from the electrostatic interaction between them via a charged surface and to tune the interlayer distance specific to the solvent molecule via its molecular weight. However, the highly studied strategy is the latter one, which utilizes the polymer’s performance. Although pristine MXene membranes were preferred due to enabling the fabrication in a form of few-nanometer nanosheets and hence the reduction in the mass transfer resistance, flux decrements were recorded for them. However, MMMs or composite membranes have an outstanding potential in practical use of solvent separation, as a result of their high stability in long-term operation even under harsh conditions. Therefore, it is not surprising to focus on MXene-based MMMs or composite membranes on solvent separation.

4. Dye Removal

Considering that more than several tons of dyes are produced annually by textile, paint, and pigment industry and nearly 10–15% of dyes are released to the water, dyes are the major pollutants for water and the environment. Thereby, removal of dyes from wastewater safely and effectively is the vital importance for a sustainable green ecosystem. The removal methods of dyes from water are the physical, chemical, and biological methods such as coagulation, flocculation, membrane filtration, adsorption, ion-exchange, oxidation, electrochemical process, photocatalysis, and biodegradation. Among these processes, membrane and adsorption technologies are the most demanding processes due to their high efficiency and operation simplicity. Inorganic nanomaterials have gained value in the application of these two technologies with their strength, high surface area, and low mass. Since MXenes possess a layered structure providing large and highly accessible surfaces,207,208 the use in dye removal as an inorganic membrane and adsorbent material emerged, and their current performances are tabulated in Table S4 and Table 2, respectively. Within this context, to benefit from short transport pathways and large amounts of nanochannels of MXenes in order to enhance water transport, Ding et al.127 fabricated MXene (Ti3C2Tx) membranes by intercalating Fe(OH)3 nanoparticles. They investigated the rejection of several dyes such as rhodamine B (RhB), 5,10,15,20-tetrakis(n-methyl-4-pyridyl)-21,23-h-porphyrin tetratosylate (TMPyP), and Evans Blue (EB). The MXene membrane prepared by embedding Fe(OH)3 at the initial step and then removing Fe(OH)3 exhibited extremely high water permeance, 1084 L/m2 × h × bar and moderate EB rejection rate, 90% due to the enlargement of interspace between nanosheets by the insertion of Fe(OH)3. Similar dye separation performance was also observed for RhB (805 L/m2 × h × bar, 85%) and TMPyP (921 L/m2 × h × bar, 93%).127 The observed superior dye separation performance for MXene membranes compared to the existing nanofiltration membranes [see Figure 4(a)] encouraged the membrane community and advanced membrane technology. On the other hand, before the application of MXene as a membrane material, it was used as an adsorbent to remove dyes from wastewater. For the first time, Mashtalir et al.209 fabricated multilayered MXene (Ti3C2Tx) and used it as an adsorbent to reveal its dye adsorption and degradation performance. They tested methylene blue (MB) (0.05 mg/mL) and acid blue (AB) (0.06 mg/mL) as cationic and anionic dyes, respectively, under dark and UV light. Under UV light, 62 and 81% decreases in concentration were observed for AB and MB, respectively. However, the decrease was only 18% for MB in the dark environment, indicating that UV light promoted the degradation of dyes significantly. Therefore, Mashtalir et al.209 highlighted the photocatalytic property of MXene and showed its high adsorption efficiency for cationic dye, MB. After these preliminary studies, MXene nanomaterials have drawn considerable attention as both membrane and adsorbent materials for dye removal.

Table 2. Survey of Dye Removal Performance of MXene Adsorbentsa.

adsorbent d-spacing/interlayer spacing (Å) surface area (m2/g) test conditions (P: atm, T: °C, pH) adsorbent dosage (mg) dye initial dye concentration (mg/L) adsorption capacity (mg/g) ref
Ti3C2Tx 9.3 1, 25, 7 10 MB 50 21 (247)
Ti3C2Tx (functionalized with −SO3H) 14.3 1, 25, 7 10 MB 50 111 (247)
Ti3C2Tx (DMSO intercalated and hydrated) 20.18 1, 25, 5 12 MB 100 125 (254)
Ti3C2Tx (hydrated) 7.52 1, 25, 5 12 MB 100 78 (254)
Ti3C2Tx (dry) 1.52 1, 25, 5 12 MB 100 7.8 (254)
Ti3C2Tx 9 1, 20, 9 25 MB 10 140 (252)
Ti3C2Tx/PhA (hydrothermal treatment with 12 h) 1, 25, 7 10 MB 12 42.5 (242)
Ti3C2Tx/PhA (hydrothermal treatment with 12 h) 1, 25, 7 10 RhB 6 22.8 (242)
Ti3C2Tx (stirring-assisted) 1, 20, 7 20 MB 5 85 (245)
Ti3C2Tx (UV-assisted with 28 kHz) 1, 20, 7 20 MB 5 130 (245)
Ti3C2Tx (UV-assisted with 580 kHz) 1, 20, 7 20 MB 5 110 (245)
Ti3C2Tx (functionalized with −COOH) 1, 25, — 10 MB 10 39.4 (246)
Ti3C2Tx (functionalized with −COOH) 1, 25, — 10 NR 20 20.2 (246)
Ti3C2Tx (functionalized with −COOH) 1, 25, — 10 ST 30 31.6 (246)
Ti3C2Tx-COOH (treated with (PEI/PAA)10 1, 25, — 10 MB 10 40.4 (246)
Ti3C2Tx-COOH (treated with (PEI/PAA)10) 1, 25, — 10 NR 20 46.1 (246)
Ti3C2Tx-COOH (treated with (PEI/PAA)10) 1, 25, — 10 ST 30 35.6 (246)
Ti3C2Tx (treated with terephthalate) 135.7 1, 20, 7 10 MB 100 209.5 (248)
Ti3C2Tx 20.44 1, 25, 7 100 MB 50 99.9 (249)
Ti3C2Tx (treated with NaOH) 26.2 1, 25, 7 100 MB 50 184.2 (249)
Ti3C2Tx (treated with LiOH) 26.4 1, 25, 7 100 MB 50 118.9 (249)
Ti3C2Tx (treated with KOH) 24.92 1, 25, 7 100 MB 50 74.2 (249)
Ti3C2Tx (decorated with Fe3O4) 1, 25, 7 25 MB 40 2.1 (250)
Ti3C2Tx (decorated with Fe3O4) 1, 40, 7 25 MB 40 4.2 (250)
Ti3C2Tx (decorated with Fe3O4) 1, 55, 7 25 MB 40 11.7 (250)
Ti3C2Tx (traditional etching)b 8.9 100 MB 50 8.5 (258)
Ti3C2Tx (hydrothermal etching)b 44.6 100 MB 50 12.5 (258)
Ti3C2Tx/PEI (modified with SA) 16.31 1, 25, 3 10 CR 150 1300 (262)
Ti3C2Tx 1, 25, 4 10 MB 5 121.6 (263)
Ti3C2Tx(grafted with polyelectrolyte (AMPS-co-AA)) 1, 25, 4 10 MB 5 68.03 (263)
Ti3C2Tx (grafted with polyelectrolyte (DAMPS-co-AA)) 1, 25, 4 10 MB 5 67.88 (263)
Ti3C2Tx 1, 25, 2 40 MB 6.4 64.3 (264)
Ti3C2Tx (modified with SA (30%)) 12.0 1, 25, 7 40 MB 100 92.0 (259)
Ti3C2Tx (functionalized with peroxo) 1, 25, 5.6 25 MB 200 558.0 (253)
Ti3C2Tx (functionalized with peroxo) 1, 25, 5.6 25 RhB 200 524.6 (253)
Ti3C2Tx (functionalized with peroxo) 1, 25, 5.6 25 CR 100 258.2 (253)
Ti3C2Tx (functionalized with peroxo) 1, 25, 5.6 25 MO 100 292.6 (253)
Ti3C2Tx 1, 25, 2 90 MG 10 4.8 (265)
Ti3C2Tx/PDA (functionalized with cellulose) 38.43 1, 25, 7 50 MB 100 112.45 (266)
Ti2CTx 12.8 18.6 1, 35, 6 35 MB 300 544.1 (266)
Ti3C2Tx 15.97 1, 25, 8 4 MR 50 61.56 (267)
Ti3C2Tx 15.38 1, 25, 8 4 MO 50 12.29 (267)
Ti3C2Tx 14.72 1, 25, 8 4 OG 50 1.15 (267)
Ti3C2Tx 12.45 1, 25, 7 CR 90 20 (253)
Ti3C2Tx 12.45 1, 25, 7 MB 90 60 (253)
Ti3C2Tx (alkalized with acrylic acid) 95.51 1, 25, 7 CR 90 265 (253)
Ti3C2Tx (alkalized with acrylic acid) 95.51 1, 25, 7 MB 90 195 (253)
Ti3C2Tx/Co3O4 —, 25, — 10 RhB 5 47.1 (256)
Ti3C2Tx/Co3O4 —, 25, — 10 MB 12.5 128.9 (256)
Ti3C2Tx 4.71 1, 25, 7 20 MB 200 105 (255)
Ti3C2Tx 4.71 1, 25, 7 20 RhB 200 58 (255)
Ti3C2Tx/Fe3O4 8.77 1, 25, 7 20 MB 200 153 (255)
Ti3C2Tx/Fe3O4 8.77 1, 25, 7 20 RhB 200 86 (255)
Ti3C2Tx 10.01 1, 25, — 15 MB 9.0 (257)
Ti3C2Tx/ZIF-8 13.43 1, 25, — 15 MB 107 (257)
V2CTx 7.9 26.6 1, 25, 11 15 MB 20 111.11 (260)
Nb2CTx (traditional etching)b 100 MB 50 3.5 (258)
Nb2CTx (hydrothermal etching)b 100 MB 50 6.5 (258)
Nb2CTx 9.7 44.69 1, 25, 7 100 MO 500 493 (261)
Nb2CTx 9.7 44.69 1, 25, 7 100 MB 500 496 (261)
Ti3C2Tx 1, 30, 6 20 MB 100 63.07 (268)
Ti3C2Tx (modified with EHL (50%)) 1, 30, 6 20 MB 100 102.5 (268)
a

AA: acrylic acid, Alk: alkalized, AMPS: 2-acrylamido-2-methylpropane sulfonic acid, DMAPS: (2-(methacryloyloxy) ethyl] dimethyl-(3-sulfopropyl) ammonium hydroxide, DMSO: dimethyl sulfoxide, EHL: enzymatic hydrolysis lignin, PhA: phytic acid, PAA: poly(acrylic acid), PEI: polyethylene polyimide, SA: sodium alginate. CR: Congo Red, MB: methylene blue, MG: malachite green, MO: methyl orange, MR: methyl red, NR: neutral red, OG: orange G, RhB: rhodamine B, and ST: safranine T.

b

Calculated based on initial dye concentration.

Figure 4.

Figure 4

Performance comparison between MXene membranes and various previously reported membranes for various dyes. (a) EB separation data measured by Ding et al. Adapted with permission from ref (127). Copyright 2017 Wiley Online Library. (b) AY79 separation data measured by Wang et al. Adapted with permission from ref (128). Copyright 2018 Wiley Online Library.

4.1. Membrane-Based Separation

As it was clarified in the previous section, OSN membranes are used for both solvent recovery and separation of high-value solutes such as dyes, pharmaceuticals, etc. from solvents and/or water. However, for simplicity, generally in lab scale tests, solute separation from water rather than solvents is tested for OSN membranes. Especially, since the MXene family is a new type of nanomaterial, the first trials for dye separation were carried using aqueous solutions. To define dye separation performance of membranes, mainly, free-standing MXene membranes were easily fabricated by a vacuum filtration method on various polymeric membrane supports such as polyethersulfone (PES),210 mixed cellulose ester (MCE),211,212 polyvinylidene difluoride (PVDF),213 nylon 66,214,215 etc. Compared to the first study performed by Ding et al.127 where free-standing MXene membrane fabricated by initially embedding and subsequently removing Fe(OH)3 nanoparticles, there are few studies which have found almost the same performances for MXene membranes. Wang et al.128 reported water permeances of Ti3C2Tx for the solvated state and the state of dried and subsequently hydrated as 2302 and 1703 L/m2 × h × bar with acid yellow 79 (AY79) rejection rates of 96.3 and 98.9%, respectively [see Figure 4(b)]. For different MXene type, Hu et al.216 prepared Nb2CTx MXene membranes composed with SA under vacuum-filtration and measured their separation performance for various dyes such as basic blue (BB), rhodamine 6G (Rh6G), and toluidine. Water fluxes of Rh6G, BB, and toluidine were reported as 2164, 2001, and 2209 L/m2 × h × bar, respectively, with ∼100% rejection rates for each dye.216 However, following studies where pure MXene membranes were synthesized revealed almost one order of magnitude lower water permeance with almost the same dye rejection rates for various dyes.210,211 Han et al.210 reported rejection rates of 92.3 and 80.3% for congo red (CR) and gentian violet (GV) with a water flux of 115 L/m2 × h at 1 bar, respectively. Similarly, Zhang et al.211 proposed the same order of water transport property as 120 and 84 L/m2 × h for pure water and MB aqueous solution, respectively. While water permeance decreased to 45 L/m2 × h, rejection rate increased to 100% with the increase in MXene loading at 1 bar.211 Then, to improve water permeance, the intercalation of nanoparticles within MXene nanochannels was examined. For instance, Pandey et al.213 used silver nanoparticles to prepare nanocomposite membranes by varying the silver content (0, 7, 14, 21, 28, and 35 wt %). While pristine MXene membrane displayed comparably lower water fluxes as 90 and 85 L/m2 × h × bar for solutions including RhB and methyl green (MG), respectively, the nanocomposite membrane with 21 wt % Ag exhibited water fluxes of 387 and 354 L/m2 × h × bar without sacrificing the dye rejection rate. Alternatively, He et al.214 synthesized nanocomposite membranes via the intercalation of various amounts of UiO-66, which is a new class of zirconium-based porous MOF, into MXene nanosheets. For different MB concentrations, flux of the nanocomposite membrane including 1.5 mg of UiO-66 was reported around 750 L/m2 × h × bar with the rejection rate greater than 99.2%. However, in the presence of various oils such as n-hexane, isooctane, 1,3,5-trimethyltoluene, and toluene in emulsions, its fluxes decreased to about 443, 488, 328, and 446 L/m2 × h × bar, respectively, without the alteration in dye rejection rates. Long et al.215 inserted Al2O3 nanoparticles in a different amount within the channels of Ti3C2Tx. Excellent improvement (%408) in water permeability of nanocomposite membrane at a 1/1 mass ratio of MXene/Al2O3 was observed increasing from ∼21 to 86 L/m2 × h × bar with a MB rejection rate of 99%. Spectacular design was proposed by Tao et al.217 via the intercalation of carboxymethyl-β-cyclodextrin (CM-β-CD) into Ti3C2Tx channels. Considerable enhancement in total permeance of aqueous MB solution was achieved for Ti3C2Tx/CM-β-CD membrane compared to the pristine MXene from 18.5 up to 431 L/m2 × h × bar (23-fold) with the rejection rate greater than 99% for five different dyes.217 These studies evidently display the importance of intercalation of specific nanoparticles into MXene channels to modulate the interlayer distance and, hence, the flux. Although the performance of membranes fabricated via the intercalation of some nanoparticles within the MXene nanochannels could not exceed those of membranes proposed by Ding et al.,127 it was suggested that they surpassed the water permeance of some commercial and novel nanocomposite membranes.211,213

To further improve separation performance of dye as well as alter the interlayer spacing of MXenes, surface functionalization was applied. Wu et al.206 modified the Ti3C2Tx surface with hydrophilic (−NH2) and hydrophobic groups (−C6H5, −C12H25). Compared to the various dye rejection rates of hydrophilic and hydrophobic MXenes, Ti3C2Tx-NH2 displayed higher rejection rates. Acid–base functionalization of the MXene surface is the other strategy to tailor the interlayer distance.218,219 Yi et al.218 modified Ti3C2Tx with a various number of organic phosphonic acid (OPA) groups. As expected, with the increase in the number of OPA in the functional group, d-spacing of MXene was improved from 12.2 to 16.5 Å. For the MXene functionalized with bulky OPA groups, water fluxes in aqueous solutions including CR (MW: 696.66 g/mol) and eriochrome black T (BT) (MW: 461.38 g/mol) were reported as 514.5 and 508.6 L/m2 × h × bar with the rejection rates of 99.6 and 98.3%, respectively.218 Functionalization with bulky groups were also carried out by the study of Yousaf et al.,212 where they reported the removal efficiency of several dyes in the range of 52–98% for membranes coupled with three different silane agents. Tong et al.219 was aimed not only to alter the interlayer distance by functionalizing with acid–base groups (tannic acid) but also to synthesize the Ti3C2Tx membrane having dye-selective ability by modifying acid groups with multivalent ion salts such as FeCl3, CuCl2, and ZnCl2. Regardless of the type of ion salt, all surface modified MXene membranes exhibited high CR rejection ratios which were above 92% and similar water permeability (∼260 L/m2 × h × bar).219 However, the MXene membrane modified with different multivalent ion salts displayed the best rejection performance for different dye types. Rather than the effect of molecular weight of dye molecules, it was attributed to their charges, where positively charged one was attracted and negatively charged one was repelled from the membrane surface, arising from the negative surface charge of MXene. As highlighted with these preliminary studies, the effect of functionalization of the MXene surface for dye separation is still more complex, where several factors should be considered.

To further enhance the water flux of MXene membranes, instead of fabrication of free-standing membranes, composite membranes via different polymers were proposed. For instance, Pandey et al.220 prepared chemically cross-linked composite membranes consisting of cellulose acetate (CA) and MXene via a phase inversion method. Among the various MXene contents, membranes including 10 wt % MXene over CA revealed pure water flux of 256 L/m2 × h × bar with RhB and MG rejections of 92 and 98%, respectively. Alternatively, the composite membrane prepared by the encapsulation of MXene via silk fibroin displayed greater dye rejection reaching to 99% and water permeance of 324.7 L/m2 × h × bar.221 Additionally, Han et al.222 synthesized chemically cross-linked MMMs using copolyimide (P84) as a continuous phase and MXene as a filler. Total fluxes of MMM containing 1 wt % of MXene for GV and CR were 268 and 380 L/m2 × h × bar, respectively, with a 100% rejection rate for GV and 78.5% for CR. Alternative to the use of commercial and highly preferred polymer type in the fabrication of MMMs, polydopamine (PDA)223 and poly(ionic liquid)224 polymers were also tested combining with MXene as a filler for the separation of several dyes. Due to the synergy between these novel polymers and MXene, which leads to the enhancement in hydrophilicity of the membrane, water flux improved almost twice. Interestingly, poly(ionic liquid)s including counteranions was proposed as a hydrophilicity modifier, which alters the water transport and rejects dyes selectively via a precise selection of the counterion.224 An intriguing approach was also suggested by the group of Wu,225 who fabricated MMM using PES and Ti3C2Tx/ZIF-8 nanocomposite synthesized via a facile one-pot microemulsion strategy. However, the dye separation performance of MMM with 3% microemulsion content was in the order of other MMMs with only MXene. Rather than mixing MXene with a polymer matrix, the final strategy is to intercalate MXene into nanofibers. For instance, Li et al.226 reinforced the Kevlar nanofibers with Ti3C2Tx and observed improved rejection performance for several dyes. The average molecular weight cutoff (MWCO) for the composite membrane was reported as in between 600 and 800 Da considering the analysis based on the different molecular weight of PEG molecules.226 Compared with the free-standing membranes where only the pure MXene was fabricated on a polymeric support,210,211 MXene/polymer composite membranes220224 offered almost twice the water flux with stable dye rejection performance.

On the other hand, it is worth mentioning that although MXene/polymer composite membranes220,222 provided a lower water flux compared to free-standing nanocomposite membranes (Ti3C2Tx/Ag213 and Ti3C2Tx/UiO-66214), they are free from leaching of the MXene layer as a result of weak binding of MXene to the support layer. Therefore, they are comparably more mechanically stable and have constant water transport properties during several operation times. This topic will be discussed in the following paragraphs.

Alternative to silver nanoparticle and porous inorganic nanomaterials such as MOFs that were used to fabricate MXene-based nanocomposite membranes, among the pool of nanomaterials, GO has been remarkably used to prepare a nanocomposite membrane with MXene for the removal of dyes from water. The main reason for this preference was its excellent properties such as large surface area (2600 m2/g), small interlayer spacing, varied functional groups, high mechanical strength, and plain structure. Considering excellent properties of MXene and GO, the design of nanocomposite membranes by combining them has become an alternative strategy in order to obtain high separation performance for dyes. Kang et al.193 prepared a free-standing Ti3C2Tx/GO nanocomposite membrane including 30 wt % of GO on a polycarbonate (PC) support to demonstrate its separation potential for methyl red (MR), MB, rose bengal (RB), and brilliant blue (BB). Total permeances of Ti3C2Tx/GO nanocomposite membranes were 2.1, 0.3, 0.67, and 0.23 L/m2 × h × bar at 5 bar with the rejection rates of 68, 99.5, 93.5, and 100% for MR, MB, RB, and BB solutions, respectively. It was suggested that dye molecules with a hydrated radii larger than 5 Å (MB: 5.04 Å, RB: 5.88 Å, BB: 7.98 Å) displayed lower permeation with higher rejection rates due to the hindrance of transport through nanochannels. Wei et al.194 tested dye separation performance of several Ti3C2Tx/GO nanocomposite membranes prepared by varying MXene loadings as 0, 50, 60, 70, 80, and 100 wt % for five different dye types. With the alteration of the ratio of MXene over GO, fluxes of the nanocomposite membrane for both pure water and aqueous dye solution reversely changed with dye rejection, indicating the requirement of a well-thought-out design for the nanocomposite membrane. Ti3C2Tx/GO nanocomposite membrane with the MXene amount of 70 wt % exhibited the optimum performance for the removal of different dyes with the rejection rates greater than 90% and total flux ranging between 18 and 24 L/m2 × h × bar. Similar behavior was also supported by the study of Liu et al.227 where the separation performances of Ti3C2Tx/GO nanocomposite membranes for five dyes, which were different than those used in the study of Wei et al.,194 were investigated. Similarly, the change of the MXene/GO ratio inversely effected total flux and dye rejection. As a consequent, the optimum performance was observed for the nanocomposite membrane having a MXene/GO ratio of 4/1 as an average total flux of 71.9 L/m2 × h × bar and dye rejection rates greater than 99.5%. A different strategy was applied to improve dye separation performances of Ti3C2Tx/GO nanocomposite membranes by Han et al.228 They fabricated nanocomposite membranes with different MXene loadings and then applied H2O2 to convert MXene into TiO2 nanocrystals via in situ oxidation. However, separation properties of the nanocomposite membrane with optimum performance were reported in the order of the results obtained by Liu et al.227 who did not apply any of the oxidation process. To enhance the adhesion between MXene and GO layers, as it was applied for other nanocomposite membranes, Feng et al.229 and Zeng et al.230 used dopamine as a cross-linking agent. Compared to the previously reported total flux values for Ti3C2Tx/GO nanocomposite membranes having optimum performances, higher fluxes were reported as 174, 90, 89, 116, and 109 L/m2 × h × bar for MB, MR, MO (methyl orange), CR, and EB dye solutions with similar dye rejection performances.229 Collectively, by varying MXene and GO content simultaneously, optimal balance was caught for the MXene/GO nanocomposite membrane. The main reason for the better separation properties of MXene/GO nanocomposite membranes compared to pure MXene and pure GO membranes was related among the above studies to the nonselective regions existing in MXene and a critical role of GO in ensuring the selective filtration of MXene. Alternative to GO, as a carbon-based material, carbon nanotubes (CNTs) were proposed to fabricate nanocomposite membranes with MXene due to the possible size-sieving ability of CNTs.231,232 Ding et al.231 reported excellent water permeance of 1270 L/m2 × h × bar for the Ti3C2Tx/CNT nanocomposite membrane and related it to the unique fusiform structures between CNT bundles and MXene nanochannels, leading to the fast water transport. This structural organization was not surprising and observed also with the insertion of COF233 and Bi2S3234 nanoparticles in MXene nanochannels. However, Sun et al.232 displayed that with thermal cross-linking of CNT and MXene nanomaterials, interlayer spacing within the membrane decreased, yielding in the shrinkage of channels and drop in water flux, which was in the order of the performance of MXene/GO nanocomposite membranes.

In addition to pressure-assisted permeation of water and dye separation in pristine MXene nanofiltration membranes, electric field-assisted permeation was also studied via the application of negative and positive voltages. It was also suggested that on top of the size-sieving ability of MXene membranes, their voltage-gated sieving ability could be used for dye separation benefiting from the electrostatic interactions between dyes and MXene, which are regulated by changing the applied external voltage.235,236 Therefore, it reveals that the separation performance of MXene membranes can be enhanced via mean contribution of different driving forces.

According to the above-mentioned studies, the dye separation performances of MXene membranes are accelerated by the modification of either the surface or interlayer distance of MXene via intercalation, functionalization, or composite and nanocomposite membrane fabrication. However, some basic analysis such as structural or methodological analysis in the synthesis of MXene should also be examined in detail, probably to enhance membrane performance. Although there are some studies revealing the effect of synthesis method, lateral size, or structural deformations of MXene on its dye separation performance, to gain more understanding, different aspects should be evaluated. For instance, Kim et al.237 fabricated pristine MXene membranes via both vacuum-filtration and slot-die coating methods. They reported three-times greater pure water permeance as 190 L/m2 × h × bar with the dye rejection rate greater than 90% for several dyes for the membrane fabricated by the slot-die coating method, compared to the one synthesized with the highly preferred and simple vacuum filtration method (66 L/m2 × h × bar and 77.9–97.9%).237 On the other hand, regarding the structural properties, Xing et al.238 identified that without following any modification methods discussed above, just fabrication crumbled MXene membranes improved the dye separation capacity of MXene compared to the flat MXene membranes. Crumbled MXene membranes fabricated via the freeze-drying method displayed a two orders of magnitude greater solvent (water, acetone, methanol, ethanol, 1-propanol, and n-butanol) permeation and with a little compromise from the dye rejection at high MXene loading.238 A similar observation was proposed by the study of Li et al.239 where pores or, in other words, new pathways [see Figure 5(a)] for the transport of water were introduced via the chemical etching of MXene with hydrogen peroxide (H2O2). Xiang et al.240 examined the effect of lateral-size of MXene nanosheets on dye separation. Since the MXene membrane having small lateral size offered a short pathway [see Figure 5(b)] for water transport, they reported better water permeance and comparable dye rejection compared to the MXene having large lateral size. Feng et al.241 combined these two strategies [Figure 5(c)] by introducing porous material as COF and fabricating a MXene/COF membrane top to each nanomaterial having different lateral sizes. Collectively, handling the pathway of molecules for their transport within MXene membranes is an efficient strategy to improve dye removal capacity of MXene membranes without following any modification approaches.

Figure 5.

Figure 5

Schematic illustration of the possible pathways for water in (a) porous MXene, Adapted with permission from ref (239). Copyright 2021 Elsevier, (b) MXene having different lateral sizes, Adapted with permission from ref (240). Copyright 2022 Elsevier, and (c) MXene having different lateral size combined with porous COF. Adapted with permission from ref (241). Copyright 2022 Elsevier.

MB (3,7-bis(dimethylamino)-phenothiazin-5-iumchloride) is a cationic dye widely used for biological staining, dyeing of paper, wool, cotton, tannin, clothes, and for the treatment of methemoglobinemia and urinary tract infections. However, above a certain concentration causes carcinogenicity and health problems such as breathing, vomiting, eye burns, etc. Therefore, it is necessary to develop effective strategies for the separation of MB from wastewater. Several studies have been carried out to remove MB effectively using MXene OSN membranes.193,194,211,214,215,227,242 Since we aimed to display the limits of MXene membranes for each separation application, we plot MB separation performances of several MXene membranes to draw a frame for a specific particle in Figure 6. In accordance with the collected data from the literature, a big portion of data dominated mainly on the region of rejection rate greater than 99%. This evidently reveals the applicability of MXene for the dye separation from wastewater, despite the varying of total water permeances of MXene membranes in the vicinity of 50 L/m2 × h × bar. Even so, fortunately, thanks to unique MXene modification approaches, there are some promising MXene membranes having the water permeance in the order of hundreds with greater MB rejection than 95%.

Figure 6.

Figure 6

Comparison of methylene blue separation performance of pristine MXene (empty symbols) and the MXene nanocomposite (full/partially full symbols) membranes measured by several groups (Kang et al.,193 Wei et al.,194 Liu et al.,227 Feng et al.,229 Zhang et al.,211 Long et al.,215 Tong et al.,219 Yi et al.,224 Gong et al.,233 Kim et al.,237 He et al.,214 and Tao et al.217). Arrows represent the alteration in MXene content in a nanocomposite membrane. LMH: L/m2 × h.

Not only is superior dye separation performance of MXene-based membranes necessary but also is their long-term stability vital for the long-term use of membranes. Since long-term stability is directly related to the applicability of membranes in industrial processes, several studies concentrated on this issue for MXene membranes.211,220,222,229,237 Feng et al.229 reported a 17% decrease in total flux of the Ti3C2Tx/GO nanocomposite membrane cross-linked with dopamine after six cycles with a stable MB rejection rate and flux recovery rate greater than 97.7% at each cycle. Similarly, Pandey et al.220 revealed the excellent flux recovery rate of 98% for thin film membranes prepared with MXene and CA during third cycle filtration of RhB solution with only a 3.5% flux decline. Stability in water permeance and slight drop in MB rejection of Ti3C2Tx intercalated with cucurbit[5]uril were reported considering great number of filtration cycles, as 30.243 Long-term stability of MMM fabricated by Han et al.222 was investigated under harsh conditions and comparably long operation times. MMMs immersed in different solvents and then GV separation performances were measured for 18 days. Dye rejection rates of each membrane soaked in each solvent kept stable. However, approximate flux decline ranging between 50 and 60% was observed after immersion.222 It was attributed to the membrane fouling by dyes, which garners attention to the importance of fouling behavior of membranes. Rather than composite or nanocomposite MXene membranes, pristine MXene membrane fabricated via the slot-die-coated method displayed high stability for 30 days without extra post-treatment,237 although it is well-known that MXene can be easily oxidized in aqueous solution to form TiO2.

One of the major limitations, which restrains the membrane applicability is the fouling described as the filtration performance decline over time. The pores of a membrane are blocked by particles and these particles start to accumulate on the surface during nanofiltration. To overcome this challenge, Pandey et al.213 used a different content of silver nanoparticles (0, 7, 14, 21, 28, and 35 wt %) and produced nanocomposite membranes. Flux recovery rates of MXene loaded with 21 wt % of Ag and pristine MXene were reported as 97 and 86% for MG, respectively. This superior performance was ascribed to the ability of Ag to improve the resistance of membrane towards organic foulants via its hydrophilic nature. Most importantly, this nanocomposite membrane exhibited high antimicrobial activity with a bacteria growth inhibition as >99%, whereas that of pristine MXene was ∼60% which is still high and that cannot be underestimated. This was the first study that indicated the importance of designing antifouling and antibiofouling membrane concurrently. In another study of the same group,220 as we mentioned above, they revealed excellent flux decline resistance to RhB with the composite membrane fabricated via the phase inversion method using MXene and CA. Additionally, they proposed that with the increase of MXene content, growth inhibition for two different bacteria (E. coli and B. subtilis) increased and reached to 98 and 96% for the membrane having a 10 wt % of MXene content, revealing a high antibiofouling performance of MXene. Flux recovery ratio is the other parameter used to define the fouling behavior of a membrane. Kallem et al.244 reported a considerably high flux recovery rate of the solution including humic acid (HA), SA, and BSA (94–97%) for the membrane fabricated with the grafting MXene via zwitterions and then mixing with PES. However, these values ranged around 60% for the pristine PES membrane and 80% for the unmodified MXene/PES nanocomposite membrane.244 Mainly, the modification to repel the adsorption of foulants on the membrane surface via electrostatic repulsion forces was suggested in the literature.213,244 With these exciting studies, the requirement to produce an antifouling membrane has been underlined significantly again.

4.2. Adsorption-Based Separation

An alternative to membrane technology, MXene nanomaterials were also used as adsorbents to remove dye from wastewater. The adsorption process is a highly efficient separation technique with its initial cost, simplicity of design, and ease of operation. Since the main target is to compose an adsorbent with high adsorption performance for dyes, MXene nanomaterials were aimed to design with a large interlayer spacing in order to use their full adsorption capacity by either functionalization via various groups242,245248 or physical treatment to intercalate various particles.249251 Jun et al.252 compared the removal performance of unmodified commercial MXene with Al-based (A100) MOF for dyes of MB (cationic) and AB (anionic). Although MXene and MOF had different surface areas as 9 and 630 m2/g, respectively, they revealed opposite removal rates for each dye. More specifically, removal rate of MB was achieved as 80% (140 mg/g) and 40% by MXene and MOF, respectively. On the other hand, the removal rate of AB was above 80% (200 mg/g) for MOF, whereas no AB was removed by MXene. Therefore, selective separation can be achieved with MXene adsorbents. To design efficient MXene as an adsorbent for the removal of dyes, the widely preferred strategy in the literature was to functionalize the MXene surface via either short or bulky groups. Jun et al.245 performed chemical modification by using ultrasonication to observe oxygenated functional groups on the surface of MXene and evaluated its adsorption performance for MB and MO removal. Compared to the pristine MXene (90 mg/g at pH 7 and 20 °C), MB adsorption capacities of MXene treated with two different frequencies (28 kHz and 580 kHz) were increased as 30.8 and 18.2%. Similarly, to increase the oxygenated functional groups on the surface of MXene, Li et al.253 observed peroxo-functionalized Ti3C2Tx with the exposure of H2O2, which revealed excellent dye adsorption performances of 558.0, 524.6, 292.6, and 258.2 mg/g for MB, RhB, MO, and CR, respectively. On the other hand, Cai et al.242 observed greater enhancement with the functionalization of MXene using phytic acid (PhA) by the hydrothermal process under different operation times (0.5, 3, 6, 12, and 30 h). Although pristine MXene had a very low adsorption performance for MB and RhB, PA-functionalized MXene for 12 h revealed 54 and an 80% improvement in dye adsorption performance and reached to 42.51 and 22.79 mg/g, respectively, at 25 °C and pH 7. The same strategy was followed by the study of Lei et al.247 where arenediazonium salts were used to intercalate and functionalize MXene via −SO3H groups. Since d-spacing of MXene increased from 9.3 to 14.3 Å after the intercalation of aromatic compounds within nanosheets, MB adsorption capacity of functionalized MXene enhanced 81% from 21 to 111 mg/g. It can be deduced that the attachment of comparably similar size of functional groups has led to almost the same adsorption capacity improvement regardless of dye type and initial performance of MXene. This information was also verified by Li et al.246 who modified MXene by −COOH groups and more bulky groups as polyethylene polyimide and then poly(acrylic acid) (PAA) using a layer-by-layer assembly method. While −COOH modified MXene revealed similar improvement as 81 and 89% in the adsorption capacity for MB and safranine T (ST), respectively, direct functionalization with a bulkier group did not lead to further significant enhancement. However, Hao et al.253 revealed that before functionalization with acrylic acid, treatment with an alkaline solution suggested a tremendous enhancement in dye adsorption performance as 3- and 13-times greater than pristine MXene for MB and CR dyes, respectively. Collectively, it can be suggested that to improve the surface adsorption capacity of MXene, functional groups should be interpenetrated through MXene channels causing the enlargement of interlayer spacing between nanosheets and serving as favorable adsorption sites. Therefore, a moderate size of functional groups is more efficient compared to a bulky one. Additionally, the stability of this enlarged interlayer spacing during the adsorption process was the other concern. Therefore, Vakili et al.248 proposed a pillaring approach to cross-link two MXene surfaces using terephthalate and reported excellent increase in the surface area from 5.4 to 135.7 m2/g and MB uptake of 209.5 mg/g.

The second way proposed to increase dye adsorption capacity of MXene in the literature was the intercalation of MXene via several molecules or nanoparticles by physical treatment. Wei et al.249 treated MXene (Ti3C2Tx) with different alkaline solutions including LiOH, NaOH, and KOH to benefit from the ionic radius of alkaline ions for the enlargement of interlayer spacing. MB adsorption capacities of NaOH, LiOH, and KOH treated and pristine MXene were reported as 189, 121, 77, and 100 mg/g, yielding change of 47, 17, and −30%, respectively.249 Alternatively, Wang et al.254 treated Ti3C2Tx with DMSO and NaOH solutions. Specifically, after DMSO treatment, an extreme increase in MB uptake was observed from 8 to 125 mg/g. Several researchers250,251,255 synthesized the MXene (Ti3C2Tx)/Fe3O4 nanocomposite material using Fe3O4 nanoparticles via the in situ growth method and investigated for MB adsorption performance. Zhu et al.250 measured the maximum MB removal performance of nanocomposite adsorbent as 11.68 mg/g at 55 °C while 2.10 mg/g at 25 °C. The same performance was also observed by the nanocomposite adsorbent fabricated by Zhang et al.251 as 3.8 and 11.68 mg/g at 25 and 55 °C, respectively. More importantly, Zhang et al.251 investigated the dye removal from binary dye solutions including separately MO and RhB along with MB. Removal efficiencies of MXene/Fe3O4 for MB from both binary solutions were 94%, whereas those for MO and RhB were reported as 17% and 5%, respectively, proving the selective adsorption performance of the nanocomposite towards MB. Alternative to the Fe3O4 nanoparticle, Luo et al.256 coordinated cubelike Co3O4 nanoparticles by the solvothermal method and observed equilibrium adsorption capacities of MB and RhB as 128.9 and 47.1 mg/g, respectively. Moreover, Gu et al.257 proposed a unique adsorbent having one order of magnitude greater MB adsorption capacity by decorating tiny ZIF-8 nanoparticles within the interlayer of Ti3C2Tx.

Since 2011, researchers have focused dominantly on one common MXene type as Ti3C2Tx. However, other members of the MXene family are waiting for their real performances to be discovered for membrane- or adsorption-based separation processes. Only Nb2CTx216 was tested for membrane-based separation, whereas there are few studies examining the dye adsorption performance of other MXene types such as Nb2CTx258,259 and V2CTx.260 Different strategies to modulate the interlayer distance between MXene nanosheets have been discussed up to now. However, the examination of the effect of the etching method in the MXene synthesis on dye separation performance was a weighty matter, which may accelerate the use of a safe and environmentally friendly synthesis approach. Peng et al.258 fabricated common MXene (Ti3C2Tx) and Nb2CTx by a traditional etching method using HF (t-Ti3C2Tx and t-Nb2CTx) and by a hydrothermal etching method using a mixture solution consisting of NaBF4 and HCl (h-Ti3C2Tx and h-Nb2CTx). SSA of h-Ti3C2Tx and t-Ti3C2Tx were 44.6 and 8.9 m2/g. The MB concentration in solution dropped to 76.4 and 82.7% for h-Ti3C2 and t-Ti3C2, respectively, whereas it reduced to 86.2 and 92.4% for h-Nb2CTx and t-Nb2CTx, proving the efficiency of the hydrothermal method. Although Peng et al.258 observed low MB adsorption capacity for Nb2CTx by either traditional or hydrothermal treatment methods, Yan et al.261 reported greater MB and MO adsorption capacities as 99.6 and 100.7 mg/g. Similar performance for MB was also suggested for V2CTx by Lei et al.260 as 111.1 mg/g at 25 °C.

MXene nanomaterials were also fabricated in the form of aerogel to further increase their dye adsorption performances.259,262 Ti3C2Tx at different loadings was immobilized by SA to form aerogel in order to remove MB.259 Comparable adsorption performance was reported for MB as 92.17 mg/g at pH 7 and 25 °C.259 However, Feng et al.262 reported an excellent CR adsorption capacity of 3568 mg/g for the mixture of Ti3C2Tx and PEI, immobilized by SA. It is worth noting that there is not any evidence ascribed to the observed improvement in dye adsorption performance after MXene incorporation into an aerogel during its formation. Therefore, it is not easy to provide the effect of MXene after immobilization on dye removal.

Prominent process design was proposed by utilizing from the potential separation performance of both membrane and adsorbent technology from Kim et al.263 For this novel process design, they created a hybrid system by combining Ti3C2Tx as an adsorbent with an ultrafiltration membrane (UF) to evaluate its removal performance for MB and MO. In the presence of MB and MO, normalized fluxes of MXene-UF were reported as 0.90 and 0.92, whereas those of single UF were 0.86 and 0.90, respectively. Likewise, when the synthetic dye wastewater was applied including MB (2 mg/L), humic acid, and salts, MXene-UF exhibited a high retention rate of 99.1% with a normalized flux of 0.90. Kim et al.263 not only suggested an alternative hybrid model for the separation of dyes but also made it to function as a bridge in transition to membrane technology.

Layered architecture of MXene not only provided shortcuts for enhanced water flux but also improved the removal of dyes thanks to their large amounts of nanochannels with tailorable interlayer distance via intercalation of nanoparticles. Surprisingly, early studies revealed superior permeances for pristine MXene membranes, whereas it could be reached by the following studies later. Nevertheless, the intercalation of some nanoparticles within the MXene nanochannels led to the considerable improvement in water permeances outperforming some commercial and novel nanocomposite membranes.211,213 Functionalization was proposed as another strategy not only to adjust the interlayer distance between MXene nanochannels but also to ensure that MXene has a dye-selective ability. However, the real improvement in water permeances was achieved by the synergy between novel polymers and MXene, boosting up the hydrophilicity of the membrane. On the other hand, GO is the nanomaterial that excessively tested with a combination of MXene as a nanocomposite membrane for the removal of dyes to catch the optimum separation performance. The main drawback in dye removal via the membrane process is the fouling phenomena, which was not overcome by MXene membranes despite the exceptions220,237 and requires more in-depth studies. Similar strategies were also followed for MXene-based adsorbents, in addition to handling different types of MXene such as Nb2CTx258,259 or V2CTx260 as adsorbents for dye removal. Collectively, although of all separation applications the most publications have been published on identifying the performance of MXene for dye removal, still there have been some aspects that need to be investigated in detail due to the large differences between dye separation performance in reported studies.

5. Separation of Oil-in-Water Emulsions

Unfortunately, industrial wastewater discharges include different types of contaminants that can pose significant long-term risks to human health and environmental safety and need to be removed before discharge. The second-ranking contaminant after dyes, which is produced in huge amounts by many industries, such as textiles, pharmaceuticals, petrochemicals, and metal/steel industries daily, is the oil–water mixtures. Depending on the region, oily wastewater effluent discharge limit is varied within 5–100 mg/L range.269 Therefore, finding effective separation or demulsification methods for oily wastewater has gained importance. Alternative to the common separation processes, membrane filtration is proposed for the removal of oil from the aqueous phase. Not only in membrane nanofiltration229,270273 which is considered as the heart of the membrane-based separation processes but also in membrane distillation130 and adsorption processes,273 the enhancement in separation of oil–water mixtures with the use of MXene nanosheets were proposed. In addition, to enhance the separation performance in these processes, MXene nanosheets were also used to solve the problem related to the concentration polarization, polarized layer, and especially membrane fouling. Since oily wastewater contains highly toxic substances, hydrocarbon compounds, heavy metals, and suspended solid particles, to keep flux rates at a high level throughout the separation process is the main problem in membrane-based oil-in-water-emulsion separation. For instance, Saththasivam et al.270 proposed an antifouling membrane, which was fabricated by coating Ti3C2Tx nanosheets on conventional print paper. They suggested that favorable characteristic properties of MXene will help to mitigate the fouling by the help of the hydrophilic nature of MXene which increases the interaction with water, leading to the formation of a water layer on the membrane surface and hence the decrease in interaction of membrane with oil.270 Their strategy was also confirmed by Tan et al.130 who used Ti3C2Tx as a coating material to fabricate a membrane via vacuum filtration onto a PVDF support for test in the direct contact membrane distillation (DCMD) process. They compared to the flux decline of PVDF and MXene-coated PVDF membranes after 21 h continuous filtration process using BSA (0.2 g/L) and NaCl (10 g/L) and reported that percentage of flux declines were 18.8 and 8.3% for PVDF and MXene-coated PVDF membranes, respectively.130 These preliminary results evidently prove the efficiency of MXene in mitigating the membrane fouling in oil-in-water-emulsion separation. The other purpose of the use of MXene was to benefit from its photothermal property. Tan et al.130 also used MXene-coated PVDF membrane to reduce the energy requirement for the DCMD process. While temperature increase for the PVDF membrane was observed approximately 6 °C, it was reported as 49 °C for PVDF coated with MXene. In view of this, Ti3C2Tx was suggested as promising nanomaterials to decrease the fouling effect and the energy requirement in addition to gaining an enhancement in separation performance.

The vacuum assisted self-assembly process is the mainly preferred method to prepare free-standing MXene membranes on different polymeric membrane supports. Membrane prepared with this method is widely studied in the oil-in-water-emulsion separation. Additionally, the mostly used MXene type was Ti3C2Tx for the fabrication of free-standing membranes. Saththasivam et al.270 reported average total fluxes for Ti3C2Tx as 544, 682, 649, 638, and 574 L/m2 × h × bar for oil-in-water emulsions including sunflower oil, hexane, petroleum ether, silicone oil, and diesel, respectively, with the oil concentration below 12 mg/L. More surprisingly, Li et al.273 observed much greater total flux around 6000 L/m2 × h × bar without any trace of oil in the permeate side for ultrathin membrane (about 0.03 μm) fabricated by a different kind of MXene (Ti2CTx) on the PES. However, the total flux decreased to 540, 488, and 437 L/m2 × h × bar for oil-in-water emulsions including toluene, soybean oil, and pump oil, respectively, with the oil level lower than 10 mg/L, when the tween-80 was used to stabilize the emulsions.273 Similar high total fluxes in four kind of oil-in-water emulsions such as toluene, petroleum ether, kerosene, and n-hexane were observed for the cracked-earth-like Ti3C2Tx membrane surface-enriched with −(OH)2 groups.274 Much more extreme dye separation performance was achieved for the membrane consisting of free-standing MXene fabricated in the form of nanoribbons by the treatment of KOH.275 Its total fluxes of oil-in-water emulsions for gasoline, n-hexane, diesel and edible oil were reported around 15,000 L/m2 × h × bar for each emulsion.275 Disappointingly, its fluxes for each oil-in-water emulsions decreased dramatically after the second cycle.

To reveal the outstanding performance of MXene-based membranes, the more challenging conditions were examined like oil/salt-water emulsion systems. Earlier studies have indicated that the MXene membrane separation performance for oil was deteriorated by the presence of both surfactant and salt in oil–water solution.273,276,277 Therefore, in order to fabricate MXene-based membranes preserving its performance in those harsh conditions, different strategies were applied and finally encouraging performances were revealed for MXene membranes. For instance, the effect of different corrosive conditions on the oil–water separation performance of MXene-based membranes was tested in the study of Zhang et al.271 They prepared a free-standing MXene (Ti3C2Tx) membrane on PVDF via cross-linking with SA to identify the separation performance of crude oil–water in acidic (HCl-3M), alkaline (NaOH-3M), and salty (NaCl-3.5 wt %) environments. Its total fluxes increased from 887 to 969, 1043, and 906 L/m2 × h × bar at each challenging conditions, respectively, with the oil removal efficiency greater than 99.4%.271 Similar analysis was performed for the cracked-earthlike, free-standing Ti3C2Tx membrane surface-enriched with −(OH)2 groups.274 However, total flux was altered from 4720 to 6957, 2796, and 2482 L/m2 × h × bar for 1 M NaOH, 1 M NaCl, and 1 M HCl emulsion, respectively.274 Although cross-linked MXene271 yielded in flux improvement in each emulsions, uncross-linked MXenes273,274 revealed a decrease depending on the oil-in-water emulsion system. In order to compare the uncross-linked and cross-linked MXene, Liu et al.278 modified Ti3C2Tx with 3-aminopropyltriethoxysilane (APTES) and various tannic acid (TA) concentrations for the separation of petroleum ether, lubricating oil, and vegetable oil emulsions. Without sacrificing from the rejection performance (≥98%), more than one order of magnitude enhancement was observed for each emulsion system after cross-linking.278 Therefore, we can conclude that the stabilization of MXene layer on the support membrane via cross-linking enhanced the total flux without changing the oil separation efficiency at harsh conditions.

This cross-linking methodology was also used to stabilize the nanocomposite membrane composed of the combination of MXene (Ti3C2Tx) and reduced graphene oxide (rGO) by Feng et al.279 PDA was used as a cross-linking agent to increase the adhesion not only between MXene and graphene but also between nanosheets and support membrane. They designed a MXene/rGO/PDA (160/40/100 mg) nanocomposite membrane to determine separation performance for dodecane, lubricating oil, and petroleum ether stabilized with a sodium dodecyl sulfate (SDS) surfactant. However, total fluxes were reported as 50.03, 56.34, and 72.31 L/m2 × h × bar for these oil-in-water emulsions, respectively.279 These values are one order of magnitude lower compared to the previous studies due to the formation of oil layer on the membrane surface related to the particle size of oil droplets. Since as a result of cross-linking, the membrane pore size was observed in the order of particle size of oil droplets, oil rejection performance was reported greater than 95% due to the blockage of membrane pores with oil droplets. Promisingly, better performance in terms of total flux was achieved in the combination of MXene with one-dimensional (1D) nanotubes, instead of 2D nanosheets.272,280 Total flux of lubricating oil-in-water emulsion reached to 4116 L/m2 × h × bar for MXene/halloysite nanotube/PDA (2/5/80 mg)272 and 1887 L/m2 × h × bar for MXene/aminated-carbon nanotube(ACNT)/APTES (2/10/10 mg)280 nanocomposite membranes, exhibiting excellent rejection rates (over 98%).

It is claimed in the literature that MXene nanocomposite membranes were effective for the separation oil-in-water emulsions and dyes simultaneously.279 Accordingly, a new question appeared for researchers: “Can we develop a highly effective nanocomposite membrane to separate multi-component pollutants-oil-in-water emulsion even under harsh conditions?” To find an answer to this question, He et al.214 synthesized a free-standing nanocomposite membrane composed of the combination of MXene (Ti3C2Tx) and UiO-66 without cross-linking. They proved that the MXene/MOF nanocomposite membrane can be preferred for the concurrent separation of oil and dye from aqueous solutions [see Figure 7(a)]. While total flux and rejection rates varied between ∼307 and 497 L/m2 × h × bar and ∼99.4 and 99.6% for various oil-in-water emulsions, respectively, these properties altered slightly between ∼328 and 488 L/m2 × h × bar for various oil-in MB water emulsions and MB rejection rates greater than 99%. Moreover, the MXene/MOF nanocomposite membrane displayed only a slight decrease in total flux (in the order of 7–10%) without a change to its rejection rates even under acidic (HCl-3M), alkaline (NaOH-1M), and salty (NaCl-saturated) environment. Thanks to these studies, which clearly demonstrate the potential of MXene-membranes, we can deduce the competence of MXene membranes in the separation of oil-in-water emulsions as well as dyes even under these very harsh environments, which may represent the actual wastewater conditions in nature. Considering separation properties of MXene-based membranes given in Table 3 for oil-in-water emulsions, we can conclude that MXene-based membranes provide an oil rejection rate greater than approximately 99.5% and total flux of around 500 L/m2 × h × bar or higher for specific oils.

Figure 7.

Figure 7

Oil-in-water emulsion separation performance of MXene membranes. (a) Fabrication and separation processes of MXene/UIO-66-(COOH)2 nanocomposite membrane for multicomponent pollutant–oil–in-water emulsion. Adapted with permission from ref (214). Copyright 2020 Elsevier. (b) (Left) Photograph and (Right) SEM image of MXene/polyimide hybrid aerogel. Reproduced with permission from ref (288). Copyright 2019 American Chemical Society. (c) Its absorption capacities for various organic liquids. Reproduced with permission from ref (288). Copyright 2019 American Chemical Society.

Table 3. Survey of Oil-in-Water Emulsion Separation Performance of MXene-Based Membranesa.

membranes oil-in-water emulsion rejection (%) total flux [L(kg*)/m2 × h × bar] ref
Ti3C2Tx on print paper (1.2 μm) sun flower oil >99 543.5 (270)
Ti3C2Tx on print paper (1.2 μm) hexane >99 682.3 (270)
Ti3C2Tx on print paper (1.2 μm) petroleum ether >99 648.7 (270)
Ti3C2Tx on print paper (1.2 μm) silicone oil >99 638 (270)
Ti3C2Tx on print paper (1.2 μm) diesel >99 573.7 (270)
Ti2CTx on PES (0.03 μm) toluene 6149 (273)
Ti3C2Tx on PES (0.03 μm) pump oil 6098 (273)
Ti3C2Tx on PES (0.03 μm) soybean oil 6115 (273)
Ti3C2Tx on PES (0.03 μm) toluene/Tween 80 99.94 540 (273)
Ti3C2Tx on PES (0.03 μm) pump oil/Tween 80 99.94 437 (273)
Ti3C2Tx on PES (0.03 μm) soybean oil/Tween 80 99.94 488 (273)
Ti3C2Tx on PES (0.03 μm) toluene/Tween 80/salt 99.98 505 (273)
Ti3C2Tx on PES (0.03 μm) pump oil/Tween 80/salt 99.98 392 (273)
Ti3C2Tx on PES (0.03 μm) soybean oil/Tween 80/salt 99.98 465 (273)
Ti3C2Tx on PVDF (18.15 μm) kerosene 99.8 444 (271)
Ti3C2Tx on PVDF (18.15 μm) crude oil 99.7 887 (271)
Ti3C2Tx on PVDF (18.15 μm) heptane 99.5 668 (271)
Ti3C2Tx on PVDF (18.15 μm) hexane 99.4 707 (271)
Ti3C2Tx on PVDF (18.15 μm) petroleum ether 99.6 762 (271)
Ti3C2Tx on PVDF (18.15 μm) crude oil in HCl (3 M) 99.76 970 (271)
Ti3C2Tx on PVDF (18.15 μm) crude oil in NaCl (3.5 wt %) 99.71 905 (271)
Ti3C2Tx on PVDF (18.15 μm) crude oil in NaOH (3 M) 99.72 1045 (271)
Ti3C2Tx/rGO with PDA on nylon lubricating oil/SDS 96 56.34 (279)
Ti3C2Tx/rGO with PDA on nylon dodecane/SDS 98 50.03 (279)
Ti3C2Tx/rGO with PDA on nylon petroleum ether/SDS 98 72.31 (279)
Ti3C2Tx/UiO-66-(COOH)2 on nylon-66 n-hexane 99.37 431 (214)
Ti3C2Tx/UiO-66-(COOH)2 on nylon-66 iso-octane 99.63 497 (214)
Ti3C2Tx/UiO-66-(COOH)2 on nylon-66 1,3,5-trimethylbenzene 99.46 344 (214)
Ti3C2Tx/UiO-66-(COOH)2 on nylon-66 toluene 99.55 457 (214)
Ti3C2Tx/UiO-66-(COOH)2 on nylon-66 hexadecane 99.37 422 (214)
Ti3C2Tx/UiO-66-(COOH)2 on nylon-66 crude oil 99.26 307 (214)
Ti3C2Tx/UiO-66-(COOH)2 on nylon-66 toluene in HCl (3 M) 99.33 398 (214)
Ti3C2Tx/UiO-66-(COOH)2 on nylon-66 toluene in NaOH (1 M) 99.46 457 (214)
Ti3C2Tx/UiO-66-(COOH)2 on nylon-66 toluene in NaCl (saturated) 99.52 439 (214)
Ti3C2Tx/UiO-66-(COOH)2 on nylon-66 n-hexane/MB 99.65 443 (214)
Ti3C2Tx/UiO-66-(COOH)2 on nylon-66 iso-octane/MB 99.35 488 (214)
Ti3C2Tx/UiO-66-(COOH)2 on nylon-66 1,3,5-trimethyltoluene/MB 99.35 328 (214)
Ti3C2Tx/UiO-66-(COOH)2 on nylon-66 toluene/MB 99.40 446 (214)
Ti3C2Tx/copolyamide (single layer) vegetable oil (10 mg/L) 97.44 11,000 (289)
Ti3C2Tx/copolyamide (single layer) vegetable oil (100 mg/L) 95.31 8000 (289)
Ti3C2Tx/copolyamide (single layer) vegetable oil (1000 mg/L) 97.89 4000 (289)
Ti3C2Tx/copolyamide (multilayer) vegetable oil (10 mg/L) 99.0 10,000 (289)
Ti3C2Tx/copolyamide (multilayer) vegetable oil (100 mg/L) 98.0 7000 (289)
Ti3C2Tx/copolyamide (multilayer) vegetable oil (1000 mg/L) 97.0 3500 (289)
Ti3C2Tx/TA/APTES on CA (50 μm) petroleum ether 98.5 4170 (278)
Ti3C2Tx/TA/APTES on CA (50 μm) lubricating oil 99.1 4108 (278)
Ti3C2Tx/TA/APTES on CA (50 μm) vegetable oil 99.8 3477 (278)
Ti3C2Tx/ACNT/APTES on CA lubricating oil 99.6 1887 (280)
Ti3C2Tx/ACNT/APTES on CA vegetable oil 99.8 1644 (280)
Ti3C2Tx/HAL/PDA on CA petroleum ether 99.9 4241 (272)
Ti3C2Tx/HAL/PDA on CA lubricating oil 99.8 4116 (272)
Ti3C2Tx (cracked-earthlike and −OH functionalized) toluene 6386 (274)
Ti3C2Tx (cracked-earthlike and −OH functionalized) n-hexane 4720 (274)
Ti3C2Tx (cracked-earthlike and −OH functionalized) petroleum ether 2365 (274)
Ti3C2Tx (cracked-earthlike and −OH functionalized) kerosene 464 (274)
Ti3C2Tx (expanded with PTFE) peanut oil 5327 (290)
Ti3C2Tx (expanded with PTFE) dichloromethane 7482 (290)
Ti3C2Tx (expanded with PTFE) paraffin oil 6577 (290)
Ti3C2Tx (expanded with PTFE) toluene 8114 (290)
Ti3C2Tx (expanded with PTFE) n-hexane 8351 (290)
Ti3C2Tx/ZnO/TA on PEN petroleum ether/SDS 99.61 2513 (286)
Ti3C2Tx/ZnO/TA on PEN isooctane/SDS 99.50 2427 (286)
Ti3C2Tx/ZnO/TA on PEN n-hexane/SDS 99.43 2412 (286)
Ti3C2Tx/ZnO/TA on PEN mesitylene/SDS 99.49 2317 (286)
Ti3C2Tx/ZnO/TA on PEN n-heptane/SDS 99.47 2196 (286)
Ti3C2Tx (coated with PDMS-PDA-PEI) on PVDF soybean oil 9.5–8* (291)
Ti3C2Tx (coated with PDMS-PDA-PEI) on PVDF soybean oil/SDS 12–6* (291)
Ti3C2Tx (nanoribbon) on MCE (38 μm) gasoline >99 ∼15,000 (275)
Ti3C2Tx (nanoribbon) on MCE (38 μm) n-hexane >99 ∼14,500 (275)
Ti3C2Tx (nanoribbon) on MCE (38 μm) diesel >99 ∼15,500 (275)
Ti3C2Tx (nanoribbon) on MCE (38 μm) edible oil >99 15,860 (275)
Ti3C2Tx /BN/PDA/PEI (5.68 μm) 1,2-dichloroethane 96.54 95.8 (287)
Ti3C2Tx /BN/PDA/PEI (5.68 μm) dichloromethane 98.83 326 (287)
Ti3C2Tx /BN/PDA/PEI (5.68 μm) hexane 94.90 875 (287)
Ti3C2Tx /BN/PDA/PEI (5.68 μm) chloroform 95.50 144 (287)
Ti3C2Tx /BN/PDA/PEI (5.68 μm) toluene 96.46 77.5 (287)
a

Membrane thickness is given in parentheses. ACNTs: amine functionalized carbon nanotubes, APTES: 3-aminopropyltriethoxysilane, BN: boron nitrate, CA: cellulose acetate, HAL: halloysite nanotube, MB: methylene blue, MCE: mixed cellulose, PDA: polydopamine, PDMS: polydimethylsiloxane, PEI: polyethylenimine, PEN: poly(arylene ether nitrile), PES: polyether sulfone, PTFE: polytetrafluoroethylene, PVDF: polyvinylidene difluoride, SA: sodium alginate, TA: tannic acid.

Challenges in oil-in-water emulsion separation such as to observe high flux and rejection rate, less fouling, and great chemical stability were targeted to be overcome by designing MXene-based nanocomposite membranes. However, the nanofiltration process is unfeasible in the case of big environmental pollution where large amounts of oil were spilled in lakes, rivers, or ocean. To separate huge volumes of oil spillage from water sources quickly and effectively, absorbents such as sponges and aerogels are considered a novel approach. There are few studies in the literature that used MXene-based adsorbents to evaluate their oil–water separation performance. Wang et al.281 used Ti3C2Tx to create the MXene/polyimide hybrid aerogel with very low density and high porosity [see Figure 7(b)]. Among the absorbents, aerogels such as polyimide aerogels are the most promising materials due to their excellent compressible features and thermal stability. Different organic liquids such as pump oil, chloroform, tetrahydrofuran, soybean oil, acetone, toluene, n-hexane, and pump oil were applied to determine the absorption capacity of the MXene/polyimide hybrid aerogel. The highest percentages of weight gain (absorption capacities) were reported as 5778 and 4508% for pump oil and chloroform, respectively, as given in Figure 7(c). Its performance was compared with the graphene/polyimide aerogel281 which absorbed motor oil 37 times its own weight, whereas the MXene/polyimide hybrid aerogel could absorb various organic liquids from ∼18 to ∼58 times its own weight.281 Thanks to the novel strategies to fabricate aerogel for oil-in-water emulsion separation, MXene was adapted successfully. For instance, melamine sponge (MS) covered with tetradecylamine (TDA)-functionalized MXene revealed absorption capacity ranges from 60 to 112 times of its own mass.282 Wood-inspired MXene ternary aerogels synthesized with a novel approach and composed of nanocrystal cellulose functionalized with the silane agent displayed absorption performance varying between 45 and 63 times its own weight for several oil-in-water emulsions.283 Due to the hybrid hydrophobic–hydrophilic surface characteristics of these novel aerogels, superior characteristic features in addition to separation performance were achieved. Moreover, since MXene nanomaterials can possess an extensive temperature limit, they have great potential for the applications in photothermal-assisted oil recovery.283285 Collectively, all these studies about MXene-based hybrid aerogels paved the way for the future absorbent studies.

Durability and long-term operation stability of the membrane are also crucial for the separation of the oily wastewater even under harsh conditions. Therefore, similar to other membrane-based separation applications, durability and stability of MXene membranes in oil-in-water emulsion separation were tested by all studies included. For instance, Saththasivam et al.270 reported a 13% decrease in total flux when sunflower oil was used in feed solution after 8 cycles of filtration with the stable oil rejection of >99% and did not observe any sign of degradation after operation/washing cycles. Similarly, the water flux of the MXene/ACNT/APTES nanocomposite membrane decreased 26% after the 8th cycle but still had high performance as 2284 L/m2 × h × bar.280 Emulsion separation flux of the MXene/ZnO nanocomposite membrane decreased 27% after the 10th cycle, revealing the high performance as 1838 L/m2 × h × bar and >99% petroleum ether rejection.286 However, fortunately, Li et al.273 did not achieve any decrease either in total flux or rejection rate even after 50 consecutive cycles for the Ti2CTx membrane. Similarly, no considerable changes were reported in the study of Zhang et al.271 where only kerosene/water emulsion was investigated for 10 cycles. Feng et al.279 proved the long-term stability and durability of the MXene membrane by washing with ethanol for 3 days and immersing in strong acidic or alkaline solutions for more than three months. Similarly, excellent durability after being immersed in water for 600 h was achieved for MXene/BN/PDA/PEI nanocomposite membrane fabricated by Zhang et al.287 Ten washing/operation cycles were also performed by He et al.214 for toluene-in-water emulsion, and a 2% increase in total flux and rejection rate drop from 99.6 to 99.4% were observed for the MXene/MOF membrane. Moreover, even after dipping into acidic, alkaline, and salty solutions for 8 h, the MXene/MOF nanocomposite membrane preserved its separation performance, confirming its excellent chemical stability under harsh conditions. A different stability and reusability issue was observed for the MXene/polyimide hybrid aerogel adsorbent proposed by Wang et al.281 Its absorption capacity for soybean oil after 10 cycles was kept nearly stable even under harsh conditions as treating with soybean oil at 400 °C in air for 1 h. However, when it was treated with soybean oil in liquid nitrogen for 10 min, its performance was deteriorated after the 5th cycle due to the occurrence of fractures during the freezing process leading to the increase in the permeation of oil molecules. Interestingly, while reusability of MXene/polyimide hybrid aerogel was possible at high temperatures, it was restricted at low temperatures due to the brittle structure of aerogel. However, absorption capacity of n-hexane and silicone oil decreased only as ∼8 and 10% for MXene/TDA/MS aerogel after 20 cycles.282 For super heavy crude oil, adsorption capacity of the ternary MXene aerogel dropped only 23.8% after 5 cycles.283 In accordance with these studies, long-term operation and chemical stability of MXene membranes and absorbents were only revealed for specific oil-in-water emulsions and for the limited number of cycles where the highest one was 50. To mesmerize the membrane market, many more cycles for MXene membranes should be tested for several oil-in-water emulsion systems.

Since the fouling phenomena is a vital factor in the oil-in-water separation membrane due to limiting its durability and long-term operation stability, MXene nanomaterials were offered as a crucial solution to alleviate the fouling problem, utilizing their hydrophilic nature. Great performance was reported for free-standing MXene membranes cross-linked with different agents,271,278 free-standing MXene-based nanocomposite membranes composed of carbon-based nanomaterials and stabilized with cross-linking,229,272,280 and MXene-based adsorbents in the form of sponge and aerogel.281283,288 However, there are a very limited number of studies investigating the oil-in-water separation performance of MXene. Instead, the ones published evidently proved its superior capability of oil-in-water separation even under the more challenging conditions like oil/salt-water emulsion systems as well as oil/dye-water emulsion systems at very harsh environments.

6. Heavy Metal Ion Removal

The other contaminant that leads to the environmental pollution and threatens all living organisms is the heavy metals. Especially, heavy metal wastes arisen from the developed industries in the 21st century such as mining, metallurgy, leather tanning, electronics, and chemicals pose a serious threat. Heavy metals are toxic for organisms as they cause the formation of free radicals. More importantly, heavy metals easily accumulate in a human body and are not biodegradable. To date, various removal methods for heavy metal ions are applied such as chemical precipitation, coagulation, photocatalytic degradation, solvent extraction, adsorption, and membrane filtration.

6.1. Adsorption-Based Separation

Within the heavy metal ion removal processes, adsorption is a highly used process. The MXene family has become one of the strongest competitors to the existing adsorbents due to their high surface area and tailorable surface chemistry. MXene nanomaterials are tested dominantly as adsorbents for the removal of several heavy metal ions such as Cd(II),254,292294 Cr(VI),143,262,294302 Cu(II),293,294,303307 Hg(II),292,308310 Ni(II),311 and Pb(II)306,307,312316 in the literature. Different mechanisms have been proposed to explain the advanced adsorption performance of MXene. The principal adsorption mechanism of the MXene nanomaterial for heavy metal ions was proposed as electrostatic attraction between adsorbent and ions.313 Oxidation of ions to their low oxidation states on the surface of MXene was proposed as a supporting mechanism for the enhancement of adsorption performance of MXene.143,300 Additionally, inner-sphere complexation which consists of pH-dependent electrostatic interaction is the other mechanism used to explain the superior performance of MXene.312,317 The final mechanism proposed for the identification of performance of MXene is the ion-exchange.312,313 However, mostly, the removal performance of MXene is explained based on the mutual effect of several mechanisms. The heavy metal ion adsorption mechanism was also aimed at being identified by first-principles calculations.318 Preliminary studies motivated the use of MXene nanomaterials as adsorbents for the removal of heavy metal ions by displaying their outstanding adsorption performance. For instance, Ti3C2Tx having SSA of 57 m2/g revealed the removal capacity of Cr(VI) as 250 mg/g.143 Its removal performance of bromate (BrO3) was suggested as 321.8 mg/g with instant 100% BrO3 removal from drinking water.317 Additionally, adsorption performance of MXene for several heavy metal ions was compared with the common adsorbents.295,312 Jun et al.312 compared the adsorption efficiency of Ti3C2Tx with the powder activated carbon (PAC) for Pb(II) ions. Removal rate of MXene was reported as ∼92% while it was ∼68% for PAC, although MXene had less surface area as ∼10 m2/g than PAC (∼470 m2/g).312 This was explained by the electrostatic interaction between the negatively charged MXene surface and Pb(II) ions along with the mechanisms of ion-exchange and inner-sphere complex formation.312 Karthikeyan et al.295 suggested that adsorption capacity of Ti3C2Tx for Cr(VI) (104 mg/g) was greater than the common adsorbents as ceramic materials. Additionally, composite adsorbents composed of different materials modified with MXene have led to an increase in removal performance. For instance, adsorption performance of alginate modified with PEI and −NH2 functionalized Ti3C2Tx (550.3 mg/g)262 for Cr(VI) was comparably higher than both pure alginate beads modified with PEI (375.3 mg/g)319 and alginate modified with PEI and Fe3O4 (175.8 mg/g).320 With these exciting studies, the MXene family started to be tested for adsorption of other heavy metal ions as listed at Table S5, and several strategies are proposed to improve their adsorption performance.

One of the strategies preferred to enhance the adsorption capability of MXene is the functionalization of the MXene surface, benefiting from its tailorable surface chemistry. Amino-functionalization of Ti3C2Tx led to a 44.2% increase in its adsorption capacities for total Cr.297 Functionalization with the silane coupling agent enhanced the 205% Pb(II) adsorption capacity of Ti3C2Tx.315 Likewise, insertion of the enzymatic hydrolysis lignin as a biosurfactant into the Ti2CTx yielded an increase of 50.2% in Pb(II) adsorption capacity.313 In another study, alkaline-treated Ti3C2Tx having intercalated with Na(I) ions displayed a superior Pb(II) adsorption capacity (∼140 mg/g) and surprisingly instant equilibration within only 2 min.314 Zhang et al.307 offered to combine the above purposed methods as alkaline treatment and amino-functionalization for the removal of Pb(II). As expected, the highest Pb(II) adsorption capacity (see Table S5) along with the lowest equilibrium time of 20 min was correlated with this synergetic modification compared to the performance of the above-mentioned MXene adsorbents modified with different methods. Comparably, modified Ti3C2Tx displayed a 69.2% improvement in the Pb(II) adsorption capacity accompanying the enhancements in interlayer spacing from 8.8 to 13.6 Å and surface area from 6.37 to 129.2 m2/g.307 Collectively, since the interlayer spacing of MXene increased after functionalization, it was claimed that this provided more active adsorption sites leading to the strong van der Waals forces and electrostatic interaction along with effective exchange of ions, as illustrated in Figure 8(a).307,313315

Figure 8.

Figure 8

Schematic illustration of the removal mechanisms of Ti3C2Tx for (a) Cr(VI) and (b) Ba(II) or Sr(II). Panel (a): Adapted with permission from ref (143). Copyright 2015 American Chemical Society. Panel (b): Adapted with permission from ref (321). Copyright 2020 Elsevier.

Fabrication of MXene-based composite adsorbents with different polymers was considered to be another effective strategy for removing heavy metal ions. Either amino acid-based or petroleum-based polymers are used to observe composite materials. To improve the efficiency of MXene for Cr(VI), the Ti3C2Tx/poly(m-phenylenediamine) (PmPD) composite was produced via functionalization of Ti3C2Tx through in situ polymerization and then intercalation of PmPD.296 Thanks to the synergistic effect between Ti3C2Tx and PmPD, the removal capacity of Cr(VI) was observed as 540.47 mg/g which was greater than the performance of pure PmPD of 384.73 mg/g and pristine Ti3C2Tx of 137.45 mg/g (40.5 and 293%, respectively).296 This was attributed to the increase in interlayer spacing from 14.6 to 17.6 Å and the specific surface area from 10.42 to 55.93 m2/g compared to the pristine MXene.296 Similarly, the MXene/polymer composite adsorbent was prepared via in situ growth of imidazoles on the surface of MXene, and its adsorption performance of Cr(VI) was examined.322 Although its adsorption performance for Cr(VI) was not greater than the result of the previous study,296 its adsorption capacity quickly increased to 66.91 mg/g within 1 min and reached to the equilibrium point of 119.5 mg/g at the end of the 80 min.322 On the other hand, amino acid polymerization, then coating Ti3C2Tx, and intercalation of amino acid into the Ti3C2Tx improved the Cu(II) adsorption capacity by 83.6%303 and 20.6%304 compared to the pristine MXene,303,305 respectively. Additionally, the nanocomposite fabricated using MXene and alginate consisting of several amino groups revealed the increased Cu(II) and Pb(II) adsorption rates of 92.1 and 114.3%, respectively, with the increase in alginate concentration from 30 to 70% in a composite adsorbent.306

Unprecedented Hg(II) removal performance was reported for MXene-based adsorbents compared to other heavy metal ions.292,308310 Hg(II) adsorption performance of pure SA was compared with GO/SA and Ti3C2Tx/SA nanocomposites by the group of Lee. Adsorption rates of GO/SA and pure SA for Hg(II) were 34.63 and 11.53%, respectively, whereas that of MXene/SA was 100% with the maximum adsorption capacity of 932.84 mg/g.292 In their following study, they observed an excellent adsorption capacity of 1128.41 mg/g with the deposition of Fe2O3 nanoparticles on the surface of MXene, instead of having lower SSA of 56.51 m2/g than the pristine MXene (63.39 m2/g).310 Hg(II) removal rates of pristine Ti3C2Tx and Ti3C2Tx/Fe2O3 nanocomposite were 58.92 and 99.27%. Similarly, molybdenum disulfide (MoS2) was deposited on the surface of MXene to further improve its Hg(II) adsorption capacity by the same group.309 With the fabrication of delaminated MXene/MoS2 nanocomposites, they observed the greatest adsorption capacity of 1435.2 mg/g with the removal efficiency of 98.5% for Hg(II).309 This exceptional Hg(II) removal of the nanocomposite was not attributed to an only distinct adsorption capacity but also to the catalytic reduction. Fu et al.308 unveiled the tremendously high Hg(II) maximum adsorption capacity (4806 mg/g) of oxygen-functionalized Ti3C2Tx and linked this performance with the ability of {001}-Ti edge and oxygen functional groups that enable catalytic reduction and electrostatic interaction, respectively. Similar Hg(II) removal performance was reported for Ti3C2Tx (5070 mg/g) by Shahzad et al.323 and compared with the maximum adsorption capacity of Ti3CNTx (4263 mg/g). The achievement of high Hg(II) adsorption in MXene-based nanocomposites encouraged researchers to combine MXene with different materials for the removal of other heavy metal ions.

Encouraging from the performance of MXene-based nanocomposites for Hg(II) removal, novel MXene-based nanocomposite designs were proposed as potential candidates for the removal of other heavy metal ions.143,298,299 Alkaline-treated Ti3C2 having intercalated nanoscale zerovalent iron (nZVI) within its nanochannels revealed a 537% improvement in Cr(VI) adsorption capacity from 30.6 to 194.87 mg/g.299 This high adsorption capacity was ascribed to the complex adsorption mechanism where negatively charged Cr(VI) ions adsorbed onto the MXene surface, Cr(VI) reduced to the Cr(III), and Fe-O-Cr(III) species were formed simultaneously.299 Alternatively, TiO2 nanoparticles were distributed regularly between MXene nanochannels for the removal of Cr(VI).143 As a result of strong electrostatic interaction and reduction of Cr2O72– to Cr (III), the removal rate of the Ti3C2/TiO2 nanocomposite for Cr(VI) was achieved as 97.7%.143 TiO2-C/TiC nanocomposite derived from Ti3C2(OH)0.8F1.2 revealed higher adsorption capacity of Cr(VI) as ∼225 mg/g with the removal rate of >95% compared to the precursor MXene (∼62 mg/g).298 Similarly, novel nanocomposite adsorbents were also proposed for Ni(II) removal. Feng et al.311 synthesized the Ti3C2Tx/layered double metal hydroxide (LDH) nanocomposite having maximum adsorption capacity of 222.7 mg/g with the removal rate of >97.35%. However, adsorption capacities of pristine LDH (38.95 mg/g) and MXene (52.86 mg/g) for Ni(II) were well-below the performance of the nanocomposite adsorbent, which was attributed to the presence of multimolecular layer adsorption. Since impressively high improvements were reached for heavy metal ion removal using novel MXene-based nanocomposite materials, this evidently highlights the importance of engineered structures like the layered 2D–2D heterogeneous nanoplatelets for the adsorbent materials.

The final identified strategy is the morphology of adsorbents. There are contradicting observations about the effect of morphology in the literature. For instance, Gu et al.316 suggested that Ti3C2Tx produced as nanofibers displayed the highest maximum Pb(II) adsorption capacity of 285.9 mg/g compared to the commercial MXene nanosheets. This was attributed to the enhanced surface area of 16.4 m2/g and plenty of oxygenated functional groups compared to its nanosheet form. However, Shahzad et al.292 reported that the Ti2CTx produced as nanosheets displayed higher maximum adsorption capacity of Cd(II) as 294.23 mg/g than its nanofiber form. A similar reason was claimed in this study to explain the high performance of the nanosheet form with a higher surface area of 66.7 m2/g compared to the nanofiber form (55.1 m2/g) of MXene. The main difference in the fabrication of MXene sheets/fibers from the corresponding MAX phase is that while the former played with the temperature by keeping molarity of NaOH constant in solution, the later changed the NaOH molar value at constant temperature to form either nanosheets or nanofibers. However, the Ti2AlC MAX phase was produced via a bottom-up approach by Shahzad et al.292 while Gu et al.316 used a top-down method for the fabrication of Ti3C2Tx nanofibers and nanosheets. In another study, by playing with the alkalization process, Ti3C2Tx was also fabricated in the form of nanofibers and nanoribbons using NaOH and KOH solutions, respectively.324 NaOH alkalization led to an urchinlike nanofiber morphology with a surface area of 62.4 m2/g, whereas KOH alkalization was formed as a kelplike nanoribbon with a surface area of 57.8 m2/g. Therefore, maximum adsorption capacity of the former for Pb(II) (328.9 mg/g) was greater than that of the latter (248.3 mg/g).324 Additionally, Dong et al.293 fabricated the precursor (Ti3AlC2, MAX phase) of MXene in a nanofiber form and tested its Cd(II) and Cu(II) adsorption capacities. Even if the heavy metal adsorption capacity of the precursor of MXene cannot reach the performance of MXene, the comparison of the nanofiber form with the original morphology of the MAX phase evidently revealed the effect of morphology on adsorption performance. While the commercial MAX phase had 4.28 and 6.57 mg/g adsorption capacity for Cu(II) and Cd(II), its nanofiber form revealed those of 11.13 and 11.50 mg/g, respectively.293 To sum up briefly, the morphology which increased the surface area and, hence, the number of active adsorption sites is more capable of improving the capacity of MXene for heavy metal ion removal.

Unfortunately, different heavy metal ions coexist in wastewater, which complicates the removal process. Heavy metal ion adsorption capacities of MXene nanomaterials were examined with the coexisting ions in the solution in order to reveal their combined effect on the removal of heavy metal ions. For instance, Cu(II) removal efficiency (∼98%) of delaminated Ti3C2Tx was not affected from the presence of common heavy metal ions of Pb(II), Cr(III), and Cd(II) in the wastewater.305 Likewise, 98% removal efficiency of Ti3C2Tx for Ba(II) ions was reported when other metal ions such as As(III), Pb(II), Cr(IV), Ca(II), and Sr(II) were present in the simulated solution.325 Similarly, Ti3C2Tx kept the removal rate of Hg(II) as nearly 100% with coexisting heavy metal ions such as Pb(II), Cd(II), Ni(II), and Cu(II), while its removal rates for other metal ions were only 47.8, 18.4, 9.8, and 7.6%, respectively, proving high selectivity of Hg(II).326 Slight drops for alkaline-treated and nZVI-intercalated Ti3C2Tx were reported where 54% removal efficiency of Cr(IV) was decreased only 2.23% for wastewater, including Ca(II), Mn(II), Zn(II), and Ni(II) coexisting ions.299 However, when single and mixture adsorption performances of Ti3C2Tx were compared in another study, 5.5, 38.0, 71.4, and 82.9% reductions in the removal efficiencies of Pb(II), Cu(II), Zn(II), and Cd(II) were reported, respectively, which follows the heavy metal hydroxide constants (log K values).299 On the contrary, Ti3C2Tx having large interlayer spacing and abundant adsorption sites due to the alkaline treatment revealed high reduction in Pb(II) removal efficiency as 37.1 and 33.9% in the presence of Ca(II) and Mg(II), respectively.314 Encouraging data obtained from the studies identifying the effect of heavy metal ion mixture may lead researchers to investigate the use of MXene in the industrial field.

The studies reviewed so far have revealed that MXene has high adsorption performance for heavy metal ions. However, demonstrating its industrial applicability is as important as adsorption capacity. In order to determine the real performance of MXene as an adsorbent in the real wastewater environment, a simulated water environment, which includes various competing ions, was applied to the MXene and MXene-based nanocomposite adsorbents in several studies. Gu et al.316 treated the layered Ti3C2Tx nanofiber (single Pb(II) adsorption capacity of 285.9 mg/g) with the simulated drinking water consisting of Ca(II) (40 mg/L), Mg(II) (80 mg/L), Na(I) (100 mg/L), and Pb(II) (0.1 mg/L). Average treatment capacity was 2500 kg water/kg nanofiber with a Pb(II) concentration of ∼0.008 mg/L, which is under the drinking water standard (0.01 mg/L) at pH 6.8–7.2. Similarly, Zou et al.298 treated the urchinlike TiO2-C/TiC nanocomposite (single Cr(VI) adsorption capacity of 225 mg/g) with synthetic wastewater including 2 mg/L of Cr(VI) and 200 mg/L of Cl(I), NO3(I), and SO4(II) at pH 5.5–6.8. They unveiled the treatment capacity of 1120 kg water/kg adsorbent for this nanocomposite with 0.1 mg/L of Cr(VI), which is the wastewater discharge standard in China.298 However, the treatment capacity was observed as nearly 800 kg water/kg adsorbent with 0.1 mg/L of Cr(VI) when real wastewater was used.298 Shahzad et al.292 prepared simulated groundwater containing Na(I) (0.1 mg/g), Mg(II) (0.08 mg/g), and Ca(II) (0.04 mg/g) ions to investigate adsorption capacity of the Ti2CTx nanofiber and nanosheet for Cd(II) removal. They reported excellent removal rates of 99.47 and 99.70% with the final Cd(II) concentration of 1.87 × 10–6 and 1.04 × 10–6 mg/g for the Ti2CTx nanofiber and nanosheet, respectively, which is in accordance with the US Environmental Protection Agency limit (EPA) (5 × 10–6 mg/g).292 Likewise, the Ti3C2Tx/MoS2 nanocomposite preserved its high Hg(II) removal efficiency of 99.75% in the presence of other metal [Pb(II), Cd(II), Cr(III), Cu(II), and Zn(II)] and salt [Mg(II), Na(I), Ca(II), and K(I)] ions in the simulated wastewater.309 The same Hg(II) removal performance was also reported for multilayered oxygen-functionalized Ti3C2Tx where Hg(II) removal efficiencies were varied between 99.6 and 97.5% in the presence of separate salt ions as Mg(II), Na(I), Ca(II), and K(I).308 However, a desperate situation is valid for the Ti3C2Tx/PEI/SA aerogel which has the highest Cr(VI) adsorption capacity within the identified MXene-based adsorbents in Table S5.262 Cr(VI) adsorption capacity decreased to 22.1 and 91.8% at low (0.001M) and high (0.1M) anion concentrations [Br(I), Cl(I), SO4(II), CO3(II)], respectively, whereas 14.8 and 41.0% drops were reported for two different cation concentrations [Mg(II), Na(I), Ca(II), and K(I)].262 Nevertheless, according to the above-discussed studies, considering the treatment capacity of water above 1000 kg with the removal efficiency of >99% for different heavy metal ions, MXene-based materials emerge as brilliant adsorbents for removing heavy metal ions from complex wastewater compositions.

The removal and recovery of precious metal ions from wastewater are necessary not only environmentally but also economically. Adsorption performance of MXene-based adsorbents and films for several precious metal ions such as Pd(II),300,327 Au(III),300,328,329 and Ag(I)300,329 were identified in the literature. For instance, the Ti3C2Tx/rGO nanocomposite was fabricated as a film, and its adsorption capacity of Au(III) reached to 1241 mg/g.300 Considerable alteration was not reported for Au(III) adsorption (1063.8 mg/g) in another study329 where Ti3C2Tx/rGO was fabricated in an aerogels form. However, the Au(III) adsorption capacity of the Ti3C2Tx/CNT nanocomposite film was reported as 2093 mg/g.328 The introduction of either rGO or CNT leads to the restacking of Ti3C2Tx and, hence, gaining high SSA. To explain the Au(III) adsorption process of the nanocomposite, the redox rejection mechanism was suggested where oxidation of MXene and reduction of Au(III) occurred at the same time.328 Additionally, while the Ti3C2Tx adsorbent treated at 45 °C revealed Pd(II) adsorption capacity of 185 mg/g,327 rGO intercalated Ti3C2Tx film displayed enhancement in the Pd(II) adsorption capacity (890 mg/g).300 Although this high Pd(II) adsorption performance was attributed to the rGO intercalation in order to hinder the restacking,300 it is worth noting that there is not any stage as exfoliation, which is performed in the film fabrication and increases the surface area of MXene, in adsorbent preparation aside from the etching stage.

In addition to the high removal performances of MXene for heavy metal ions, long-term applicability and reusability of MXene nanomaterials were mostly tested within several continuous cycles. Shahzad et al.309 unveiled the high stability of the Ti3C2Tx/MoS2 heterogeneous nanocomposite for Hg(II) adsorption. Nanocomposite treated with 5 M HCl for 3 h revealed the adsorption and desorption rates of Hg(II) as 100 and 37.3% after five consecutive cycles, respectively. However, as contact time increased to 30 h, the adsorption rate remained constant whereas the desorption rate reached to 55.0%. Likewise, alkaline-treated MXene/LDH nanocomposite fabricated by Feng et al.311 exhibited a superior Ni(II) adsorption rate of >85.32% even after eight continuous adsorption–desorption cycles at 25 °C. The maximum adsorption/desorption cycle number was carried out by Dong et al.306 where they tested Pb(II) and Cu(II) adsorption rates of cross-linked and uncross-linked Ti3C2Tx/alginate adsorbents within 10 continuous cycles. Following the cleaning process of Ti3C2Tx/alginate by 0.1 M nitric acid solution, only 8.9 and 5.4% drops were observed in the adsorption rates of the cross-linked adsorbent for Pb(II) and Cu(II), respectively, while those of the uncross-linked one were 13.1 and 12.7%. Since the structural stability is preserved with the cross-linking, this also leads to excellent regenerability, in other words, the stability in heavy metal ion removal performance. Considering pristine MXene, rather than the composite, Hg(II) removal performance of Ti3C2Tx exhibited only a 9% decline after 5 consecutive adsorption–desorption cycles,323 which is still promising. However, the removal of Hg(II) for the other type of MXene, Ti3CNTx, decreased almost 25% under the same conditions.323 Unfortunately, a desperate situation is present in the regenerability test of delaminated MXene treated with an acidic mixture including nitric acid and calcium nitrate for 5 h after centrifugation.305 The removal rate of Cu(II) was 80% in the first cycle; however, in the second and third cycles, the removal rate dropped to 47 and 30%, respectively, due to the formation of TiO2 and incomplete Cu(II) desorption after the regeneration process. They suggested that the ion-exchange reaction between Cu(II) and the surface functional groups of MXene occurred during the adsorption process. As a result of the ion-exchange mechanism which was generally named as the inner-sphere complex formation, the formation of CuO2 via the reduction of Cu(II) was observed, proving the existing challenge in desorption process. Similarly, the Cr(VI) removal rate of the Ti3C2Tx/δ-MnO2 nanocomposite adsorbent sharply changed from 86 to 45% after five consecutive cycles, since the adsorption is partly governed by the chemisorption process, which hinders the regeneration of adsorbent at mild conditions.330

6.2. Membrane-Based Separation

Although MXene-based nanomaterials were dominantly used as adsorbents for heavy metal ion removal, its usage is limited in the membrane technology. MXene/Fe3O4 nanocomposite membranes, which were fabricated using various amounts of Fe3O4 nanoparticles, displayed a high water flux of 125.1 L/m2 × h due to the enhanced interlayer spacing. Likewise, moderate rejection values were determined as 63.2, 64.1, and 70.2% for Cu(II), Cd(II), and Cr(VI), respectively.294 This was attributed to the accessible −(OH)2 groups on the surface of the nanocomposite membrane with the insertion of nanoparticles. Similarly, MXene/CNT nanocomposite membranes were identified for the separation of Au(III).328 Encouraging results were reported where water permeability under a pressure of 1.0 bar was 437.6 L/m2 × h × bar, and Au(III) rejection was 99.8%.328 The high rejection performance was attributed to the spontaneous electron transfer ability of the nanocomposite material, leading to the reduction of Au(III) to zerovalent Au nanoparticles and, hence, easy rejection of Au nanoparticles through the nanochannels of MXene due to the size-sieving. Collectively, the proper use of MXene as a membrane material alternative to the adsorbent material is not a negligible strategy in heavy metal ion removal and requires further investigation.

Due to the high surface area and designable surface chemistry, the removal efficiency of the MXene family for heavy metal ions of Cd(II), Cr(VI), Cu(II), Hg(II), Ni(II), and Pb(II) was dominantly investigated and adsorption mechanisms under this performance were aimed to be revealed in-depth. Instead of a unique separation mechanism, it was suggested that the removal performance of MXene can be explained by the mutual contribution of electrostatic interaction, oxidation of ions, inner-sphere complexation, and ion-exchange. Similarly, the interlayer distance was accepted as a key point for MXene nanomaterials and aimed to tune via functionalization of the MXene surface and fabrication of the MXene-based composite adsorbents with different polymers. The most encouraging conclusion was derived as the instant equilibration (within only 2 min) of the MXene adsorbent during adsorption of specific heavy metal ions, suggesting the opportunity to carry out a great number of adsorption–desorption cycles within a limited time in the adsorption process. The other conclusion is that most of the MXene-based adsorbents have an unprecedented Hg(II) removal performance compared to other heavy metal ions, which makes them unique material for Hg(II) separation. However, on the other hand, there are contradicting observations especially about the effect of morphology and inadequate research on its efficiency in the industrial field and reusability. Instead of proven track records on the application of the MXene adsorbent for several continuous cycles, 11 cycles that were examined as a maximum in the literature are not enough to prove its regenerability. Therefore, we believe that heavy metal ion removal using MXene adsorbent requires a special consideration instead of the presence of published overarching studies.

7. Removal of Radionuclides

Considering the use of nuclear energy to a large extent, significant amounts of radionuclides are produced and released into nature as a nuclear disposal. Since radionuclides cause several disorders in human health and damage the environment, nuclear waste treatment with effective methods has become an important issue. Since the MXene family provides wide resistance to severe radiation, they are proposed as an effective adsorbent for removing radionuclides such as Ba(II),321,325,331 Cs(I),332334 Eu(III),273,335 Re(VII),336 Sr(II),321 Th(IV),336 and U(VI)126,254,335,337,338 from nuclear waste as summarized in Table S6.

Uranium has both chemical and radiological toxicity, and its release to the environment is a serious concern. Therefore, several adsorbent materials including MXene have been widely explored for the removal of U(VI). The first study examining the removal of U(VI) from nuclear wastewater was carried out using the V2CTx type of MXene.126 Maximum adsorption capacity of multilayered V2CTx was measured as 174 mg/g. However, DFT calculation displayed that experimental measurements cannot reveal the actual performance of the adsorbent which was computed as 536 mg/g.126 Likewise, for the other type of MXene, Ti3C2(OH)2, DFT simulations also revealed high [UO2(H2O)5]2+ adsorption capacity of 595.3 mg/g.126 Therefore, further studies are aimed to improve the U(VI) adsorption performance of the MXene family either by intercalation, functionalization, or fabrication of composite adsorbents. For instance, it reached 214 mg/g by the intercalation of DMSO and NaOH254 and 344.8 mg/g by the functionalization with carboxyl groups for Ti3C2Tx.335 Promisingly, U(VI) adsorption capacity hit the peak for Ti2CTx as 470 mg/g, revealing the complete removal within 48 h.337 However, the functionalization with amidoxime resulted in a mild increase for a U(VI) uptake of 294 mg/g with a removal rate of 95% after 5 min.338 Perfect conductivity of MXene nanomaterials are provided as a benefit for the adsorption, and its electrochemical adsorption performance for U(VI) was tested under periodic alternating potential. Surprisingly, U(VI) adsorption capacity reached to 626 mg/g at pH 5 which was higher than the data observed from the conventional adsorption process and predicted from DFT simulations.338 This was the first study to obtain improved U(VI) adsorption using the electro-sorption process on functionalized MXene but hopefully will not be the last one. To further increase U(VI) adsorption capacity of MXene adsorbents, they were modified with several other materials. For instance, alkali-treated Ti3C2Tx was modified with nZVI, and resultant novel nanocomposite adsorbent unveiled the highest U(VI) adsorption capacity of 1246 mg/g.339 The great performance of Ti3C2Tx/nZVI was attributed to the simultaneous formation of reduction precipitates and their adsorption from aqueous solution. Therefore, mainly to explain the adsorption mechanism of U(VI) on MXene, either a stable inner-sphere complex formation335 or a strategy of simultaneous adsorption and reduction337,339 was suggested.

Nuclear waste includes a high amount of Cs(I), a fission product of uranium.340 Therefore, its efficient removal using MXene-based adsorbents was also identified in the literature. However, a higher adsorption capacity of Cs(I) could not be reached as it was observed for U(VI) in MXene adsorbents. The removal rate of Ti3C2Tx fabricated via an in situ HF method was observed as 62.7% with a maximum adsorption capacity of 25.4 mg/g for Cs(I) within 1 min.332 Alternatively, Jun et al.341 reported the adsorption capacity of 148 mg/g using the same type of MXene for Cs(I) derived from nonradioactive CsNO3. To further increase Cs(I) adsorption performance of Ti3C2Tx, it was fabricated as an aerogel sphere using sodium alginate.334 Encouraging results were observed as a maximum adsorption capacity of 315.91 mg/g for Cs(I).334 This was attributed to the mean contribution of trapping of Cs(I) in the aerogel sphere and ion-exchange reaction with either alginate or MXene.334

The final radionuclides for which the removal performance of MXene adsorbents was excessively investigated is Ba(II). While pristine Ti3C2Tx fabricated in different studies revealed almost the same Ba(II) adsorption capacity such as 9.3325 and 11.98331 mg/g, alkali treatment with 5% NaOH leads to three times the enhancement in the Ba(II) adsorption capacity of Ti3C2Tx.331 With the increase in adsorbent dose, maximum Ba(II) uptake reached to 175.1 mg/g.321 It was proposed that inner-sphere complexation and ion-exchange mechanisms were responsible for the high performance of MXene, as given in Figure 8(b).

The reusability and regeneration property of MXene adsorbents are also investigated in the removal of radionuclides. Amidoxime-functionalized Ti3C2Tx kept its removal performance of U(VI) as nearly 100%, even after five adsorption–desorption cycles in 0.1 M Na2CO3 aqueous solution.338 Similar high regeneration performance was also proposed for DMSO-hydrated Ti3C2Tx used in U(VI) removal.254 Only 13% reduction in U(VI) uptake capacity was observed even after one month of the testing period. Furthermore, MXene treated with 0.2 M HNO3 after U(VI) saturation revealed the desorption efficiency of 98.5%, proving its excellent regenerability.254 In another study,321 slight decrements in removal rates of Ba(II) and Sr(II) as ∼5% and ∼8% were observed for pristine Ti3C2Tx after four cycles, respectively. However, there are few exceptions like the MXene/Ag2Ox/PDA nanocomposite. Its recycle efficiency for I(I) decreased to nearly 50% of its initial I(I) removal capacity after the 5th adsorption cycle.342

The critical issue in adsorbents is that they should not easily release the adsorbed ions, especially the radioactive ones to the environment. Therefore, leaching tests are performed by immersing the adsorbent into different aqueous media. It was also tested for some radionuclides trapped in MXene. For instance, Mu et al.331 immersed alkaline-treated Ti3C2Tx into the aqueous solutions for 3, 7, and 10 days to test the Ba(II) leaching. Ba(II) ion concentration in aqueous solutions was reported under the detection limit indicating the presence of irreversible interaction between MXene and Ba(II) at the adsorption conditions. Another study was carried out by Wang et al.254 where the U(VI) leaching test was performed for DMSO-hydrated Ti3C2Tx under various temperatures (200, 400, 450, and 500 °C) and atmospheres (air and N2). The optimum radionuclide imprisonment temperature was suggested as 400 °C in air for U(VI). Additionally, a long-term test of U(VI) leaching revealed that the leaching ratio was below 3% at the end of the 10 days. This study evidently proves the applicability of MXene nanomaterials for the encapsulation process.

To further evaluate the efficiency of MXene adsorbents, treatment of simulated or artificial wastewater-containing varying radioactive amounts was studied. The target was to display the potential of MXenes for industrial applications. For instance, different U(VI)-containing simulated water was tested for DMSO-hydrated Ti3C2Tx, and almost 5000 kg of U(VI)-containing wastewater was treated using only 1 kg of adsorbent.254 After the treatment, it was proposed that the remaining U(VI) content in wastewater was under the standard of the World Health Organization (WHO) (0.015 mg/L). On the other hand, it was demonstrated that Ti3C2Tx had a superior potential to remove U(VI) with a removal rate of 95.7% in the presence of a high amount of several coexisting ions and under an anaerobic environment throughout many days.337 Removal efficiencies of the alkali-treated Ti3C2Tx/nZVI nanocomposite adsorbent were observed as 88.9, 95.1, and 69.5% under various environmental media such as a 1.0 mM NaHCO3 aqueous solution, a solution mimicking groundwater, and a 10 mg/L humic acid aqueous solution, respectively.339 U(VI) removal performance of amidoxime-functionalized Ti3C2Tx was tested under simulated groundwater including U(VI), Ca(NO3)2, CaBr2, MgSO4, Na2SO4, NaHCO3, KHCO3, and Na2CO3 and compared to that of organic polyamidoxime (PAO).338 The adsorption capacity of amidoxime-functionalized Ti3C2Tx for U(VI) was around 230 mg/g while this value was 164 mg/g for PAO for a one day continuous measurement test.338 Similarly, Mu et al.331 tested the Ba(II) adsorption capacity of alkaline-treated Ti3C2Tx using simulated nuclear wastewater containing coexisting ions. The removal rate of MXene for Ba(II) was proposed as greater than 99%. Collectively, these studies proved that MXene are the favorable adsorbents for capturing radionuclides from actual nuclear wastewater.

MXene nanomaterials were only used as adsorbents rather than membranes for the removal of radionuclides utilizing the inner-sphere complex and ion-exchange mechanisms. Radionuclides of U(VI), Cs(I), and Ba(II) were highly investigated by MXene adsorbents, while the greatest performance was found for U(VI) removal. The most important feature of MXene nanomaterials demonstrated in the literature is that they effectively sequester radionuclides by hindering leaching. However, more research on the capacity of MXene adsorbents for removing radionuclides should be carried out, especially considering their well-proven ability of capturing radionuclides from actual nuclear wastewater.

8. Desalination

In parallel with the increase in industrialization and urbanizing population, a rise in water use is expected soon. Still, approximately 40% of the population on a global scale is affected by water scarcity, and unfortunately, this rate is expected to increase to 60% in 2025.343 Although there is a plenty of water in the world, only 2.5% of this water is available for utilization. Collectively, the increase in water usage and decrease in freshwater sources make the water treatment process as a key process in overcoming these problems, especially in water-stressed countries. Thermal-based processes such as evaporation, solar distillation, gas hydrate formation, etc. are the ones that have been used widely to remove the dissolved mineral salts in seawater.

8.1. Membrane-Based Separation

Membrane-based processes such as reverse osmosis (RO), nanofiltration, and electrodialysis are becoming more promising technologies because of their lower energy consumption, environmental footprint, and higher capacity. Desalination capacity via membrane-based separation techniques was around 97.5 × 106 m3/day in 2015344 and, hopefully, was estimated to reach to 192 × 106 m3/day by 2050.345 Industry that uses membrane-based desalination processes is dominated by polymeric membranes. However, novel advanced membrane materials for the desalination are emerging rapidly. Considering properties such as hydrophilicity, low contact angle, and easily functionalization, MXene nanomaterials are the most suitable candidates as the membrane material for the desalination processes. Table 4 summarizes the water permeance and salt rejection performances of existing MXene membranes. Ren et al.125 was the first group that introduced the MXene 2D nanomaterial for the desalination membranes. They synthesized a 1.5 μm-thick Ti3C2Tx membrane by vacuum-assisted filtration and obtained a water flux of 37.4 L/m2 × h × bar, which exceeded the performance of the industrial polymeric membranes.125 This unexpected performance was explained with the high water content of MXene layers in the wet state creating a free path for water flux. From permeation rates of ions such as Na+, Li+, K+, Ca2+, Mg2+, and Al3+, those of Na+, Li+, and K+ increased with the time due to the repulsive interaction between ions and MXene nanosheets arising from the formation of an electric double layer on the surface of the nanosheet due to the accumulation of salt ions. However, permeation rates of Ca2+, Mg2+, and Al3+ decreased due to attractive interactions between ions and highly negatively charged MXene nanosheets leading to their shrinkage. Simply, Ren et al.125 revealed the effect of radii and charge of ions on the water permeation for the first time where the existence of the ion-sieving mechanism for MXene membranes was displayed. Berdiyorov et al.346 carried out DFT calculations to provide an understanding about the size-charge selective ion sieving mechanism of MXene (Ti3C2) proposed by Ren et al.125 They suggested that this mechanism was interrelated with the electrostatic interactions between ions and the MXene surface, so the selectivity of ions originated from the surface charge network of MXene, which leads to a dynamic response to the penetrating ions by expanding or shrinking the interlayer spacing between the MXene nanosheets. Moving on the observations of two promising studies, a big door has been opened for researchers to create high separation performance membranes using MXene 2D nanomaterials. Generally, for the fast, reliable, and cost-effective observation, atomic-scale simulation approaches have been carried out to identify the effect of slit size and functional group of MXene, and the electric field was applied to define the desalination performance of MXene and, hence, to guide the experimentalists.347350

Table 4. Survey of Desalination Performance of MXene-Based Membranesa.

membranes operation conditions (P: bar, T: °C) ion type (concentration in g/L) water (* pure) permeate (L/m2 × h × bar) salt rejection (%) ref
Ti3C2Tx on PVDF (1.5 μm) AlCl3 (−) 18.0 (125)
Ti3C2Tx on PVDF (1.5 μm) MgCl2 (−) 24.2 (125)
Ti3C2Tx on PVDF (1.5 μm) NaCl (−) 45.7 (125)
Ti3C2Tx (Na+-intercalated) on PVDF (1.5 μm) AlCl3 (−) 25.6 (125)
Ti3C2Tx/GO (30 wt %) on PC (0.09 μm) 5, – NaCl (5.85) 2.3 0 (193)
Ti3C2Tx/GO (30 wt %) on PC (0.09 μm) 5, – MgSO4 (24.65) 2.5 10 (193)
Ti3C2Tx on PES (0.065 μm) 2, – Na2SO4 (1) 30* 97.5 (351)
Ti3C2Tx on PES (0.065 μm) 2, – MgSO4 (1) 30* 95.4 (351)
Ti3C2Tx on PES (0.065 μm) 2, – CaCl2 (1) 30* 52.0 (351)
Ti3C2Tx on PES (0.065 μm) 2, – NaCl (1) 30* 20.0 (351)
Ti3C2Tx on PES 1, 20 NaCl (1) 435 13.8 (210)
Ti3C2Tx on PES 1, 20 Na2SO4 (1) 632 13.2 (210)
Ti3C2Tx on PES 1, 20 MgCl2 (1) 460 23 (210)
Ti3C2Tx/CA (123 μm) 1, – NaCl (2) 256 28 (220)
Ti3C2Tx/CA (123 μm) 1, – Na2SO4 (2) 256 59 (220)
Ti3C2Tx/CA (123 μm) 1, – MgSO4 (2) 256 56 (220)
Ti3C2Tx/CA (123 μm) 1, – MgCl2 (2) 256 40 (220)
Ti3C2Tx/GO (30 wt %) (H2O2-treated) on MCE (0.262 μm) 1, 25 NaCl (0.29) 90 ∼39 (228)
Ti3C2Tx/GO (30 wt %) (H2O2-treated) on MCE (0.262 μm) 1, 25 Na2SO4 (0.71) 90 60.6 (228)
Ti3C2Tx/GO (30 wt %) (H2O2-treated) on MCE (0.262 μm) 1, 25 MgSO4 (1.23) 90 ∼31 (228)
Ti3C2Tx/GO (30 wt %) (H2O2-treated) on MCE (0.262 μm) 1, 25 MgCl2 (0.48) 90 22.5 (228)
Ti3C2Tx on α-Al2O3 (0.1 μm) 3, 25 NaCl (0.59) 6.2 55.3 (354)
Ti3C2Tx on α-Al2O3 (0.1 μm) 3, 25 Na2SO4 (1.42) 3.5 75.9 (354)
Ti3C2Tx on α-Al2O3 (0.1 μm) 3, 25 MgSO4 (2.46) 4.7 67.3 (354)
Ti3C2Tx on α-Al2O3 (0.1 μm) 3, 25 MgCl2 (0.95) 8.5 46.1 (354)
Ti3C2Tx (Al3+-intercalated) on PES (0.34 μm) OPb NaCl (5.85) 8.5 89.5 (367)
Ti3C2Tx (Al3+-intercalated) on PES (0.58 μm) OPb NaCl (5.85) 4.8 92.3 (367)
Ti3C2Tx/PA on PSF (0.205–0.375 μm) 16, 25 NaCl (2) 2.53 98.5 (352)
Ti3C2Tx/SA (Mn2+-intercalated) on PVDF (0.05 μm) 1, – Na2SO4 (0.05) 16.5 84 (357)
Ti3C2Tx/SA (Mn2+-intercalated) on PVDF (0.07 μm) 1, – Na2SO4 (0.05) 12.7 100 (357)
Ti3C2Tx/Al13 on PVDF (∼0.05 μm) OPb NaCl (5.85) 0.30 99 (357)
Ti3C2Tx on PAN (∼0.06 μm) 0.004, 65 NaCl (35) 85.4 99.5 (382)
Ti3C2Tx on PAN (∼0.06 μm) 0.004, 30 NaCl (35) 48.2 99.5 (382)
Ti3C2Tx (MA-cross-linked) on nylon (∼0.03 μm) 0.0013, 30 NaCl (35) 22.8 99.9 (356)
Ti3C2Tx (MA-cross-linked) on nylon (∼0.03 μm) 0.0013, 30 KCl (35) 22.99 99.90 (356)
Ti3C2Tx (MA-cross-linked) on nylon (∼0.03 μm) 0.0013, 30 Na2SO4 (35) 21.97 99.98 (356)
Ti3C2Tx (MA-cross-linked) on nylon (∼0.03 μm) 0.0013, 30 MgSO4 (35) 20.89 99.99 (356)
Ti3C2Tx (MA-cross-linked) on nylon (∼0.03 μm) 0.0013, 30 MgCl2 (35) 21.35 99.97 (356)
Ti3C2Tx/PVA/SSA composite on PTFE (0.23 μm) 0.0019, 70 KCl (44.73) ∼65c 99.81 (356)
Ti3C2Tx/PVA/SSA composite on PTFE (0.23 μm) 0.0019, 70 Na2SO4 (85.22) ∼55.3c 99.91 (356)
Ti3C2Tx/PVA/SSA composite on PTFE (0.23 μm) 0.0019, 70 MgSO4 (147.88) ∼55.4c 99.93 (356)
Ti3C2Tx/PVA/SSA composite on PTFE (0.23 μm) 0.0019, 70 MgCl2 (57.13) ∼52.3c 99.93 (356)
Ti3C2Tx/PVA/SSA composite on PTFE (0.23 μm) 0.0019, 70 CaCl2 (66.59) ∼54.2c 99.92 (356)
Ti3C2Tx/P84 mixed matrix (150–20 μm) 1, 20 Na2SO4 (1) 423 0 (222)
Ti3C2Tx/P84 mixed matrix (150–20 μm) 1, 20 MgCl2 (1) 418 0 (222)
Ti3C2Tx (casted with EPD) (0.2 μm) OPb NaCl (11.7) ∼99.6 (359)
Ti3C2Tx (casted with EPD) (0.2 μm) OPb MgCl2 (19.04) ∼99.6 (359)
Ti3C2Tx (casted with EPD) (0.2 μm) OPb AlCl3 (26.67) ∼99.6 (359)
Ti3C2Tx on PES (0.2 μm) OPb NaCl (11.7) ∼94.8 (359)
Ti3C2Tx on PES (0.2 μm) OPb MgCl2 (19.04) ∼96.5 (359)
Ti3C2Tx on PES (0.2 μm) OPb AlCl3 (26.67) ∼98.2 (359)
Ti3C2Tx/PA on PSF 3, 25 NaCl (2) 3.1 13.84 (358)
Ti3C2Tx/PA on PSF 3, 25 Na2SO4 (2) 4.7 97 (358)
Ti3C2Tx/PA on PSF 3, 25 MgSO4 (2) 4.9 92.35 (358)
Ti3C2Tx/PA on PSF 3, 25 MgCl2 (2) 3.2 48.36 (358)
Ti3C2Tx/PAN hollow fiber (∼0.09 μm) 1, 25 Na2SO4 (0.05) ∼5.9 ∼70 (383)
Ti3C2Tx/PAN hollow fiber (∼0.09 μm) 1, 25 MgSO4 (0.05) ∼6.0 ∼45 (383)
Ti3C2Tx/PAN hollow fiber (∼0.09 μm) 1, 25 MgCl2 (0.05) ∼6.9 ∼36 (383)
Ti3C2Tx/PAN hollow fiber (∼0.09 μm) 1, 25 NaCl (0.05) ∼6.2 ∼28 (383)
Ti3C2Tx (5 mg)/PA on PES OPb NaCl (0.23) 13.5 (362)
Ti3C2Tx (15 mg)/PA on PES OPb NaCl (0.23) 15.2 (362)
Ti3C2Tx (25 mg)/PA on PES OPb NaCl (0.23) 14.1 (362)
Ti3C2Tx (35 mg)/PA on PES OPb NaCl (0.23) 14.0 (362)
Ti3C2Tx/PA (immersed in organic phase) on PSF 4, 25 Na2SO4 (2) 2.33 98.6 (360)
Ti3C2Tx/PA (immersed in aqueous phase) on PSF 4, 25 Na2SO4 (2) 4.70 97.6 (360)
Ti3C2Tx (modified with PFDTMS) on PVDF (120 μm) –, 20 seawater 1.88c 100 (371)
Ti3C2Tx (modified with HTEOS) on PVDF (120 μm) –, 20 seawater 0.85c 100 (371)
Ti3C2Tx/PA on PDA OPb NaCl (58.5) 27.5 (384)
Ti3C2Tx/CNT/PA on PDA OPb NaCl (58.5) 31.5 (384)
a

Membrane thickness is given in parentheses. CA: cellulose acetate, CNT: carbon nanotube, EPD: electrophoretic deposition, GO: graphene oxide, HTEOS: hexadecyltrimethoxy silane, MA: maleic acid, MCE: mixed cellulose ester, PA: polyamide, PAN: polyacrylonitrile, PDA: polydopamine, PES: polyethersulfone, PFDTMS: 1H,1H,2H,2H-hepta decafluoro decyltrimethoxy silane, PSF: polysulfone, PTFE: polytetrafluoroethylene, PVA: poly (vinyl alcohol), PVDF: polyvinylidene difluoride, P84: commercial copolyimide, SA: sodium alginate, SSA: sulfosuccinic acid.

b

Osmotic pressure (OP) corresponding to 2 M sucrose.

c

Unit of permeate in kg/m2 × h.

Membrane technologies focused on designing highly selective nanofiltration, RO, and promising forward osmosis (FO) MXene-based membranes, which can selectively separate salt ions by either size- or charge-exclusion. Initially, TFN membranes were engineered by the deposition of MXene (Ti3C2Tx) nanosheets on the PES support via vacuum filtration and then an interfacial polymerization reaction was carried out to test the nanofiltration performance for divalent salt ions. For example, Xu et al.351 reported 97 and 95.3% rejection rates of this TFN membrane for Na2SO4 and MgSO4 with 45.7 L/m2 × h × bar of water permeance where it exhibited a selective separation factor (α(NaCl/MgSO4) of 14.5. To further enhance the performance of the MXene TFN membrane, MXene nanosheets were dispersed in aqueous solution during in situ interfacial polymerization and higher selectivity of the monovalent salt ions were achieved for RO technology. Wang et al.352 achieved the highest NaCl rejection rate of 98.5% and water flux of 2.5 L/m2 × h × bar compared to the other TFN membranes [see Figure 9(a)]. In addition to the experimental measurements, molecular simulation approaches support the optimal performance of MXene membranes for the RO process. Meidani et al.353 calculated the water and salt ion transport through the pore of three types of MXene (Ti2C, Ti3C2, and Ti4C3), graphene, and MoS2 membranes. Permeability coefficients were reported in the order of Ti3C2 > Ti2C > Mo2S > Ti4C3 > graphene under low pressure (<100 bar), which is generally applied in the simulation approaches, showing the applicability of MXene membranes for the desalination process.353 The Ti3C2 membrane exhibited the highest permeation rate of 11.4 L/cm2 × day × MPa (475 L/m2 × h × bar) with an ion rejection rate of 100% more than the rest of the investigated membranes.

Figure 9.

Figure 9

Desalination performance of MXene membranes. (a) Comparison of water permeability and salt rejection for the polyamide-based nanocomposite RO membranes (NaCl: 2 g/L, water pressure: 16 bar, temperature: 25 °C). Adapted with permission from ref (352). Copyright 2020 Elsevier. (b) Variation of permeate flux and salt rejection as a function of time (MXene-H and -P represent the hexadecyl trimethoxysilane (HTEOS) and 1H,1H,2H,2H-heptadecafluorodecyl trimethoxysilane (PFDTMS) grafted MXene, respectively). Adapted with permission from ref (371). Copyright 2022 Elsevier. (c) Illustration of the electrophoresis deposition process for the preparation of MXene membranes. Adapted with permission from ref (359). Copyright 2021 Elsevier. (d) Digital photo of the EPD-MXene membrane having a large area (top) and SEM images of the cross-section of the EPD-MXene membrane (bottom). Adapted with permission from ref (359). Copyright 2021 Elsevier. Schematic illustration for hybrid capacitive deionization process of (e) the charging (desalination) and (f) discharging (regeneration) steps. Adapted with permission from ref (381). Copyright 2021 Elsevier.

GO is widely studied and accepted as the next-generation material in membrane technology due to its chemical resistance, mechanical robustness, and selective separation of ions. Although the fabrication of the nanocomposite membrane by combining GO and MXene has been suggested by researchers as another strategy to design high-performance membranes for desalination, satisfactory results have not been achieved yet. For instance, Ti3C2Tx/GO nanocomposite membranes were prepared by mixing different GO concentrations (10–30 wt %) with Ti3C2Tx by Kang et al.193 Rejection rates of NaCl and MgSO4 were below 11% with the permeances of 2.25 and 2.35 L/m2 × h × bar, respectively, due to the swelling where the optimum interlayer spacing was determined around 5 Å for the nanocomposite membrane. Likewise, Han et al.228 fabricated Ti3C2Tx/GO nanocomposite membranes with various MXene amounts and then applied H2O2 to create TiO2 nanocrystals via in situ oxidation. Pure water permeability of nanocomposite membranes reached from ∼31.34 to 108.7 L/m2 × h × bar with an increase in the MXene amount from 10 to 40%. However, salt rejection rates for Na2SO4, NaCl, MgSO4, and MgCl2 were still low as 60.6, ∼39, ∼31, and 22.5%, respectively. Nevertheless, in order to obtain higher separation efficiency in the desalination process, more research on MXene/GO nanocomposite membranes should be carried out.

To improve the desalination performance of MXene-based membranes, one of the efficient strategies is adjusting the interlayer spacing of nanosheets. Sun et al.354 fabricated MXene-based membranes with 0.1 μm thickness on α-Al2O3 tubular supports and used a sintering-temperature regulation method to design the interlayer distance that selectively separated ions. The interlayer distance decreased from 3.71 to 2.68 Å with the increase in temperature from 60 to 400 °C due to the change in functional group concentration on the surface of MXene, which results in the cross-linking via thermal treatment. The pure water permeance of membrane treated at 400 °C was reported as 11.5 L/m2 × h × bar, and the improvement in rejection rates of the MXene membrane with the rise in treatment temperature from 200 to 400 °C was obtained as 32.3, 31.4, 34, and 43.3% for Na2SO4, MgSO4, NaCl, and MgCl2, respectively. Using a similar approach as the thermal treatment at 180 °C, Lu et al.355 confined the interlayer spacing of MXene via a self-cross-linking strategy from 13.61 to 12.87 Å in the dry state and from 16.68 to 15.54 Å in the wet state. Thanks to the self-cross-linking mechanism of MXene, the permeation rate of the monovalent ions such as K+, Na+, and Li+ reduced from the order of 10–1 to 10–3 mol/m2 × h and a comprehensive increase was observed in the NaCl rejection rate, reaching to 98.6%. They attributed the interlayer distance shrinkage to the thermal treatment induced by the self-cross-linking behavior. With the increase in temperature, bonded water and −(OH)2 groups on the surface of MXene nanosheets were decreased and then Ti–O–Ti bonds (each Ti from neighboring sheets) were formed leading to the self-cross-linked MXene and confined interlayer spacing. As a result, the swelling of 2D membranes and poor ion rejection were overcome.

Swelling is considered as a big challenge for 2D membranes, since it leads to a low rejection rate of small ions during desalination. With the exposure of membrane to water, they are usually swelled resulting in the enhanced d-spacing, decreased stability, and poor ion-sieving. Although MXene nanomaterials are being used in the membrane separation technology more day by day because of their outstanding properties, their hydrophilic nature was proposed as the prospective reason for swelling. Several strategies have been applied to prevent enhancement of the interlayer space between nanosheets such as self-cross-linking,354,355 chemical cross-linking,356 and cation intercalation.357 Ding et al.356 used maleic acid for the chemical cross-linking of Ti3C2Tx nanosheets, which led to the change in d-spacing as 12.5% in contact with water. However, the d-space of pristine Ti3C2Tx increased 149% in the wet state. Chemical cross-linking between hydroxyl groups of MXene and carboxyl groups in maleic acid prevented the membrane from swelling. A covalently bridged Ti3C2Tx membrane revealed high salt rejection of >99.9% in different salt solutions (3.5 wt %) as KCl, MgCl2, Na2SO4, and MgSO4 and water flux ranged between 20 and 23 kg/m2 × h. In another study, to overcome the swelling of the PVA pervaporation desalination membrane, MXene (Ti3C2Tx) nanosheets were inserted and PVA was cross-linked with sulfosuccinic acid.356 The swelling degree of this composite membrane (48%) was lower than the pure PVA membrane (283%). This severe hindrance in swelling was attributed to the mutual effect of the formation of an ester linkage between PVA and SFA, and hydrogen bonds between sulfonic/ester groups and functional groups on the surface of MXene, possibly occurring during annealing. The water flux of the composite membrane was 62.2 kg/m2 × h with a NaCl rejection rate of 99.8% and was kept constant during a 50 h desalination process at 30 °C. Wang et al.357 suggested an efficient combined strategy using SA and various cations such as Ca2+, Ba2+, and Mn2+ together, where the former has hydrogen bonding ability with the MXene layer and the latter can electrostatically interact with the former. Unfortunately, with the immersion of pristine MXene in various salty solutions including NaCl, LiCl, RbCl, MgCl2, CaCl2, AlCl3, and NH4Cl, its interlayer spacing difference was reported as 1.79 Å between the solutions leading to the maximum and minimum interlayer distance. However, it was only 0.52 Å for the MXene membrane anchored with SA on the nanosheet surface and intercalated with Ca(II). Similar swelling resistance performance was observed for MXene membranes intercalated with Ba2+ and Mn2+ with an interlayer spacing constrained at 16.2 ± 0.2 Å. As a result, an improved separation performance was achieved, although ion-sieving mechanisms changed due to the affinity of SA molecules to the cations. For instance, the MXene membrane anchored with SA on the nanosheet surface and intercalated with Mn2+ having thicknesses of 0.05 and 0.08 μm showed high permeances of 16.5 and 12.7 L/m2 × h × bar and Na2SO4 rejection rates of 84 and 100%, respectively. The other study that proves the success of intercalation of the cation within MXene nanosheets was carried out by Zhu et al.357 They fabricated a Keggin Al13 polycation intercalated Ti3C2Tx membrane via vacuum filtration and aimed to utilize from the electrostatic interaction between the Keggin Al13 ion (zeta potential: +24 mV) and Ti3C2Tx (zeta potential: −31.3 mV). The pristine Ti3C2Tx membrane swelled after soaking into the various salt solutions such as LiCl, NaCl, KCl, MgCl2, CaCl2, and AlCl3 for 1 h, and its d-spacing varied between 14.7 and 22.1 Å (difference: 7.4 Å). However, the Keggin Al13 ion intercalated Ti3C2Tx membrane showed little variance for d-spacing as 0.3 Å with nearly suppressed value at around 11.5 Å even at wet or dry conditions. Suppressed swelling phenomena led to the NaCl rejection of 99% under osmotic pressure (2 M sucrose). Therefore, the proposed antiswelling strategies are encouraging for the desalination MXene membranes in adjusting the interlayer spacing with angstrom precision and, hence, improving the salt rejection.

Maintaining ion-sieving performance of membranes as long as possible during and after long consecutive operation cycles is the other challenge in desalination. Research to test the long-term structural and performance-based stability of membranes is emerging. Therefore, either long-term measurements355,356,358 or long time water immersion tests357,359,360 are carried out in order to provide knowledge about the stability. The longest desalination test was carried out about 58 days for Ti3C2Tx-based TFN membranes without the decrease in salt rejection.358 The other study carried out by Lu et al.355 displayed the effect of self-cross-linking via thermal treatment on the long permeation measurements as >70 h. Additionally, self-cross-linked MXene displayed structural stability up to 12 h after immersion in solutions at different pH values such as 1.5, 7.0, and 11.3. Although not as long as the permeation tests of Lu et al.,355 Yang et al.356 also achieved stable water permeation for the MXene-based nanocomposite membrane for 50 h. Wang et al.357 soaked the MXene membrane, prepared by anchoring SA on a nanosheet surface and intercalating Ca2+ ions, in water for 20 days and observed almost the same NaCl permeation rate of a fresh MXene membrane. On the other hand, at the end of the third permeation/drying cycles, its NaCl permeation rate still did not change, proving its long-term usability. Likewise, different salt solutions such as NaCl, MgCl2, KCl, and AlCl3 were applied for the consecutive cycles lasting 4 h, and ion permeation rates did not change significantly, indicating its superb regeneration ability. The longest immersion was carried out by Xue et al.,360 where the MXene-based TFN membrane was immersed in water for 105 days with no change in Na2SO4 rejection and slight decrease in flux.

Performance-based stability of membranes in desalination is mainly hindered by fouling phenomena, which is the blockage of membrane pores by foulants such as proteins, lipids, bacteria, etc., and their accumulation on the membrane surface, leading to the reduction in efficiency of process and increase in cost due to the requirement of greater pressure. Three types of fouling are observed: organic fouling, deposition of salts, and biofouling. There are many factors affecting the membrane fouling such as water characteristics, process conditions, and most importantly properties of the membrane itself. Thus, to develop a membrane with low fouling tendency is so important for the effective desalination performance. Generally, BSA solution is used to test the fouling behavior of membranes. For instance, Shen et al.361 reported that the flux recovery ratio of the Ti3C2Tx membrane on a polysulfone (PSF) support was 76.1%, while it was only 48.3% for the pristine PSF membrane after physical cleaning, indicating improved antifouling properties of MXene membranes. Another study that used BSA to test the membrane fouling was done by Pandey et al.220 The Ti3C2Tx/CA composite membrane cross-linked with formaldehyde exhibited 100% rejection rate for BSA, whereas the pristine CA membrane showed a lower rejection rate of 65.3%.220 An improvement of 35% for the BSA rejection was explained as the formation of small pores and reduction in macrovoids via addition of MXene into the CA matrix and chemically cross-linking. Encouragingly, Xu et al.362 fabricated a novel membrane where MXene nanoplatelets were deposited on the outer surface of PES via vacuum filtering and selective polyamide (PA) layer was synthesized by the interfacial polymerization on the other side of PES. This novel layered membrane revealed the decrement in water flux of only 8.2% and 4.5%, respectively, without and with an applied electric field of +2 V.362 On the other hand, Tan et al.130 fabricated MXene (Ti3C2Tx) as a coating material via vacuum filtration onto the PVDF support membrane to mitigate fouling in a DCMD process. They observed a water flux decrement in the absence and presence of light irradiation as only 8.3 and 6.6% for the MXene-coated membrane, whereas those of 18.8 and 18.2% were reported for the PVDF membrane, respectively, after 21 h of continuous filtration process.130 It was ascribed to its repulsive interaction with BSA and its ability to accommodate BSA within its intercalated structure.

Similar as fouling, performance stability of a membrane is also influenced by biofouling, which is a complex, reversible, or irreversible adhesion of microorganisms on a membrane surface and their growth within a couple of weeks or months. Biofouling has severe adverse effects on membrane processes such as flux decline, membrane biodegradation, rise in energy consumption, and decrease in salt rejection. Although biofouling has been known as a contributing factor to more than 45% of all membrane fouling,363 limited studies have investigated the biofouling resistance of MXene membranes. Pandey et al.46,220 showed the antibacterial capacity of Ti3C2Tx and two niobium-based carbide MXenes (Nb2CTx and Nb4C3Tx). They reported the growth inhibition of bacteria of almost >98% and 96% for E. coli (Gram-negative bacteria) and B. subtilis (Gram-positive bacteria) for the membrane consisting of 10% MXene content, respectively.220 Similarly, high growth inhibitions of Nb2CTx and Nb4C3Tx were recorded as 94.2 and 96.1% for E. coli and 91.6 and 93.7% for S. aureus within 3 h of incubation, respectively.46 Pandey and colleagues attributed this exceptional antibacterial activity of MXenes to their nanoknife behavior toward the bacteria by the help of their sharp edges. However, instead of fabricating MXene-based composite membranes, Zha et al.364 suggested that coating Ti3C2Tx MXene nanomaterials on the surface of a cellulose membrane was more efficient. Cell viability was much more hindered, and the highest antibacterial efficiency of 99.99 and 99.98% for E. coli and S. aureus were observed after 24 h of contact time, respectively.364 Collectively, not only the MXene type but also the membrane preparation procedure has significant effects on cell viability. In addition to the studies of the group of Pandey,46,220 Jastrzȩbska et al.365 identified the biological activity of different types of MXene nanomaterials as Ti3C2Tx and Ti2CTx for E. coli. They concluded that atomic structure with specific stoichiometry is crucial for the bioactivity of MXenes rather than the chemical composition. Antibacterial capacity of specific types of MXene was elaborated in the review of Seidi et al. in depth.366 However, there are more than hundreds of stoichiometric MXene compositions (at least) and a countless number of solid solutions. Therefore, more research needs to be done to determine complete antibacterial capacity of the MXene family.

The chlorination process is mostly used in water remediation to reduce membrane biofouling. However, chlorine can damage the membrane and reduce its separation performance. Therefore, producing a chlorine-resistant membrane is the key point for the desalination process. Ding et al.367 investigated the chlorine resistance of pristine and Al3+-intercalated Ti3C2Tx membranes by treating them with 200 mg/L NaClO for 24 h. While Na+ permeation rates of pristine MXene were increased almost 5-fold, those of the Al3+-intercalated MXene membrane were only doubled after treatment.367 Moreover, Al3+-intercalated MXene membranes showed high and nearly stable Na+ rejection rates of 98% after chlorine treatment, whereas it was below 40% for the pristine MXene membrane. Harsher conditions were also tested for defining the chlorine resistance of MXene membranes. For instance, Wang et al.352 applied 5 chlorine cleaning cycles using 2000 mg/L NaClO solution to test the Ti3C2Tx-based TFN membrane. Almost stable water flux and salt rejection of >97% were reported for the TFN membrane, whereas the water flux of the PA membrane increased 3.5-fold along with the decrease in salt rejection rate to 94% during 5 chlorine cleaning cycles. Most importantly, the TFN membrane with a constant salt rejection rate of 97.1% was unveiled after the chlorination test of 10,000 mg/L × hour NaClO, proving its excellent chlorine resistance property.352 The reason for the excellent chlorine resistance of the TFN membrane was attributed to the interaction between the surface functional groups of the MXene nanosheets and chlorine, which was utilized to prevent the chlorine attack on the PA matrix.

Seawater contains salts, organic compounds, minerals, and microbial organisms, which influence the selective separation of membranes. The most feasible way to reveal the accurate performance of membranes in a lab or pilot scale is to investigate their performance for artificial/simulated water, which mimics real seawater. Zhao et al.368 tested the solar desalination performance of the Ti3C2Tx membrane from the simulated water containing RhB and MO dyes and Cu2+ and Cr6+ ions and achieved high salt rejection rates varied in the range of 99.6–100%. Likewise, the MXene membrane displayed a decrease in salt ion concentrations below ∼10 mg/L, which is suitable for the drinking water standards of WHO, from Bohai seawater including Na+, K+, Mg2+, and Ca2+ ions.368 Similarly, for the separation of various real seawater samples with the different salt concentrations (Yellow sea: 31‰, Bohai sea: 30‰, Alkali lake 1–2:12–76‰), Zhang et al.369 tested the vertically aligned Janus MXene (Ti3C2Tx) aerogel. The salinity ratio decreased under the drinking water standards of WHO (1‰) and EPA (0.5‰) after performing solar desalination tests using the MXene aerogel. Moreover, Wang et al.370 reached the ion concentration ranging between 0.16 and 1.12 mg/L for Na+, K+, Mg2+, and Ca2+ ions in the purified water using five-layered cotton fabrics coated with MXene and CNT. Zhang et al.371 suggested that the silane-modified MXene membrane was stable during the 120 h treatment of actual seawater, as illustrated in Figure 9(b). The encouraging desalination performance of MXene-based membranes from both real and simulated seawater proves that the MXene nanomaterial can be used safely in membrane technology.

The main desire is for a novel membrane material to find a place in the membrane market. Therefore, scientists and/or engineers should concentrate on its scalability, which is the most important challenge in the membrane production step. Wu et al.372 proposed a facile and novel method to create a Ti3C2Tx interlayered polyamide membrane having a large effective membrane area. This method consists of brush coating of the MXene solution and then interfacial polymerization. The MXene TFN membrane with an area of 24 × 8 cm2 exhibited low salt fluxes of 0.07 and 0.27 g/L and high-water fluxes of 15.6 and 23.9 L/m2 × h in the active layer faces to the feed and draw solutions, respectively.372 In a different way, Deng et al.359 prepared the Ti3C2Tx membrane with a large area of 575 cm2 via EPD within only 10 min. Rejection rates of Na+, Mg2+, and Al3+ reached to 99.6% for the membranes fabricated via EPD compared to the MXene membranes synthesized via traditional vacuum filtration with a salt rejection of 94.7%. It was revealed that EPD led to the deposition of big MXene nanosheets onto the support, which was called “smart selection” [see Figure 9(c)], resulting in a more ordered MXene membrane, as given in Figure 9(d). This method seems as scalable as brush coating and interfacial polymerization, but unfortunately it is cost-intensive. Unfortunately, more research still needs to be carried out in the membrane engineering on the design of both scalable and easy production methods to fabricate MXene membranes with a high desalination performance.

MXene-based membranes were used in nanofiltration, RO, and FO processes for the separation of salt ions by either size- or charge-exclusion. With the help of electrostatic interactions between ions and the MXene surface, the expansion or shrinkage of interlayer distance was directed by the ion-sieving mechanism successfully in the desalination process. Either to enhance their desalination performance or inhibit the swelling phenomena, the main strategy followed was adjusting the interlayer spacing of the nanosheets by self-cross-linking via thermal treatment,354,355 chemical cross-linking,356,373 and cation intercalation.357,374 Considerable studies were carried out for identification of swelling, long-term stability, and fouling behavior of MXene nanomaterials with great success. Especially, we believe that the stability of the MXene membrane during the desalination process is well-documented, reaching up to 105 days with no alteration in Na2SO4 rejection. Since antifouling, antibiofouling, and chlorination resistance of a specific type of MXene were revealed, other members of this fastest growing 2D inorganic nanomaterial family should be identified as well. The most promising and thrilling observation for MXene-based membranes in desalination is their scalability, which will enable its usage in the near future in industrial application.

8.2. Solar Desalination

Alternative energy sources have gained vital importance these days, where the limit of global warming is expected reach to above 1.9 °C by 2100.375 Therefore, the use of plentiful sunlight in the desalination process has garnered increasing attention due to its environmentally friendly and low-cost technology. Considering the capacity of MXene to convert the light efficiently to heat, it would not be a surprise to encounter the usage of the MXene family for the solar desalination process.364,368370,376379 Zhao et al.368 used a Ti3C2Tx membrane, which was modified with trimethoxy (1H,1H,2H,2H-per-fluorodecyl) silane (PFDTMS), in solar desalination for the first time. The solar evaporation rate of the MXene membrane was reported as 1.31 kg/m2 × h with a solar steam conversion efficiency of 71% under 1 sun. Additionally, its solar steam generation performance was almost constant for 10 cycles under 1 sun, proving its excellent durability. Salts were accumulated under the membrane, and no deterioration was observed onto the upper surface of the membrane. For the MXene/cellulose composite membrane, both higher steam conversion efficiency (85.8%) and water evaporation rate (1.44 kg/m2 × h) were observed by Zha et al.364 than the data provided by Zhao et al.368 Similar results (conversion efficiency of 87%, and water evaporation rate of 1.46 kg/m2 × h under 1 sun for 15 days) were displayed by Zhang et al.369 for the vertically aligned Janus Ti3C2Tx aerogel modified with the fluorinated alkyl silane. Solar-assisted water evaporation technology is also used for the textile water purification, in addition to the seawater desalination. Wang et al.370 designed cotton fabrics by coating first with Ti3C2Tx and then with CNT in a manner of layer-by-layer assembly. Evaporation rates of distilled and textile wastewater were proposed as 1.35 and >1.16 kg/m2 × h, respectively, with high efficiency of 88.2% for five-layered cotton fabrics under 1 sun. The other nanocomposite membrane was fabricated by Ming et al.376 for the same purpose by combining GO and Ti3C2Tx using SA as a binder. The evaporation rate of this aerogel increased to ∼1.27 kg/m2 × h within only 10 min with an evaporation efficiency up to 90.7% under 1 sun. Moreover, its evaporation rate and efficiency enhanced to 6.70 kg/m2 × h and ∼93.9% under 5 suns. Collectively, these research studies suggested that MXene-based membranes have a great potential for water remediation via solar driven membrane-based desalination processes.

8.3. Capacitive Deionization

The capacitive deionization (CDI) is an alternative system for the desalination where dissolved ions were separated from water by electrodes via applying a small electrical voltage, as illustrated in Figure 9(e,f). The efficiency of this electrochemical separation process is hindered by the selection of electrode material. MXene nanomaterials are also promising electrode materials for CDI due to their high electrical conductivity, hydrophilicity, booklike structure, and tunable surface chemistry. Ma et al.380 investigated the desalination capacity of Ti3C2Tx films having less fluorine terminal groups. Average desalination capacity of MXene was observed as 68 mg/g with the initial NaCl concentration of 585 mg/L at a current density of 20 mA/g and voltage window of 1.2 V for 3 cycles. Similar performance (desalination capacity of 72 mg/g) but higher durability was observed by Shen et al.380 for the hybrid MXene electrode produced by the intercalation of MXene nanosheets with small size into the larger ones by vacuum filtration. The desalination capacity was kept constant even after 50 cycles at 1.4 V with a 5 mM NaCl initial concentration. The best desalination performance was obtained for the Ti3C2Tx/Na2Ti2(PO4)3 nanohybrid electrode designed by Chen et al.381 Desalination performance of the nanohybrid electrode increased from 32.3 to 128.6 mg/g at the operation voltage of 1.8 V with a NaCl concentration from 250 to 1000 mg/L. In addition, the desalination capacity was maintained above 100 mg/g without significant impairment for 20 cycles. This success was related to the cooperation of MXene, which increases electrical conductivity and provides significant interlayer spacing for ion transport. As a result of these outstanding data, MXene will increasingly take place in CDI as an electrode. The list of other promising studies that incorporate MXene in the CDI process are listed in Table S7.

9. Other Prominent Applications

Since the discovery of the MXene family in 2011 by Gogotsi and co-workers, it has become one of the fastest growing family of 2D inorganic materials. To date, approximately 150 MAX phases20 have been explored experimentally and dozens of MXene forms have been revealed by the computational approaches. Due to their attractive properties such as metallic conductivity, elastic mechanical strength, hydrophilicity, surface chemistry, and interlayer engineering properties, the MXene family has been dominantly tested for the different emerging areas such as the removal of toxic substances131133 and pharmaceuticals134136,385387 from water as well as biomedical138141 and controlled release142 applications.

Phosphate and nitrate ions are soluble, toxic, and carcinogenic substances for living organisms. When their amount increased in water, methemoglobinemia and eutrophication threaten human health and water safety directly. MXene nanomaterials were also evaluated for the separation of these substances. The first study reported by Zhang et al.132 investigated the phosphate adsorption performance of a sandwich-like layered Ti3C2OH0.8F1.2/Fe3O4 nanocomposite. Thanks to the Fe3O4 molecules coordinated on the MXene surface, adsorption of the phosphate ion significantly increased within the first 20 min. The MXene/Fe3O4 nanocomposite composed of a weight ratio of 2/1 displayed a phosphate adsorption of 9.42 mg/g from an initial ion concentration of 10 mg/L with the removal efficiency of ∼87% at pH ∼6.132 Importantly, the treatment capacity was 2100 kg per water with 2 mg/L phosphate and 200 mg/L nitrate in the effluent (sewage discharge standard) when synthetic wastewater including various ions was used, whereas it improved to 2400 kg per water with the same phosphate concentration in the effluent using real wastewater.132 Although it was stated that the reason for this performance was the synergy between MXene and Fe3O4, Karthikeyan et al.131 proved that MXene alone has the same separation performance. Adsorption performance of pristine Ti3C2Tx for phosphate and nitrate ions were reported as 84 and 66 mg/g from the initial ion concentration of 100 mg/L, respectively, at pH 6. The removal mechanism of such toxic anions was attributed to the electrostatic interaction and surface complexation. Although promising studies exist, more research is required on the removal of toxic anions via MXene in the near future.

Another factor that threatens water safety is pharmaceutical compounds that have been found at trace levels in most aquatic environments due to the discharge of domestic sewage and industry. Considering the adverse effects of these compounds on human health as well as the ecosystem because of their toxicity and resistance manner, the application of efficient separation methods is the main issue for the water remediation. Adsorption has been preferred as a separation method related to its simplicity and low energy requirement even though it was restricted by the trade-off between adsorption capacity and recyclability involving as many cycles as possible. For the separation of pharmaceutical compounds, the potential of Ti3C2Tx as an adsorbent and membrane was identified by a few studies134136,386,387 and further investigation is essential to completely propose its capability. Kim and colleagues136 investigated its removal performance and mechanism toward five pharmaceuticals with a changing molecular weight from 206 to 296 g/mol. The effect of sonication time and frequency on the removal rate of MXene was identified, where the removal rate of amitriptyline (AMT) reached to ∼66% within 120 min for the sonicated Ti3C2Tx with 28 kHz while it was only ∼47% for the pristine MXene gained within 720 min.136 As accepted, the more negatively charged MXene surface with the functional groups of −(OH)2 and −(COOH)2 due to the sonication yielded in the greater removal rate. Due to the positively charged AMT compared to the other investigated pharmaceuticals, MXene affinity to this cationic pharmaceutical was greater (>80 mg/g).136 A similar observation was proposed by the study of Ghani et al.,134 which defined a high adsorption capacity of Na+ intercalated MXene nanosheets for another positively charged pharmaceutical, ciprofloxacin (CPX). Its maximum adsorption capacity for CPX was reported as 208.2 mg/g with the removal efficiency of ∼84% at pH 5.5. To further improve CPX removal of Ti3C2Tx, Ren et al.385 applied a rotating magnetic field for a composite adsorbent as MXene/CoFe2O4 cross-linked with SA. The created magnetic field induced the spin polarization of electrons, which facilitates the activity of functional groups of adsorbents and, hence, reduces the adsorption energy. In addition to the adsorption-based separation application of MXene, they can be used for a membrane-based separation process. Recently, Li et al.135 proposed a comprehensive study revealing the organic solvent nanofiltration performance of Ti3C2Tx membranes from seven different pharmaceuticals. As the molecular weight of the antibiotics increased, water or ethanol permeances enhanced with the improved rejection rates regardless of whether the antibiotic is soluble in water and ethanol, or not. The highest water permeance and rejection rate were recorded as 340.5 L/m2 × h × bar and 99.5% for bacitracin with a molecular weight of 1422 g/mol, whereas for penicillin with a molecular weight of 334 g/mol, the lowest permeance and rejection rate were determined as 223.1 L/m2 × h × bar and 89.5%.135 This proved that the separation mechanism of antibiotics depended on the size-selective molecular sieving with the help of the optimum interlayer distance. On the other hand, the rejection rate of tetracycline was about 97% at both pH 2 and 7, whereas it reduced to 91% at pH 4, ascribed to the change in the charges of membrane surface and antibiotic.135 Similarly, Gao and Chen387 reported a lower rejection rate of trimethoprim (80.2%) than that of sulfamethoxazole (85.5%) for the TFN membrane fabricated via interfacial polymerization of the MXene/cellulose nanocrystal nanocomposite, although the former has larger molecular weight as 290.3 g/mol than the latter (253.3 g/mol). Therefore, it was evidently revealed that electrostatic interactions dominated by the surface charge of the membrane also participates in the rejection performance of MXene. Thanks to these preliminary tests, they serve as a good starting point for this virgin field.

The separation of urine and excess water from the blood of humans is the kidney’s duty. The drop in its functioning below to 10% leads to the last phase of chronic kidney disease, end-stage renal disease (ESRD). For the treatment, renal replacement therapy with either dialysis or a kidney transplant was given to 2.6 million patients all over the world and this is estimated to double in number by 2030.388 In recent years, the wearable artificial kidney (WAK) emerged as a crucial alternative to transplantation and dialysis. Although there are some major challenges in the application of WAK, research on its development continues unabated. Few members of the MXene family were also examined by some groups to identify their adsorption performance of urea.138140 Meng et al.138 tested the adsorption capacity of three different MXene nanomaterials, namely Ti3C2Tx, Ti2CTx, and Mo2TiC2Tx for urea separation from the dialysate. The highest urea adsorption capacity was determined as 9.7 mg/g with a removal rate of 84% for Ti3C2Tx (0.155 g) at room temperature, whereas the adsorption capacity of Mo2TiC2Tx and Ti2CTx were only ∼2.82 mg/g and ∼1.21 mg/g for the initial urea concentration of ∼30 mg/dL, respectively. Increasing temperature to the 37 °C, which is the body temperature, caused a further improvement in the maximum adsorption capacity of Ti3C2Tx as 10.4 mg/g.138 Fortunately, it was also proved that Ti3C2Tx displayed a good biocompatibility and did not have any negative effect on the blood clotting cascade, according to the pro-thrombin and partial thromboplastin coagulation analyses. Considering the actual conditions where there are some other competitive species such as creatinine and uric acid that exist in the dialysate, its removal efficiency of urea dropped to below 20%.138 This challenge was handled by the study of Zhao et al.139 via etching Ti3C2Tx using different concentrations of HF. Ti3C2Tx etched with 10 and 30 wt % HF displayed the removal efficiency of 99.7 and 55.3% for the creatinine with the initial adsorbent amount of 100 mg.139 This high removal rate at the low HF concentration was attributed to the presence of more hydroxyl and oxy groups but less fluoride surface terminations. To clarify the removal capacity of MXene under actual dialysis conditions, simulated dialysate including Na+, K+, Mg2+, Ca2+, chloride, acetate, bicarbonate, and dextrose were prepared. For simulated dialysate, adsorption capacities of creatinine and uric acid were 38.4 and 20.0 mg/g whereas those from aqueous solution were 45.7 and 17.0 mg/g, respectively.139 Therefore, it was suggested that by only adjusting the HF concentration during the etching process of MXene, the removal rate of Ti3C2Tx under real conditions could be improved. After these promising trials for the specific well-known MXene types,138,139 one important question has emerged that “What is the urea adsorption performance of the different MXene types?” To answer this question, Zandi et al.140 carried out molecular dynamics simulations to predict urea adsorption capacity of six different members of the MXene family such as Mn2C, Cd2C, Cu2C, Ti2C, W2C, and Ta2C. Considering the longer time requirement for the laboratory or clinical research, molecular simulation approaches are the best method to find a suitable nanomaterial within a large material pool in a time- and cost-effective way. Zandi et al.140 predicted the greatest urea adsorptions for Cd2C and Mn2C as ∼13.35 and ∼10.73 mg/g, respectively, in accordance with the order of vdW interaction energies (Cd2C > Mn2C > Cu2C) and total numbers of hydrogen bonding. These studies revealed that the rapidly worldwide-recognized MXene family could find a strong place in biomedical application.

The applicability of these MXene adsorbents was determined by the stability tests via continuous adsorption–desorption cycles. Removal efficiency of MXene131 for phosphate was around 80% with around a 17% drop after six consecutive cycles, while those for the MXene/Fe3O4 nanocomposite132 was ∼77% with around a 9.1% drop at the end of the 5th cycle. On the other hand for the pharmaceutical separation, MXene regenerated with the electrochemical regeneration method reached the removal efficiency of 99.7% after four cycles for CPX.134 Moreover, adsorption performance of the sonicated MXene at 28 kHz for CPX decreased only 1% at the end of the 5th cycle at pH 7.0.136 However, that of MXene/CoFe2O4 for CPX decreased ∼17% after five cycles.385 Although promising regenerability performance was suggested by these studies for MXene, unfortunately, the tested number of cycles are not enough to evaluate their applicability.

In summary, MXene membranes and adsorbents have also become rising stars for the above-mentioned distinct applications such as the separation of toxic ions, pharmaceutical, and biological compounds. Although there is not sufficient data for its practical and long-term use, its separation capacity was revealed with success. To sum up, the journey of the MXene 2D family has just started; therefore, scientists need long, detailed, and arduous studies to unveil the real performance of the MXene family.

10. Summary and Perspectives

The MXene family is a burgeoning class of inorganic nanomaterial families with more than 50 members, having a proven track record of intriguing characteristics. Since MXene 2D nanosheets can be stacked in parallel into ultrathin films, forming the subnanometer channels between nanosheets with a molecular size-sieving ability, they are accepted as the potential membrane nanomaterials in many separation applications. Moreover, the MXene family is also very popular for the adsorption-based separation applications due to having a high surface area, tailorable surface chemistry, and adjustable interlayer distance. As such, here in this contribution, we have reviewed the membrane and adsorbent fabrication strategies and their related properties for several separation applications in detail. In particular, we have proposed a roadmap for the development of the application of the MXene family in both membrane- and adsorption-based separation technologies. This map reveals the efficient adaptation of MXene nanomaterials into high-performance membranes and adsorbents by highlighting the relevant figures of merit. Therefore, researchers who aim to benefit from MXene nanomaterials in these two emerging separation technologies can find a foothold and determine where to begin in this booming investigation field. To unveil the potential of each MXene member for each separation application that was covered in this review, we believe future efforts should be dedicated to the following: (i) Almost thousands of distinct compositions have been accumulated, including surface terminations, multielement, multilayered MXenes, etc.157 However, many more layered ternary metal carbides and nitrides or more complex ones are waiting to be transformed to MXene using either MAX or non-MAX phase etching methods. Their synthesis is an important research direction to further improve the performance of the MXene family in the target areas. (ii) Additionally, it is also crucial to investigate MXene types, which have already existed in the published literature but have not been considered for the specific separation applications where they would excel. (iii) The surface engineering was used to tune the surface-related MXene properties, which further influence the separation properties of membranes and adsorbents. To determine the impact of different modification methods of MXene surfaces on separation performance, coupled with either varying Tx or covalently bonding specific molecules to the Tx, should be investigated in more depth. There are some contradictory and incompatible results. (iv) New strategies to regulate the nanochannels to a stable interlayer distance are also in high demand in order to prevent swelling, not to encounter flux decline during membrane-based separation and to hinder the surface area required in adsorption-based separation. (v) Nanocomposites of MXene are proposed as another strategy to further enhance the separation performance with the mean contribution of MXene and other inorganic nanomaterials. Similarly, MXene nanomaterials have been widely envisaged as fillers in diverse polymeric matrices to fabricate MMMs. However, several hurdles still remain to be solved for MXene-based MMMs and nanocomposite membranes. These are the swelling, fouling, and instability problems, which hinder the application of MXene-based MMMs and nanocomposite membranes. (vi) In the adsorption-based separation processes, the improvement in the regeneration cycle of MXene adsorbents is a critical issue in a bid to boost their cost-effectiveness. Although promising regenerability performance compared to MOF and COF nanomaterial families was reported theoretically for the MXene family,157 unfortunately, the tested number of cycles are not enough to evaluate their applicability in terms of experimental studies. (vii) Similarly, in the membrane-based processes, mitigation of fouling is an important issue to retain the membrane flux and hence its applicability. Fouling appears to be a complex phenomenon that may involve multiple counteracting factors. Ti3C2Tx was suggested as the promising nanomaterial to decrease the fouling effect and the energy requirement for some of the separation processes. However, it is required to identify the fouling behavior of not only Ti3C2Tx but also other MXene types. (viii) Thick membranes endow the prolonged mass transfer path with the increased mass transfer resistance, which is against high flux. To hinder mass transfer resistance, thin membranes are required, which can be enabled by benefiting from the MXene 2D nanomaterials. However, the fabrication of thin MXene membranes in a defect free form should be investigated in depth. (ix) The most critical issue for MXene nanomaterials is the oxidation stability of the delaminated MXene dispersions. To extend the shelf life of MXene aqueous dispersions is required without the need for any additional chemicals. In addition to the promising experimental advancements in the oxidation stability of the Ti3C2Tx MXene type, the mechanism of the oxidative degradation reaction of the dispersion of other MXene types is required to be fully understood, and hence, their stabilities are also ensured. (x) Fabrication of a large amount of MXene nanomaterials99 as well as scalable film synthesis389 are required in order to use MXene nanomaterials in membrane or adsorption technologies, which bring the investigation of etching and delaminating processes for the large quantities. (xi) In addition to the lab-scale measurements, the performance of MXene membranes and adsorbents should be conducted in a real industrial process390 with a real feed stream composition371 and under practical conditions. For each separation application, that kind of investigation is very rare. (xii) For the nanomedicine and biomedical applications, mostly the cytotoxicity and biocompatibility tests of MXene to the cancer and healthy cells were performed.391 Unfortunately, these attempts are still in the beginning stage. Therefore, systematic investigation of toxic behavior of MXene nanomaterials and their impact on the environment and human beings should be carried out.

Acknowledgments

Ş.M. and S.V. would like to thank the Scientific and Technological Research Council of Turkey for funding support under Career Development Program (TUBITAK, Grant No. 120M180)

Supporting Information Available

The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.3c01182.

  • Survey of gas adsorption capacity of MXene adsorbents, solvent transport performance of MXene-based OSN and pervaporation membranes, dye removal performances of MXene-based membranes, performance of MXene-based adsorbents for heavy metal ion and radionuclides removal, and CDI desalination performance of MXene electrodes (PDF)

The authors declare no competing financial interest.

Supplementary Material

ao3c01182_si_001.pdf (233.2KB, pdf)

References

  1. Appannagari R. R. Environmental pollution causes and consequences: a study. North Asian Inter. Res. J. Social Science Humanities 2017, 3 (8), 151–161. [Google Scholar]
  2. Çakır G.; Coşkun E.. The importance of using renewable energy resources for ecological balance. Ecology Symposium 2012.
  3. Huang Y.; Su W.; Wang R.; Zhao T. Removal of typical industrial gaseous pollutants: From carbon, zeolite, and metal-organic frameworks to molecularly imprinted adsorbents. Aerosol Air Qual. Res. 2019, 19 (9), 2130–2150. 10.4209/aaqr.2019.04.0215. [DOI] [Google Scholar]
  4. Bozoǧlu B.What Turkey needs to do about the early warning system within the scope of the Paris climate agreement. Phd Thesis, Ankara University, Institue of Social Sciences, 2018. [Google Scholar]
  5. Rodríguez-Eugenio N.; McLaughlin M.; Pennock D.. Soil pollution: A hidden reality. FAO: Rome, 2018. [Google Scholar]
  6. World Health Organization: African Region, https://www.afro.who.int/health-topics/water (accessed 2022-11-29). [Google Scholar]
  7. Stern N.; Stern N. H.. The economics of climate change: the Stern review; Cambridge University Press, 2007. [Google Scholar]
  8. Fuller R.; Landrigan P. J; Balakrishnan K.; Bathan G.; Bose-O'Reilly S.; Brauer M.; Caravanos J.; Chiles T.; Cohen A.; Corra L.; Cropper M.; Ferraro G.; Hanna J.; Hanrahan D.; Hu H.; Hunter D.; Janata G.; Kupka R.; Lanphear B.; Lichtveld M.; Martin K.; Mustapha A.; Sanchez-Triana E.; Sandilya K.; Schaefli L.; Shaw J.; Seddon J.; Suk W.; Tellez-Rojo M. M.; Yan C. Pollution and health: A progress update. Lancet Planet. Health 2022, 6, e535 10.1016/S2542-5196(22)00090-0. [DOI] [PubMed] [Google Scholar]
  9. Briggs D. Environmental pollution and the global burden of disease. Br. Med. Bulletin 2003, 68 (1), 1–24. 10.1093/bmb/ldg019. [DOI] [PubMed] [Google Scholar]
  10. Saravanan A.; Senthil Kumar P.; Jeevanantham S.; Karishma S.; Tajsabreen B.; Yaashikaa P.R.; Reshma B. Effective water/wastewater treatment methodologies for toxic pollutants removal: Processes and applications towards sustainable development. Chemosphere 2021, 280, 130595. 10.1016/j.chemosphere.2021.130595. [DOI] [PubMed] [Google Scholar]
  11. Demir H.; Aksu G. O.; Gulbalkan H. C.; Keskin S. MOF membranes for CO2 capture: Past, present and future. Carbon Capture Sci. Technol. 2022, 2, 100026. 10.1016/j.ccst.2021.100026. [DOI] [Google Scholar]
  12. Liu R. S.; Shi X. D.; Wang C. T.; Gao Y. Z.; Xu S.; Hao G. P.; Chen S.; Lu A. H. Advances in post-combustion CO2 capture by physical adsorption: From materials innovation to separation practice. ChemSusChem 2021, 14 (6), 1428–1471. 10.1002/cssc.202002677. [DOI] [PubMed] [Google Scholar]
  13. Naquash A.; Qyyum M. A.; Chaniago Y. D.; Riaz A.; Yehia F.; Lim H.; Lee M. Separation and purification of syngas-derived hydrogen: A comparative evaluation of membrane-and cryogenic-assisted approaches. Chemosphere 2023, 313, 137420. 10.1016/j.chemosphere.2022.137420. [DOI] [PubMed] [Google Scholar]
  14. Székely G.; Gil M.; Sellergren B.; Heggie W.; Ferreira F. C. Environmental and economic analysis for selection and engineering sustainable API degenotoxification processes. Green Chem. 2013, 15 (1), 210–225. 10.1039/C2GC36239B. [DOI] [Google Scholar]
  15. Bolisetty S.; Peydayesh M.; Mezzenga R. Sustainable technologies for water purification from heavy metals: Review and analysis. Chem. Soc. Rev. 2019, 48 (2), 463–487. 10.1039/C8CS00493E. [DOI] [PubMed] [Google Scholar]
  16. Fortune Business Insights. https://www.fortunebusinessinsights.com/membranes-market-102982 (accessed 2022-12-5).
  17. Beuscher U.; Kappert E.; Wijmans J. Membrane research beyond materials science. J. Membr. Sci. 2022, 643, 119902. 10.1016/j.memsci.2021.119902. [DOI] [Google Scholar]
  18. Cheng L.; Liu G.; Zhao J.; Jin W. Two-dimensional-material membranes: manipulating the transport pathway for molecular separation. Acc.of Mat. Res. 2021, 2 (2), 114–128. 10.1021/accountsmr.0c00092. [DOI] [Google Scholar]
  19. Lei J.-C.; Zhang X.; Zhou Z. Recent advances in MXene: Preparation, properties, and applications. Front. Phys. 2015, 10 (3), 276–286. 10.1007/s11467-015-0493-x. [DOI] [Google Scholar]
  20. Naguib M.; Barsoum M. W.; Gogotsi Y. J. A. M. Ten years of progress in the synthesis and development of MXenes. Adv. Mater. 2021, 33 (39), 2103393. 10.1002/adma.202103393. [DOI] [PubMed] [Google Scholar]
  21. Naguib M.; Kurtoglu M.; Presser V.; Lu J.; Niu J.; Heon M.; Hultman L.; Gogotsi Y.; Barsoum M. W. Two-dimensional nanocrystals produced by exfoliation of Ti3AlC2. Adv. Mater. 2011, 23 (37), 4248–4253. 10.1002/adma.201102306. [DOI] [PubMed] [Google Scholar]
  22. Naguib M.; Mashtalir O.; Carle J.; Presser V.; Lu J.; Hultman L.; Gogotsi Y.; Barsoum M. W. Two-dimensional transition metal carbides. ACS Nano 2012, 6 (2), 1322–1331. 10.1021/nn204153h. [DOI] [PubMed] [Google Scholar]
  23. Alhabeb M.; Maleski K.; Anasori B.; Lelyukh P.; Clark L.; Sin S.; Gogotsi Y. Guidelines for synthesis and processing of two-dimensional titanium carbide (Ti3C2Tx MXene). Chem. Mater. 2017, 29 (18), 7633–7644. 10.1021/acs.chemmater.7b02847. [DOI] [Google Scholar]
  24. Urbankowski P.; Anasori B.; Makaryan T.; Er D.; Kota S.; Walsh P. L.; Zhao M.; Shenoy V. B.; Barsoum M. W.; Gogotsi Y. Synthesis of two-dimensional titanium nitride Ti4N3 (MXene). Nanoscale 2016, 8 (22), 11385–11391. 10.1039/C6NR02253G. [DOI] [PubMed] [Google Scholar]
  25. Li Y.; Shao H.; Lin Z.; Lu J.; Liu L.; Duployer B.; Persson P. O. A.; Eklund P.; Hultman L.; Li M.; Chen K.; Zha X.-H.; Du S.; Rozier P.; Chai Z.; Raymundo-Pinero E.; Taberna P.-L.; Simon P.; Huang Q. A general Lewis acidic etching route for preparing MXenes with enhanced electrochemical performance in non-aqueous electrolyte. Nat. Mater. 2020, 19 (8), 894–899. 10.1038/s41563-020-0657-0. [DOI] [PubMed] [Google Scholar]
  26. Yang S.; Zhang P.; Wang F.; Ricciardulli A. G.; Lohe M. R.; Blom P. W.; Feng X. Fluoride-free synthesis of two-dimensional titanium carbide (MXene) using a binary aqueous system. Angew. Chem. 2018, 130 (47), 15717–15721. 10.1002/ange.201809662. [DOI] [PubMed] [Google Scholar]
  27. Jeon M.; Jun B.-M.; Kim S.; Jang M.; Park C. M.; Snyder S. A.; Yoon Y. A review on MXene-based nanomaterials as adsorbents in aqueous solution. Chemosphere 2020, 261, 127781. 10.1016/j.chemosphere.2020.127781. [DOI] [PubMed] [Google Scholar]
  28. Firestein K. L.; von Treifeldt J. E.; Kvashnin D. G.; Fernando J. F. S.; Zhang C.; Kvashnin A. G.; Podryabinkin E. V.; Shapeev A. V.; Siriwardena D. P.; Sorokin P. B.; Golberg D. Young’s modulus and tensile strength of Ti3C2 MXene nanosheets as revealed by in situ TEM probing, AFM nanomechanical mapping, and theoretical calculations. Nano Lett. 2020, 20 (8), 5900–5908. 10.1021/acs.nanolett.0c01861. [DOI] [PubMed] [Google Scholar]
  29. Borysiuk V. N.; Mochalin V. N.; Gogotsi Y. Molecular dynamic study of the mechanical properties of two-dimensional titanium carbides Tin+1Cn (MXenes). Nanotechnology 2015, 26 (26), 265705. 10.1088/0957-4484/26/26/265705. [DOI] [PubMed] [Google Scholar]
  30. Feng L.; Zha X.-H.; Luo K.; Huang Q.; He J.; Liu Y.; Deng W.; Du S. Structures and mechanical and electronic properties of the Ti2CO2 MXene incorporated with neighboring elements (Sc, V, B and N). J. Electron. Mater. 2017, 46 (4), 2460–2466. 10.1007/s11664-017-5311-5. [DOI] [Google Scholar]
  31. Li Y.; Huang S.; Wei C.; Zhou D.; Li B.; Wu C.; Mochalin V. N. Adhesion between MXenes and other 2D materials. ACS Appl. Mater. Interfaces 2021, 13 (3), 4682–4691. 10.1021/acsami.0c18624. [DOI] [PubMed] [Google Scholar]
  32. Khazaei M.; Arai M.; Sasaki T.; Chung C. Y.; Venkataramanan N. S.; Estili M.; Sakka Y.; Kawazoe Y. Novel electronic and magnetic properties of two-dimensional transition metal carbides and nitrides. Adv. Funct. Mater. 2013, 23 (17), 2185–2192. 10.1002/adfm.201202502. [DOI] [Google Scholar]
  33. Xie Y.; Kent P. Hybrid density functional study of structural and electronic properties of functionalized Tin+1Xn (X= C, N) monolayers. Phys. Rev. B 2013, 87 (23), 235441. 10.1103/PhysRevB.87.235441. [DOI] [Google Scholar]
  34. Gao G.; Ding G.; Li J.; Yao K.; Wu M.; Qian M. Monolayer MXenes: Promising half-metals and spin gapless semiconductors. Nanoscale 2016, 8 (16), 8986–8994. 10.1039/C6NR01333C. [DOI] [PubMed] [Google Scholar]
  35. Khazaei M.; Ranjbar A.; Ghorbani-Asl M.; Arai M.; Sasaki T.; Liang Y.; Yunoki S. Nearly free electron states in MXenes. Phys. Rev. B 2016, 93 (20), 205125. 10.1103/PhysRevB.93.205125. [DOI] [Google Scholar]
  36. Enyashin A.; Ivanovskii A. Atomic structure, comparative stability and electronic properties of hydroxylated Ti2C and Ti3C2 nanotubes. Comp. Theor. Chem. 2012, 989, 27–32. 10.1016/j.comptc.2012.02.034. [DOI] [Google Scholar]
  37. Zha X.-H.; Luo K.; Li Q.; Huang Q.; He J.; Wen X.; Du S. Role of the surface effect on the structural, electronic and mechanical properties of the carbide MXenes. EPL 2015, 111 (2), 26007. 10.1209/0295-5075/111/26007. [DOI] [Google Scholar]
  38. Papadopoulou K. A.; Chroneos A.; Parfitt D.; Christopoulos S.-R. G. A perspective on MXenes: Their synthesis, properties, and recent applications. J. Appl. Phys. 2020, 128 (17), 170902. 10.1063/5.0021485. [DOI] [Google Scholar]
  39. Zhang N.; Hong Y.; Yazdanparast S.; Asle Zaeem M. Superior structural, elastic and electronic properties of 2D titanium nitride MXenes over carbide MXenes: a comprehensive first principles study. 2D Materials 2018, 5 (4), 045004. 10.1088/2053-1583/aacfb3. [DOI] [Google Scholar]
  40. Wang K.; Zhou Y.; Xu W.; Huang D.; Wang Z.; Hong M. Fabrication and thermal stability of two-dimensional carbide Ti3C2 nanosheets. Ceram. Int. 2016, 42 (7), 8419–8424. 10.1016/j.ceramint.2016.02.059. [DOI] [Google Scholar]
  41. Zhou J.; Zha X.; Chen F. Y.; Ye Q.; Eklund P.; Du S.; Huang Q. A two-dimensional zirconium carbide by selective etching of Al3C3 from nanolaminated Zr3Al3C5. Angew. Chem., Int. Ed. 2016, 55 (16), 5008–5013. 10.1002/anie.201510432. [DOI] [PubMed] [Google Scholar]
  42. Han M.; Maleski K.; Shuck C. E.; Yang Y.; Glazar J. T.; Foucher A. C.; Hantanasirisakul K.; Sarycheva A.; Frey N. C.; May S. J.; Shenoy V. B.; Stach E. A.; Gogotsi Y. Tailoring electronic and optical properties of MXenes through forming solid solutions. J. Am. Chem. Soc. 2020, 142 (45), 19110–19118. 10.1021/jacs.0c07395. [DOI] [PubMed] [Google Scholar]
  43. Halim J.; Persson I.; Moon E. J.; Kühne P.; Darakchieva V.; Persson P. O. Å.; Eklund P.; Rosen J.; Barsoum M. W. Electronic and optical characterization of 2D Ti2C and Nb2C (MXene) thin films. J. Phys.: Condens. Matter 2019, 31 (16), 165301. 10.1088/1361-648X/ab00a2. [DOI] [PubMed] [Google Scholar]
  44. Han M.; Shuck C. E.; Rakhmanov R.; Parchment D.; Anasori B.; Koo C. M.; Friedman G.; Gogotsi Y. Beyond Ti3C2Tx: MXenes for electromagnetic interference shielding. ACS Nano 2020, 14 (4), 5008–5016. 10.1021/acsnano.0c01312. [DOI] [PubMed] [Google Scholar]
  45. Gao Q.; Zhang H. Magnetic i-MXenes: a new class of multifunctional two-dimensional materials. Nanoscale 2020, 12 (10), 5995–6001. 10.1039/C9NR10181K. [DOI] [PubMed] [Google Scholar]
  46. Pandey R. P.; Rasheed P. A.; Gomez T.; Rasool K.; Ponraj J.; Prenger K.; Naguib M.; Mahmoud K. A. Effect of sheet size and atomic structure on the antibacterial activity of Nb-MXene nanosheets. ACS Appl.Nano Mater. 2020, 3 (11), 11372–11382. 10.1021/acsanm.0c02463. [DOI] [Google Scholar]
  47. Naguib M.; Mashtalir O.; Lukatskaya M. R.; Dyatkin B.; Zhang C.; Presser V.; Gogotsi Y.; Barsoum M. W. One-step synthesis of nanocrystalline transition metal oxides on thin sheets of disordered graphitic carbon by oxidation of MXenes. Chem. Commun. 2014, 50 (56), 7420–7423. 10.1039/C4CC01646G. [DOI] [PubMed] [Google Scholar]
  48. Lotfi R.; Naguib M.; Yilmaz D. E.; Nanda J.; Van Duin A. C. A comparative study on the oxidation of two-dimensional Ti3C2 MXene structures in different environments. J. Mater. Chem. A 2018, 6 (26), 12733–12743. 10.1039/C8TA01468J. [DOI] [Google Scholar]
  49. Iqbal A.; Hong J.; Ko T. Y.; Koo C. M. Improving oxidation stability of 2D MXenes: Synthesis, storage media, and conditions. Nano Convergence 2021, 8 (1), 1–22. 10.1186/s40580-021-00259-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Mathis T. S.; Maleski K.; Goad A.; Sarycheva A.; Anayee M.; Foucher A. C.; Hantanasirisakul K.; Shuck C. E.; Stach E. A.; Gogotsi Y. Modified MAX phase synthesis for environmentally stable and highly conductive Ti3C2 MXene. ACS Nano 2021, 15 (4), 6420–6429. 10.1021/acsnano.0c08357. [DOI] [PubMed] [Google Scholar]
  51. Natu V.; Hart J. L.; Sokol M.; Chiang H.; Taheri M. L.; Barsoum M. W. Edge capping of 2D-MXene sheets with polyanionic salts to mitigate oxidation in aqueous colloidal suspensions. Angew. Chem. 2019, 131 (36), 12785–12790. 10.1002/ange.201906138. [DOI] [PubMed] [Google Scholar]
  52. Athavale S.; Micci-Barreca S.; Arole K.; Kotasthane V.; Blivin J.; Cao H.; Lutkenhaus J. L.; Radovic M.; Green M. J. Advances in the Chemical Stabilization of MXenes. Langmuir 2023, 39, 918. 10.1021/acs.langmuir.2c02051. [DOI] [PubMed] [Google Scholar]
  53. Zhang C. J.; Pinilla S.; McEvoy N.; Cullen C. P.; Anasori B.; Long E.; Park S.-H.; Seral-Ascaso A.; Shmeliov A.; Krishnan D.; Morant C.; Liu X.; Duesberg G. S.; Gogotsi Y.; Nicolosi V. Oxidation stability of colloidal two-dimensional titanium carbides (MXenes). Chem. Mater. 2017, 29 (11), 4848–4856. 10.1021/acs.chemmater.7b00745. [DOI] [Google Scholar]
  54. Zhao H.; Ding J.; Zhou M.; Yu H. Air-stable titanium carbide MXene nanosheets for corrosion protection. ACS Appl. Nano Mater. 2021, 4 (3), 3075–3086. 10.1021/acsanm.1c00219. [DOI] [Google Scholar]
  55. Lu M.; Han W.; Li H.; Zhang W.; Zhang B. There is plenty of space in the MXene layers: The confinement and fillings. J. Energy Chem. 2020, 48, 344–363. 10.1016/j.jechem.2020.02.032. [DOI] [Google Scholar]
  56. Munir S.; Rasheed A.; Rasheed T.; Ayman I.; Ajmal S.; Rehman A.; Shakir I.; Agboola P. O.; Warsi M. F. Exploring the influence of critical parameters for the effective synthesis of high-quality 2D MXene. ACS Omega 2020, 5 (41), 26845–26854. 10.1021/acsomega.0c03970. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Jastrzȩbska A.; Szuplewska A.; Wojciechowski T.; Chudy M.; Ziemkowska W.; Chlubny L.; Rozmysłowska A.; Olszyna A. In vitro studies on cytotoxicity of delaminated Ti3C2 MXene. J. Hazard. Mater. 2017, 339, 1–8. 10.1016/j.jhazmat.2017.06.004. [DOI] [PubMed] [Google Scholar]
  58. Lim G. P.; Soon C. F.; Ma N. L.; Morsin M.; Nayan N.; Ahmad M. K.; Tee K. S. Cytotoxicity of MXene-based nanomaterials for biomedical applications: A mini review. Environ. Res. 2021, 201, 111592. 10.1016/j.envres.2021.111592. [DOI] [PubMed] [Google Scholar]
  59. Ghidiu M.; Lukatskaya M. R.; Zhao M.-Q.; Gogotsi Y.; Barsoum M. W. Conductive two-dimensional titanium carbide ‘clay’ with high volumetric capacitance. Nature 2014, 516 (7529), 78–81. 10.1038/nature13970. [DOI] [PubMed] [Google Scholar]
  60. Li T.; Yao L.; Liu Q.; Gu J.; Luo R.; Li J.; Yan X.; Wang W.; Liu P.; Chen B.; Zhang W.; Abbas W.; Naz R.; Zhang D. Fluorine-free synthesis of high-purity Ti3C2Tx (T= OH, O) via alkali treatment. Angew. Chem., Int. Ed. 2018, 57 (21), 6115–6119. 10.1002/anie.201800887. [DOI] [PubMed] [Google Scholar]
  61. Alhabeb M.; Maleski K.; Mathis T. S.; Sarycheva A.; Hatter C. B.; Uzun S.; Levitt A.; Gogotsi Y. Selective etching of silicon from Ti3SiC2 (MAX) to obtain 2D titanium carbide (MXene). Angew. Chem. 2018, 130 (19), 5542–5546. 10.1002/ange.201802232. [DOI] [PubMed] [Google Scholar]
  62. 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 (6506), 979–983. 10.1126/science.aba8311. [DOI] [PubMed] [Google Scholar]
  63. Mashtalir O.; Naguib M.; Mochalin V. N.; Dall’Agnese Y.; Heon M.; Barsoum M. W.; Gogotsi Y. Intercalation and delamination of layered carbides and carbonitrides. Nat. Commun. 2013, 4 (1), 1–7. 10.1038/ncomms2664. [DOI] [PubMed] [Google Scholar]
  64. Mashtalir O.; Lukatskaya M. R.; Zhao M. Q.; Barsoum M. W.; Gogotsi Y. Amine-assisted delamination of Nb2C MXene for Li-ion energy storage devices. Adv. Mater. 2015, 27 (23), 3501–3506. 10.1002/adma.201500604. [DOI] [PubMed] [Google Scholar]
  65. Naguib M.; Unocic R. R.; Armstrong B. L.; Nanda J. Large-scale delamination of multi-layers transition metal carbides and carbonitrides MXenes. Dalton Trans. 2015, 44 (20), 9353–9358. 10.1039/C5DT01247C. [DOI] [PubMed] [Google Scholar]
  66. 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 (12), 1600255. 10.1002/aelm.201600255. [DOI] [Google Scholar]
  67. Tao Q.; Dahlqvist M.; Lu J.; Kota S.; Meshkian R.; Halim J.; Palisaitis J.; Hultman L.; Barsoum M. W.; Persson P. O.A.; Rosen J. Two-dimensional Mo1. 33C MXene with divacancy ordering prepared from parent 3D laminate with in-plane chemical ordering. Nat. Commun. 2017, 8 (1), 1–7. 10.1038/ncomms14949. [DOI] [PMC free article] [PubMed] [Google Scholar]
  68. Mei J.; Ayoko G. A.; Hu C.; Bell J. M.; Sun Z. Two-dimensional fluorine-free mesoporous Mo2C MXene via UV-induced selective etching of Mo2Ga2C for energy storage. Sustainable Mater. Technol. 2020, 25, e00156 10.1016/j.susmat.2020.e00156. [DOI] [Google Scholar]
  69. Xue N.; Li X.; Zhang M.; Han L.; Liu Y.; Tao X. Chemical-combined ball-milling synthesis of fluorine-free porous MXene for high-performance lithium ion batteries. ACS Appl. Energy Mater. 2020, 3 (10), 10234–10241. 10.1021/acsaem.0c02081. [DOI] [Google Scholar]
  70. Abdolhosseinzadeh S.; Jiang X.; Zhang H.; Qiu J.; Zhang C. J. Perspectives on solution processing of two-dimensional MXenes. Mater. Today 2021, 48, 214–240. 10.1016/j.mattod.2021.02.010. [DOI] [Google Scholar]
  71. Vural M.; Pena-Francesch A.; Bars-Pomes J.; Jung H.; Gudapati H.; Hatter C. B.; Allen B. D.; Anasori B.; Ozbolat I. T.; Gogotsi Y.; Demirel M. C. Inkjet printing of self-assembled 2D titanium carbide and protein electrodes for stimuli-responsive electromagnetic shielding. Adv. Funct. Mater. 2018, 28 (32), 1801972. 10.1002/adfm.201801972. [DOI] [Google Scholar]
  72. Abdolhosseinzadeh S.; Schneider R.; Verma A.; Heier J.; Nüesch F.; Zhang C. Turning trash into treasure: additive free MXene sediment inks for screen-printed micro-supercapacitors. Adv. Mater. 2020, 32 (17), 2000716. 10.1002/adma.202000716. [DOI] [PubMed] [Google Scholar]
  73. Cao W. T.; Ma C.; Mao D. S.; Zhang J.; Ma M. G.; Chen F. MXene-reinforced cellulose nanofibril inks for 3D-printed smart fibres and textiles. Adv. Funct. Mater. 2019, 29 (51), 1905898. 10.1002/adfm.201905898. [DOI] [Google Scholar]
  74. Salles P.; Quain E.; Kurra N.; Sarycheva A.; Gogotsi Y. Automated scalpel patterning of solution processed thin films for fabrication of transparent MXene microsupercapacitors. Small 2018, 14 (44), 1802864. 10.1002/smll.201802864. [DOI] [PubMed] [Google Scholar]
  75. Zhang C.; Anasori B.; Seral-Ascaso A.; Park S. H.; McEvoy N.; Shmeliov A.; Duesberg G. S.; Coleman J. N.; Gogotsi Y.; Nicolosi V. Transparent, flexible, and conductive 2D titanium carbide (MXene) films with high volumetric capacitance. Adv. Mater. 2017, 29 (36), 1702678. 10.1002/adma.201702678. [DOI] [PubMed] [Google Scholar]
  76. Zhang J.; Kong N.; Uzun S.; Levitt A.; Seyedin S.; Lynch P. A.; Qin S.; Han M.; Yang W.; Liu J.; Wang X.; Gogotsi Y.; Razal J. M. Scalable manufacturing of free-standing, strong Ti3C2Tx MXene films with outstanding conductivity. Adv. Mater. 2020, 32 (23), 2001093. 10.1002/adma.202001093. [DOI] [PubMed] [Google Scholar]
  77. Ali I.; Faraz Ud Din M.; Gu Z.-G. MXenes thin films: From fabrication to their applications. Molecules 2022, 27 (15), 4925. 10.3390/molecules27154925. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Deng J.; Lu Z.; Ding L.; Li Z.-K.; Wei Y.; Caro J.; Wang H. Fast electrophoretic preparation of large-area two-dimensional titanium carbide membranes for ion sieving. Chem. Eng. J. 2021, 408, 127806. 10.1016/j.cej.2020.127806. [DOI] [Google Scholar]
  79. Xu J.; You J.; Wang L.; Wang Z.; Zhang H. MXenes serving aqueous supercapacitors: Preparation, energy storage mechanism and electrochemical performance enhancement. Sustainable Mater. Technol. 2022, 33, e00490 10.1016/j.susmat.2022.e00490. [DOI] [Google Scholar]
  80. Anasori B.; Lukatskaya M. R.; Gogotsi Y. 2D metal carbides and nitrides (MXenes) for energy storage. Nature Rev. Mater. 2017, 2 (2), 1–17. 10.1038/natrevmats.2016.98. [DOI] [Google Scholar]
  81. Hu M.; Cheng R.; Li Z.; Hu T.; Zhang H.; Shi C.; Yang J.; Cui C.; Zhang C.; Wang H.; Fan B.; Wang X.; Yang Q.-H. Interlayer engineering of Ti3C2Tx MXenes towards high capacitance supercapacitors. Nanoscale 2020, 12 (2), 763–771. 10.1039/C9NR08960H. [DOI] [PubMed] [Google Scholar]
  82. Hu Q.; Wang H.; Wu Q.; Ye X.; Zhou A.; Sun D.; Wang L.; Liu B.; He J. Two-dimensional Sc2C: A reversible and high-capacity hydrogen storage material predicted by first-principles calculations. Inter. J. Hydrogen Energy 2014, 39 (20), 10606–10612. 10.1016/j.ijhydene.2014.05.037. [DOI] [Google Scholar]
  83. Gouveia J. D.; Morales-García A.; Vines F.; Illas F.; Gomes J. R. MXenes as promising catalysts for water dissociation. Appl. Catal. B: Environ. 2020, 260, 118191. 10.1016/j.apcatb.2019.118191. [DOI] [Google Scholar]
  84. Fu B.; Sun J.; Wang C.; Shang C.; Xu L.; Li J.; Zhang H. MXenes: Synthesis, optical properties, and applications in ultrafast photonics. Small 2021, 17 (11), 2006054. 10.1002/smll.202006054. [DOI] [PubMed] [Google Scholar]
  85. Ihsanullah I. Potential of MXenes in water desalination: current status and perspectives. Nano-Micro Lett. 2020, 12 (1), 1–20. 10.1007/s40820-020-0411-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Dixit F.; Zimmermann K.; Dutta R.; Prakash N. J.; Barbeau B.; Mohseni M.; Kandasubramanian B. Application of MXenes for water treatment and energy-efficient desalination: A review. J. Hazard. Mater. 2022, 423, 127050. 10.1016/j.jhazmat.2021.127050. [DOI] [PubMed] [Google Scholar]
  87. Ahmed Z.; Rehman F.; Ali U.; Ali A.; Iqbal M.; Thebo K. H. Recent advances in MXene-based separation membranes. ChemBioEng. Reviews 2021, 8 (2), 110–120. 10.1002/cben.202000026. [DOI] [Google Scholar]
  88. Huang L.; Ding L.; Wang H. MXene-Based Membranes for Separation Applications. Small Science 2021, 1 (7), 2100013. 10.1002/smsc.202100013. [DOI] [Google Scholar]
  89. Wu Y.; Li X.; Zhao H.; Yao F.; Cao J.; Chen Z.; Huang X.; Wang D.; Yang Q. Recent advances in transition metal carbides and nitrides (MXenes): Characteristics, environmental remediation and challenges. Chem. Eng. J. 2021, 418, 129296. 10.1016/j.cej.2021.129296. [DOI] [Google Scholar]
  90. Kwon O.; Choi Y.; Kang J.; Kim J. H.; Choi E.; Woo Y. C.; Kim D. W. A comprehensive review of MXene-based water-treatment membranes and technologies: Recent progress and perspectives. Desalination 2022, 522, 115448. 10.1016/j.desal.2021.115448. [DOI] [Google Scholar]
  91. Ibrahim Y.; Meslam M.; Eid K.; Salah B.; Abdullah A. M.; Ozoemena K. I.; Elzatahry A.; Sharaf M. A.; Sillanpää M. A review of MXenes as emergent materials for dye removal from wastewater. Sep. Purif. Technol. 2022, 282, 120083. 10.1016/j.seppur.2021.120083. [DOI] [Google Scholar]
  92. Rasheed T.; Kausar F.; Rizwan K.; Adeel M.; Sher F.; Alwadai N.; Alshammari F. H. Two dimensional MXenes as emerging paradigm for adsorptive removal of toxic metallic pollutants from wastewater. Chemosphere 2022, 287, 132319. 10.1016/j.chemosphere.2021.132319. [DOI] [PubMed] [Google Scholar]
  93. Hwang S. K.; Kang S.-M.; Rethinasabapathy M.; Roh C.; Huh Y. S. MXene: An emerging two-dimensional layered material for removal of radioactive pollutants. Chem. Eng. J. 2020, 397, 125428. 10.1016/j.cej.2020.125428. [DOI] [Google Scholar]
  94. Dong X.; Bhattacharjee S.; Zhang C.; Chai F. Titanium carbide-based adsorbents for removal of heavy metal ions and radionuclides: From nanomaterials to 3D architectures. Adv. Mater. Inter. 2021, 8 (21), 2100703. 10.1002/admi.202100703. [DOI] [Google Scholar]
  95. Szuplewska A.; Kulpińska D.; Jakubczak M.; Dybko A.; Chudy M.; Olszyna A.; Brzózka Z.; Jastrzȩbska A. M. The 10th anniversary of MXenes: Challenges and prospects for their surface modification toward future biotechnological applications. Adv. Drug Delivery Rev. 2022, 182, 114099. 10.1016/j.addr.2021.114099. [DOI] [PubMed] [Google Scholar]
  96. Huang H.; Dong C.; Feng W.; Wang Y.; Huang B.; Chen Y. Biomedical engineering of two-dimensional MXenes. Adv. Drug Delivery Rev. 2022, 184, 114178. 10.1016/j.addr.2022.114178. [DOI] [PubMed] [Google Scholar]
  97. Ding G.; Yang B.; Chen R.-S.; Zhou K.; Han S.-T.; Zhou Y. MXenes for memristive and tactile sensory systems. Appl. Phys. Reviews 2021, 8 (1), 011316. 10.1063/5.0026093. [DOI] [Google Scholar]
  98. Shuck C. E.; Sarycheva A.; Anayee M.; Levitt A.; Zhu Y.; Uzun S.; Balitskiy V.; Zahorodna V.; Gogotsi O.; Gogotsi Y. Scalable synthesis of Ti3C2Tx mxene. Adv. Eng. Mater. 2020, 22 (3), 1901241. 10.1002/adem.201901241. [DOI] [Google Scholar]
  99. Chen N.; Duan Z.; Cai W.; Wang Y.; Pu B.; Huang H.; Xie Y.; Tang Q.; Zhang H.; Yang W. Supercritical etching method for the large-scale manufacturing of MXenes. Nano Energy 2023, 107, 108147. 10.1016/j.nanoen.2022.108147. [DOI] [Google Scholar]
  100. Khosla A.; Sonu; Awan H. T. A.; Singh K.; Gaurav; Walvekar R.; Zhao Z.; Kaushik A.; Khalid M.; Chaudhary V. Emergence of MXene and MXene-polymer hybrid membranes as future-environmental remediation strategies. Advanced Science 2022, 9, 2203527. 10.1002/advs.202203527. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Lim G. P.; Soon C. F.; Al-Gheethi A.; Morsin M.; Tee K. S. Recent progress and new perspective of MXene-based membranes for water purification: A review. Ceram. Int. 2022, 48, 16477. 10.1016/j.ceramint.2022.03.165. [DOI] [Google Scholar]
  102. Lin Q.; Liu Y.; Yang Z.; He Z.; Wang H.; Zhang L.; Ang M. B. M. Y.; Zeng G. Construction and application of two-dimensional MXene-based membranes for water treatment: A mini-review. Results in Eng. 2022, 15, 100494. 10.1016/j.rineng.2022.100494. [DOI] [Google Scholar]
  103. Ahmaruzzaman M. MXenes and MXene-supported nanocomposites: a novel materials for aqueous environmental remediation. RSC Adv. 2022, 12 (53), 34766–34789. 10.1039/D2RA05530A. [DOI] [PMC free article] [PubMed] [Google Scholar]
  104. Ahmaruzzaman M. M-xenes and mxene based nanocomposites: a new generation potential materials for removal of organic contaminants from water. Inter. J. Environ. Analy. Chem. 2022, 1–17. 10.1080/03067319.2022.2076226. [DOI] [Google Scholar]
  105. Jena R. K.; Das H. T.; Patra B. N.; Das N. MXene-based nanomaterials as adsorbents for wastewater treatment: a review on recent trends. Frontiers of Mater. Sci. 2022, 16 (1), 220592. 10.1007/s11706-022-0592-x. [DOI] [Google Scholar]
  106. Sun Y.; Li Y. Potential environmental applications of MXenes: A critical review. Chemosphere 2021, 271, 129578. 10.1016/j.chemosphere.2021.129578. [DOI] [PubMed] [Google Scholar]
  107. Yaqub A.; Shafiq Q.; Khan A. R.; Husnain S. M.; Shahzad F. Recent advances in the adsorptive remediation of wastewater using two-dimensional transition metal carbides (MXenes): a review. New J. Chem. 2021, 45 (22), 9721–9742. 10.1039/D1NJ00772F. [DOI] [Google Scholar]
  108. Yu F.; Pan J.; Zhang X.; Bai X.; Ma J. Adsorption of contaminants from aqueous solutions by modified biochar: A review. Environ. Chem. 2022, 19, 53. 10.1071/EN22014. [DOI] [Google Scholar]
  109. Castro-Muñoz R. MXene: A two-dimensional material in selective water separation via pervaporation. Arabian J. Chem. 2022, 15 (1), 103524. 10.1016/j.arabjc.2021.103524. [DOI] [Google Scholar]
  110. Chen Y.; Liu C.; Guo S.; Mu T.; Wei L.; Lu Y. CO2 capture and conversion to value-added products promoted by MXene-based materials. Green Energy Environ. 2022, 7 (3), 394–410. 10.1016/j.gee.2020.11.008. [DOI] [Google Scholar]
  111. Dixit F.; Zimmermann K.; Alamoudi M.; Abkar L.; Barbeau B.; Mohseni M.; Kandasubramanian B.; Smith K. Application of MXenes for air purification, gas separation and storage: A review. Renewable and Sustainable Energy Rev. 2022, 164, 112527. 10.1016/j.rser.2022.112527. [DOI] [Google Scholar]
  112. Isfahani A. P.; Arabi Shamsabadi A.; Soroush M. MXenes and other two-dimensional materials for membrane gas separation: Progress, challenges, and potential of MXene-based membranes. Ind. Eng. Chem. Res. 2023, 62 (5), 2309. 10.1021/acs.iecr.2c02042. [DOI] [Google Scholar]
  113. Mahar I.; Memon F. H.; Lee J.-W.; Kim K. H.; Ahmed R.; Soomro F.; Rehman F.; Memon A. A.; Thebo K. H.; Choi K. H. Two-dimensional transition metal carbides and nitrides (MXenes) for water purification and antibacterial applications. Membranes 2021, 11 (11), 869. 10.3390/membranes11110869. [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Wang L.; Xu Z.; Fu D.; Zhang G. Challenging MXene-based nanomaterials and composite membranes for water treatment. Curr. Anal. Chem. 2021, 17 (6), 731–736. 10.2174/1573411016666200123144242. [DOI] [Google Scholar]
  115. Chen L.; Wakeel M.; Ul Haq T.; Alharbi N. S.; Chen C.; Ren X. Recent progress in environmental remediation, colloidal behavior and biological effects of MXene: A review. Environ. Sci.: Nano 2022, 9, 3168. 10.1039/D2EN00340F. [DOI] [Google Scholar]
  116. Ihsanullah I.; Bilal M. Recent advances in the development of MXene-based membranes for oil/water separation: A critical review. Appli. Mater. Today 2022, 29, 101674. 10.1016/j.apmt.2022.101674. [DOI] [Google Scholar]
  117. Yanan H.; Jiahui H.; Anrui Z.; Ruoxuan G.; Kexin L.; Yuejie A. A Review on MXene and its applications in environmental remediation. Progress in Chem. 2022, 34 (5), 1229. 10.7536/PC210614. [DOI] [Google Scholar]
  118. Kumar J. A.; Prakash P.; Krithiga T.; Amarnath D. J.; Premkumar J.; Rajamohan N.; Vasseghian Y.; Saravanan P.; Rajasimman M. Methods of synthesis, characteristics, and environmental applications of MXene: A comprehensive review. Chemosphere 2022, 286, 131607. 10.1016/j.chemosphere.2021.131607. [DOI] [PubMed] [Google Scholar]
  119. Bilal M.; Ihsanullah I. What makes MXenes emergent materials for the adsorption of heavy metals from water? A critical review. J. Water Process Eng. 2022, 49, 103010. 10.1016/j.jwpe.2022.103010. [DOI] [Google Scholar]
  120. Sheth Y.; Dharaskar S.; Chaudhary V.; Khalid M.; Walvekar R. Prospects of titanium carbide-based MXene in heavy metal ion and radionuclide adsorption for wastewater remediation: A review. Chemosphere 2022, 293, 133563. 10.1016/j.chemosphere.2022.133563. [DOI] [PubMed] [Google Scholar]
  121. Qu K.; Huang K.; Xu Z. Recent progress in the design and fabrication of MXene-based membranes. Frontiers of Chem. Sci. and Eng. 2021, 15, 820–836. 10.1007/s11705-020-1997-7. [DOI] [Google Scholar]
  122. Wang Y.; Niu B.; Zhang X.; Lei Y.; Zhong P.; Ma X. Ti3C2Tx MXene: An emerging two-dimensional layered material in water treatment. ECS J. Solid State Sci. and Technol. 2021, 10 (4), 047002. 10.1149/2162-8777/abf2de. [DOI] [Google Scholar]
  123. Chen H.; Ma H.; Li C. Host-guest intercalation chemistry in MXenes and its implications for practical applications. ACS Nano 2021, 15 (10), 15502–15537. 10.1021/acsnano.1c04423. [DOI] [PubMed] [Google Scholar]
  124. Mozafari M.; Soroush M. Surface functionalization of MXenes. Mater. Adv. 2021, 2 (22), 7277–7307. 10.1039/D1MA00625H. [DOI] [Google Scholar]
  125. Ren C. E.; Hatzell K. B.; Alhabeb M.; Ling Z.; Mahmoud K. A.; Gogotsi Y. Charge-and size-selective ion sieving through Ti3C2Tx MXene membranes. J. Phys. Chem. Lett. 2015, 6 (20), 4026–4031. 10.1021/acs.jpclett.5b01895. [DOI] [PubMed] [Google Scholar]
  126. Wang L.; Yuan L.; Chen K.; Zhang Y.; Deng Q.; Du S.; Huang Q.; Zheng L.; Zhang J.; Chai Z.; Barsoum M. W.; Wang X.; Shi W. Loading actinides in multilayered structures for nuclear waste treatment: the first case study of uranium capture with vanadium carbide MXene. ACS Appl. Mater. Interfaces 2016, 8 (25), 16396–16403. 10.1021/acsami.6b02989. [DOI] [PubMed] [Google Scholar]
  127. Ding L.; Wei Y.; Wang Y.; Chen H.; Caro J.; Wang H. A two-dimensional lamellar membrane: MXene nanosheet stacks. Angew. Chem. 2017, 129 (7), 1851–1855. 10.1002/ange.201609306. [DOI] [PubMed] [Google Scholar]
  128. Wang J.; Chen P.; Shi B.; Guo W.; Jaroniec M.; Qiao S. Z. A regularly channeled lamellar membrane for unparalleled water and organics permeation. Angew. Chem., Int. Ed. 2018, 57 (23), 6814–6818. 10.1002/anie.201801094. [DOI] [PubMed] [Google Scholar]
  129. Ding L.; Wei Y.; Li L.; Zhang T.; Wang H.; Xue J.; Ding L.-X.; Wang S.; Caro J.; Gogotsi Y. MXene molecular sieving membranes for highly efficient gas separation. Nat. Commun. 2018, 9 (1), 1–7. 10.1038/s41467-017-02529-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
  130. Tan Y. Z.; Wang H.; Han L.; Tanis-Kanbur M. B.; Pranav M. V.; Chew J. W. Photothermal-enhanced and fouling-resistant membrane for solar-assisted membrane distillation. J. Membr. Sci. 2018, 565, 254–265. 10.1016/j.memsci.2018.08.032. [DOI] [Google Scholar]
  131. Karthikeyan P.; Elanchezhiyan S. S.; Preethi J.; Talukdar K.; Meenakshi S.; Park C. M. Two-dimensional (2D) Ti3C2Tx MXene nanosheets with superior adsorption behavior for phosphate and nitrate ions from the aqueous environment. Ceram. Int. 2021, 47 (1), 732–739. 10.1016/j.ceramint.2020.08.183. [DOI] [Google Scholar]
  132. Zhang Q.; Teng J.; Zou G.; Peng Q.; Du Q.; Jiao T.; Xiang J. Efficient phosphate sequestration for water purification by unique sandwich-like MXene/magnetic iron oxide nanocomposites. Nanoscale 2016, 8 (13), 7085–7093. 10.1039/C5NR09303A. [DOI] [PubMed] [Google Scholar]
  133. Le T.; Jamshidi E.; Beidaghi M.; Esfahani M. R. Functionalized-MXene thin-film nanocomposite hollow fiber membranes for enhanced PFAS removal from water. ACS Appl. Mater. Interfaces 2022, 14, 25397. 10.1021/acsami.2c03796. [DOI] [PubMed] [Google Scholar]
  134. Ghani A. A.; Shahzad A.; Moztahida M.; Tahir K.; Jeon H.; Kim B.; Lee D. S. Adsorption and electrochemical regeneration of intercalated Ti3C2Tx MXene for the removal of ciprofloxacin from wastewater. Chem. Eng. J. 2021, 421, 127780. 10.1016/j.cej.2020.127780. [DOI] [Google Scholar]
  135. Li Z. K.; Wei Y.; Gao X.; Ding L.; Lu Z.; Deng J.; Yang X.; Caro J.; Wang H. Antibiotics separation with MXene membranes based on regularly stacked high-aspect-ratio nanosheets. Angew. Chem. 2020, 132 (24), 9838–9843. 10.1002/ange.202002935. [DOI] [PubMed] [Google Scholar]
  136. Kim S.; Gholamirad F.; Yu M.; Park C. M.; Jang A.; Jang M.; Taheri-Qazvini N.; Yoon Y. Enhanced adsorption performance for selected pharmaceutical compounds by sonicated Ti3C2TX MXene. Chem. Eng. J. 2021, 406, 126789. 10.1016/j.cej.2020.126789. [DOI] [Google Scholar]
  137. Ren J.; Zhu Z.; Qiu Y.; Yu F.; Zhou T.; Ma J.; Zhao J. Enhanced adsorption performance of alginate/MXene/CoFe2O4 for antibiotic and heavy metal under rotating magnetic field. Chemosphere 2021, 284, 131284. 10.1016/j.chemosphere.2021.131284. [DOI] [PubMed] [Google Scholar]
  138. Meng F.; Seredych M.; Chen C.; Gura V.; Mikhalovsky S.; Sandeman S.; Ingavle G.; Ozulumba T.; Miao L.; Anasori B.; Gogotsi Y. MXene sorbents for removal of urea from dialysate: A step toward the wearable artificial kidney. ACS Nano 2018, 12 (10), 10518–10528. 10.1021/acsnano.8b06494. [DOI] [PubMed] [Google Scholar]
  139. Zhao Q.; Seredych M.; Precetti E.; Shuck C. E.; Harhay M.; Pang R.; Shan C.-X.; Gogotsi Y. Adsorption of uremic toxins using Ti3C2Tx MXene for dialysate regeneration. ACS Nano 2020, 14 (9), 11787–11798. 10.1021/acsnano.0c04546. [DOI] [PMC free article] [PubMed] [Google Scholar]
  140. Zandi P.; Ghasemy E.; Khedri M.; Rashidi A.; Maleki R.; Miri Jahromi A. Shedding light on miniaturized dialysis using MXene 2D materials: a computational chemistry approach. ACS Omega 2021, 6 (9), 6312–6325. 10.1021/acsomega.0c06118. [DOI] [PMC free article] [PubMed] [Google Scholar]
  141. Unal M. A.; Bayrakdar F.; Fusco L.; Besbinar O.; Shuck C. E.; Yalcin S.; Erken M. T.; Ozkul A.; Gurcan C.; Panatli O.; Summak G. Y.; Gokce C.; Orecchioni M.; Gazzi A.; Vitale F.; Somers J.; Demir E.; Yildiz S. S.; Nazir H.; Grivel J.-C.; Bedognetti D.; Crisanti A.; Akcali K. C.; Gogotsi Y.; Delogu L. G.; Yilmazer A. 2D MXenes with antiviral and immunomodulatory properties: A pilot study against SARS-CoV-2. Nano Today 2021, 38, 101136. 10.1016/j.nantod.2021.101136. [DOI] [PMC free article] [PubMed] [Google Scholar]
  142. Ciou J. H.; Li S.; Lee P. S. Ti3C2 MXene paper for the effective adsorption and controllable release of aroma molecules. Small 2019, 15 (38), 1903281. 10.1002/smll.201903281. [DOI] [PubMed] [Google Scholar]
  143. Ying Y.; Liu Y.; Wang X.; Mao Y.; Cao W.; Hu P.; Peng X. Two-dimensional titanium carbide for efficiently reductive removal of highly toxic chromium (VI) from water. ACS Appl. Mater. Interfaces 2015, 7 (3), 1795–1803. 10.1021/am5074722. [DOI] [PubMed] [Google Scholar]
  144. Fan Y.; Wei L.; Meng X.; Zhang W.; Yang N.; Jin Y.; Wang X.; Zhao M.; Liu S. An unprecedented high-temperature-tolerance 2D laminar MXene membrane for ultrafast hydrogen sieving. J. Membr. Sci. 2019, 569, 117–123. 10.1016/j.memsci.2018.10.017. [DOI] [Google Scholar]
  145. Wang Q.; Fan Y.; Wu C.; Jin Y.; Li C.; Sunarso J.; Meng X.; Yang N. Palladium-intercalated MXene membrane for efficient separation of H2/CO2: Combined experimental and modeling work. J. Membr. Sci. 2022, 653, 120533. 10.1016/j.memsci.2022.120533. [DOI] [Google Scholar]
  146. Qu K.; Dai L.; Xia Y.; Wang Y.; Zhang D.; Wu Y.; Yao Z.; Huang K.; Guo X.; Xu Z. Self-crosslinked MXene hollow fiber membranes for H2/CO2 separation. J. Membr. Sci. 2021, 638, 119669. 10.1016/j.memsci.2021.119669. [DOI] [Google Scholar]
  147. Fan Y.; Li J.; Wang S.; Meng X.; Jin Y.; Yang N.; Meng B.; Li J.; Liu S. Nickel (II) ion-intercalated MXene membranes for enhanced H2/CO2 separation. Frontiers of Chem. Sci. Eng. 2021, 15 (4), 882–891. 10.1007/s11705-020-1990-1. [DOI] [Google Scholar]
  148. Liu F.; Zhou A.; Chen J.; Zhang H.; Cao J.; Wang L.; Hu Q. Preparation and methane adsorption of two-dimensional carbide Ti2C. Adsorption 2016, 22 (7), 915–922. 10.1007/s10450-016-9795-8. [DOI] [Google Scholar]
  149. 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. 10.1016/j.apsusc.2017.04.239. [DOI] [Google Scholar]
  150. Wang B.; Zhou A.; Liu F.; Cao J.; Wang L.; Hu Q. Carbon dioxide adsorption of two-dimensional carbide MXenes. Journal of Advanced Ceramics 2018, 7 (3), 237–245. 10.1007/s40145-018-0275-3. [DOI] [Google Scholar]
  151. Morales-García A.; Fernández-Fernández A.; Viñes F.; Illas F. CO2 abatement using two-dimensional MXene carbides. J. Mater. Chem. A 2018, 6 (8), 3381–3385. 10.1039/C7TA11379J. [DOI] [Google Scholar]
  152. Zhang Y.; Zhou Z.; Lan J.; Zhang P. Prediction of Ti3C2O2 MXene as an effective capturer of formaldehyde. Appl. Surf. Sci. 2019, 469, 770–774. 10.1016/j.apsusc.2018.11.018. [DOI] [Google Scholar]
  153. Shen J.; Liu G.; Ji Y.; Liu Q.; Cheng L.; Guan K.; Zhang M.; Liu G.; Xiong J.; Yang J.; Jin W. 2D MXene nanofilms with tunable gas transport channels. Adv. Funct. Mater. 2018, 28 (31), 1801511. 10.1002/adfm.201801511. [DOI] [Google Scholar]
  154. Li L.; Zhang T.; Duan Y.; Wei Y.; Dong C.; Ding L.; Qiao Z.; Wang H. Selective gas diffusion in two-dimensional MXene lamellar membranes: insights from molecular dynamics simulations. J. Mater. Chem. A 2018, 6 (25), 11734–11742. 10.1039/C8TA03701A. [DOI] [Google Scholar]
  155. Emerenciano A. A.; do Nascimento R. M.; Barbosa A. P. C.; Ran K.; Meulenberg W. A.; Gonzalez-Julian J. Ti3C2 MXene membranes for gas separation: Influence of heat treatment conditions on d-spacing and surface functionalization. Membranes 2022, 12 (10), 1025. 10.3390/membranes12101025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  156. Jin Y.; Fan Y.; Meng X.; Zhang W.; Meng B.; Yang N.; Liu S. Theoretical and experimental insights into the mechanism for gas separation through nanochannels in 2D laminar MXene membranes. Processes 2019, 7 (10), 751. 10.3390/pr7100751. [DOI] [Google Scholar]
  157. Massoumılari S.; Doǧancı M.; Velioǧlu S. Unveiling the potential of MXenes for H2 purification and CO2 capture as an emerging family of nanomaterials. AIChE J. 2022, 68, e17837 10.1002/aic.17837. [DOI] [Google Scholar]
  158. Massoumılari Ş.High-throughput computational screening of MXene for CO2 capture from syngas and the effect of interlayer distance on ideal gas separation. Msc Thesis, Gebze Techical University, Institue of Nanotechnology, 2022. [Google Scholar]
  159. Lin H.; Gong K.; Hykys P.; Chen D.; Ying W.; Sofer Z.; Yan Y.; Li Z.; Peng X. Nanoconfined deep eutectic solvent in laminated MXene for efficient CO2 separation. Chem. Eng. J. 2021, 405, 126961. 10.1016/j.cej.2020.126961. [DOI] [Google Scholar]
  160. Liu G.; Cheng L.; Chen G.; Liang F.; Liu G.; Jin W. Pebax-based membrane filled with two-dimensional MXene nanosheets for efficient CO2 capture. Chem.-An Asian J. 2020, 15 (15), 2364–2370. 10.1002/asia.201901433. [DOI] [PubMed] [Google Scholar]
  161. Shamsabadi A. A.; Isfahani A. P.; Salestan S. K.; Rahimpour A.; Ghalei B.; Sivaniah E.; Soroush M. Pushing rubbery polymer membranes to be economic for CO2 separation: Embedment with Ti3C2Tx MXene nanosheets. ACS Appl. Mater. Interfaces 2020, 12 (3), 3984–3992. 10.1021/acsami.9b19960. [DOI] [PubMed] [Google Scholar]
  162. Guan W.; Yang X.; Dong C.; Yan X.; Zheng W.; Xi Y.; Ruan X.; Dai Y.; He G. Prestructured MXene fillers with uniform channels to enhance CO2 selective permeation in mixed matrix membranes. J. Appl. Polym. Sci. 2021, 138 (8), 49895. 10.1002/app.49895. [DOI] [Google Scholar]
  163. Shi F.; Sun J.; Wang J.; Liu M.; Yan Z.; Zhu B.; Li Y.; Cao X. MXene versus graphene oxide: Investigation on the effects of 2D nanosheets in mixed matrix membranes for CO2 separation. J. Membr. Sci. 2021, 620, 118850. 10.1016/j.memsci.2020.118850. [DOI] [Google Scholar]
  164. Wang D.; Ning H.; Xin Y.; Wang Y.; Li X.; Yao D.; Zheng Y.; Pan Y.; Zhang H.; He Z.; Liu C.; Qin M.; Wang Z.; Yang R.; Li P.; Yang Z. Transforming Ti3C2Tx MXenes into nanoscale ionic materials via an electronic interaction strategy. J. Mater. Chem. A 2021, 9 (27), 15441–15451. 10.1039/D1TA01744F. [DOI] [Google Scholar]
  165. Luo W.; Niu Z.; Mu P.; Li J. MXene/poly (ethylene glycol) mixed matrix membranes with excellent permeance for highly efficient separation of CO2/N2 and CO2/CH4. Colloids Surf.A: Physicochem. Eng.Aspects 2022, 640, 128481. 10.1016/j.colsurfa.2022.128481. [DOI] [Google Scholar]
  166. Hong X.; Lu Z.; Zhao Y.; Lyu L.; Ding L.; Wei Y.; Wang H. Fast fabrication of freestanding MXene-ZIF-8 dual-layered membranes for H2/CO2 separation. J. Membr. Sci. 2022, 642, 119982. 10.1016/j.memsci.2021.119982. [DOI] [Google Scholar]
  167. Li R.; Fu X.; Liu G.; Li J.; Zhou G.; Liu G.; Jin W. Room-temperature in situ synthesis of MOF@MXene membrane for efficient hydrogen purification. J. Membr. Sci. 2022, 664, 121097. 10.1016/j.memsci.2022.121097. [DOI] [Google Scholar]
  168. Anasori B.; Gogotsi Y. MXenes: Trends, growth, and future directions. Graphene and 2D Mater. 2022, 7, 75–79. 10.1007/s41127-022-00053-z. [DOI] [Google Scholar]
  169. Petukhov D.; Kan A.; Chumakov A.; Konovalov O.; Valeev R.; Eliseev A. MXene-based gas separation membranes with sorption type selectivity. J. Membr. Sci. 2021, 621, 118994. 10.1016/j.memsci.2020.118994. [DOI] [Google Scholar]
  170. Persson I.; Halim J.; Lind H.; Hansen T. W.; Wagner J. B.; Näslund L. Å.; Darakchieva V.; Palisaitis J.; Rosen J.; Persson P. O. 2D transition metal carbides (MXenes) for carbon capture. Adv. Mater. 2019, 31 (2), 1805472. 10.1002/adma.201805472. [DOI] [PubMed] [Google Scholar]
  171. Liu S.; Liu J.; Liu X.; Shang J.; Xu L.; Yu R.; Shui J. Hydrogen storage in incompletely etched multilayer Ti2CTx at room temperature. Nat. Nanotechnol. 2021, 16 (3), 331–336. 10.1038/s41565-020-00818-8. [DOI] [PubMed] [Google Scholar]
  172. Hajian S.; Khakbaz P.; Moshayedi M.; Maddipatla D.; Narakathu B. B.; Turkani V. S.; Bazuin B. J.; Pourfath M.; Atashbar M. Z. In Impact of Different Ratios of Fluorine, Oxygen, and Hydroxyl Surface Terminations on Ti3C2Tx MXene as Ammonia Sensor: A First-Principles Study; IEEE, 2018; pp 1–4. [Google Scholar]
  173. Lee E.; VahidMohammadi A.; Yoon Y. S.; Beidaghi M.; Kim D.-J. Two-dimensional Vanadium Carbide MXene for gas sensors with ultrahigh sensitivity toward nonpolar gases. ACS Sensors 2019, 4 (6), 1603–1611. 10.1021/acssensors.9b00303. [DOI] [PubMed] [Google Scholar]
  174. Koh H.-J.; Kim S. J.; Maleski K.; Cho S.-Y.; Kim Y.-J.; Ahn C. W.; Gogotsi Y.; Jung H.-T. Enhanced selectivity of MXene gas sensors through metal ion intercalation: In situ X-ray diffraction study. ACS Sensors 2019, 4 (5), 1365–1372. 10.1021/acssensors.9b00310. [DOI] [PubMed] [Google Scholar]
  175. Lee E.; VahidMohammadi A.; Prorok B. C.; Yoon Y. S.; Beidaghi M.; Kim D.-J. Room temperature gas sensing of two-dimensional titanium carbide (MXene). ACS Appl. Mater. Interfaces 2017, 9 (42), 37184–37190. 10.1021/acsami.7b11055. [DOI] [PubMed] [Google Scholar]
  176. Wu M.; He M.; Hu Q.; Wu Q.; Sun G.; Xie L.; Zhang Z.; Zhu Z.; Zhou A. Ti3C2 MXene-based sensors with high selectivity for NH3 detection at room temperature. ACS Sensors 2019, 4 (10), 2763–2770. 10.1021/acssensors.9b01308. [DOI] [PubMed] [Google Scholar]
  177. Zhao L.; Wang K.; Wei W.; Wang L.; Han W. High-performance flexible sensing devices based on polyaniline/MXene nanocomposites. InfoMat 2019, 1 (3), 407–416. 10.1002/inf2.12032. [DOI] [Google Scholar]
  178. Yu X.-F.; Li Y.-C.; Cheng J.-B.; Liu Z.-B.; Li Q.-Z.; Li W.-Z.; Yang X.; Xiao B. Monolayer Ti2CO2: A promising candidate for NH3 sensor or capturer with high sensitivity and selectivity. ACS Appl. Mater. Interfaces 2015, 7 (24), 13707–13713. 10.1021/acsami.5b03737. [DOI] [PubMed] [Google Scholar]
  179. Junkaew A.; Arroyave R. Enhancement of the selectivity of MXenes (M2C, M= Ti, V, Nb, Mo) via oxygen-functionalization: promising materials for gas-sensing and-separation. Phys. Chem. Chem. Phys. 2018, 20 (9), 6073–6082. 10.1039/C7CP08622A. [DOI] [PubMed] [Google Scholar]
  180. Azofra L. M.; Li N.; MacFarlane D. R.; Sun C. Promising prospects for 2D d2-d4 M3C2 transition metal carbides (MXenes) in N2 capture and conversion into ammonia. Energy Environ. Sci. 2016, 9 (8), 2545–2549. 10.1039/C6EE01800A. [DOI] [Google Scholar]
  181. Kim S.; Lee J.; Doo S.; Kang Y. C.; Koo C. M.; Kim S. J. Metal-ion-intercalated MXene nanosheet films for NH3 gas detection. ACS Appl. Nano Mater. 2021, 4 (12), 14249–14257. 10.1021/acsanm.1c03814. [DOI] [Google Scholar]
  182. Ma S.; Yuan D.; Jiao Z.; Wang T.; Dai X. Monolayer Sc2CO2: A promising candidate as a SO2 gas sensor or capturer. J. Phys. Chem. C 2017, 121 (43), 24077–24084. 10.1021/acs.jpcc.7b07921. [DOI] [Google Scholar]
  183. Wu L.; Lu X.; 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. 10.1016/j.bios.2018.02.021. [DOI] [PubMed] [Google Scholar]
  184. Kim S. J.; Koh H.-J.; Ren C. E.; Kwon O.; Maleski K.; Cho S.-Y.; Anasori B.; Kim C.-K.; Choi Y.-K.; Kim J.; Gogotsi Y.; Jung H.-T. Metallic Ti3C2Tx MXene gas sensors with ultrahigh signal-to-noise ratio. ACS Nano 2018, 12 (2), 986–993. 10.1021/acsnano.7b07460. [DOI] [PubMed] [Google Scholar]
  185. Rathi K.; Arkoti N. K.; Pal K. Fabrication of delaminated 2D metal carbide MXenes (Nb2CTx) by CTAB-based NO2 gas sensor with enhanced stability. Adv. Mater. Interfaces 2022, 9, 2200415. 10.1002/admi.202200415. [DOI] [Google Scholar]
  186. Yuan W.; Yang K.; Peng H.; Li F.; Yin F. A flexible VOCs sensor based on a 3D MXene framework with a high sensing performance. J. Mater. Chem. A 2018, 6 (37), 18116–18124. 10.1039/C8TA06928J. [DOI] [Google Scholar]
  187. Sun S.; Wang M.; Chang X.; Jiang Y.; Zhang D.; Wang D.; Zhang Y.; Lei Y. W18O49/Ti3C2Tx MXene nanocomposites for highly sensitive acetone gas sensor with low detection limit. Sens. Actuators, B 2020, 304, 127274. 10.1016/j.snb.2019.127274. [DOI] [Google Scholar]
  188. Zou S.; Gao J.; Liu L.; Lin Z.; Fu P.; Wang S.; Chen Z. Enhanced gas sensing properties at low working temperature of iron molybdate/MXene composite. J. Alloys Compd. 2020, 817, 152785. 10.1016/j.jallcom.2019.152785. [DOI] [Google Scholar]
  189. Bhardwaj R.; Hazra A. MXene-based gas sensors. J. Mat. Chem. C 2021, 9 (44), 15735–15754. 10.1039/D1TC04085E. [DOI] [Google Scholar]
  190. Devaraj M.; Rajendran S.; Hoang T. K.; Soto-Moscoso M. A review on MXene and its nanocomposites for the detection of toxic inorganic gases. Chemosphere 2022, 302, 134933. 10.1016/j.chemosphere.2022.134933. [DOI] [PubMed] [Google Scholar]
  191. Mehdi Aghaei S.; Aasi A.; Panchapakesan B. Experimental and theoretical advances in MXene-based gas sensors. ACS Omega 2021, 6 (4), 2450–2461. 10.1021/acsomega.0c05766. [DOI] [PMC free article] [PubMed] [Google Scholar]
  192. Regmi C.; Azadmanjiri J.; Mishra V.; Sofer Z.; Ashtiani S.; Friess K. Cellulose triacetate-based mixed-matrix membranes with MXene 2D filler—CO2/CH4 separation performance and comparison with TiO2-based 1D and 0D fillers. Membranes 2022, 12 (10), 917. 10.3390/membranes12100917. [DOI] [PMC free article] [PubMed] [Google Scholar]
  193. Kang K. M.; Kim D. W.; Ren C. E.; Cho K. M.; Kim S. J.; Choi J. H.; Nam Y. T.; Gogotsi Y.; Jung H.-T. Selective molecular separation on Ti3C2Tx-Graphene Oxide membranes during pressure-driven filtration: Comparison with Graphene Oxide and MXenes. ACS Appl. Mater. Interfaces 2017, 9 (51), 44687–44694. 10.1021/acsami.7b10932. [DOI] [PubMed] [Google Scholar]
  194. Wei S.; Xie Y.; Xing Y.; Wang L.; Ye H.; Xiong X.; Wang S.; Han K. Two-dimensional graphene Oxide/MXene composite lamellar membranes for efficient solvent permeation and molecular separation. J. Membr. Sci. 2019, 582, 414–422. 10.1016/j.memsci.2019.03.085. [DOI] [Google Scholar]
  195. Qin Y.; Zhao N.; Zhu Y.; Zhang Y.; Gao Q.; Dai Z.; You Y.; Lu X. Molecular insights into the microstructure of ethanol/water binary mixtures confined within typical 2D nanoslits: The role of the adsorbed layers induced by different solid surfaces. Fluid Phase Equilib. 2020, 509, 112452. 10.1016/j.fluid.2019.112452. [DOI] [Google Scholar]
  196. Liu G.; Shen J.; Ji Y.; Liu Q.; Liu G.; Yang J.; Jin W. Two-dimensional Ti2CTx MXene membranes with integrated and ordered nanochannels for efficient solvent dehydration. J. Mater. Chem. A 2019, 7 (19), 12095–12104. 10.1039/C9TA01507H. [DOI] [Google Scholar]
  197. Liu G.; Liu S.; Ma K.; Wang H.; Wang X.; Liu G.; Jin W. Polyelectrolyte functionalized Ti2CTx MXene membranes for pervaporation dehydration of isopropanol/water mixtures. Ind. Eng. Chem. Res. 2020, 59 (10), 4732–4741. 10.1021/acs.iecr.9b06881. [DOI] [Google Scholar]
  198. Wu Y.; Ding L.; Lu Z.; Deng J.; Wei Y. Two-dimensional MXene membrane for ethanol dehydration. J. Membr. Sci. 2019, 590, 117300. 10.1016/j.memsci.2019.117300. [DOI] [Google Scholar]
  199. Xu Z.; Liu G.; Ye H.; Jin W.; Cui Z. Two-dimensional MXene incorporated chitosan mixed-matrix membranes for efficient solvent dehydration. J. Membr. Sci. 2018, 563, 625–632. 10.1016/j.memsci.2018.05.044. [DOI] [Google Scholar]
  200. Li S.; Dai J.; Geng X.; Li J.; Li P.; Lei J.; Wang L.; He J. Highly selective sodium alginate mixed-matrix membrane incorporating multi-layered MXene for ethanol dehydration. Sep. Purif. Technol. 2020, 235, 116206. 10.1016/j.seppur.2019.116206. [DOI] [Google Scholar]
  201. Cai W.; Cheng X.; Chen X.; Li J.; Pei J. Poly (vinyl alcohol)-modified membranes by Ti3C2Tx for ethanol dehydration via pervaporation. ACS Omega 2020, 5 (12), 6277–6287. 10.1021/acsomega.9b03388. [DOI] [PMC free article] [PubMed] [Google Scholar]
  202. Yang G.; Xie Z.; Thornton A. W.; Doherty C. M.; Ding M.; Xu H.; Cran M.; Ng D.; Gray S. Ultrathin poly (vinyl alcohol)/MXene nanofilm composite membrane with facile intrusion-free construction for pervaporative separations. J. Membr. Sci. 2020, 614, 118490. 10.1016/j.memsci.2020.118490. [DOI] [Google Scholar]
  203. Wu X.; Hao L.; Zhang J.; Zhang X.; Wang J.; Liu J. Polymer-Ti3C2Tx composite membranes to overcome the trade-off in solvent resistant nanofiltration for alcohol-based system. J. Membrane Sci. 2016, 515, 175–188. 10.1016/j.memsci.2016.05.048. [DOI] [Google Scholar]
  204. Hao L.; Zhang H.; Wu X.; Zhang J.; Wang J.; Li Y. Novel thin-film nanocomposite membranes filled with multi-functional Ti3C2Tx nanosheets for task-specific solvent transport. Composites Part A: Appl. Sci. and Manufacturing 2017, 100, 139–149. 10.1016/j.compositesa.2017.05.003. [DOI] [Google Scholar]
  205. Wu X.; Chen Y.; Li W.; Chen C.; Zhang J.; Wang J. Heterostructured membranes with selective solvent-capture coatings and low-resistance 2D nanochannels for efficient mixed solvent separation. Sep. Purif.Technol. 2022, 283, 120217. 10.1016/j.seppur.2021.120217. [DOI] [Google Scholar]
  206. 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. 2019, 131 (51), 18695–18700. 10.1002/ange.201912570. [DOI] [PubMed] [Google Scholar]
  207. Liu P.; Hou J.; Zhang Y.; Li L.; Lu X.; Tang Z. Two-dimensional material membranes for critical separations. Inorg. Chem. Fronti. 2020, 7, 2560. 10.1039/D0QI00307G. [DOI] [Google Scholar]
  208. Hyun T.; Jeong J.; Chae A.; Kim Y. K.; Koh D.-Y. 2D-enabled membranes: materials and beyond. BMC Chem. Eng. 2019, 1, 12. 10.1186/s42480-019-0012-x. [DOI] [Google Scholar]
  209. 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 (35), 14334–14338. 10.1039/C4TA02638A. [DOI] [Google Scholar]
  210. Han R.; Ma X.; Xie Y.; Teng D.; Zhang S. Preparation of a new 2D MXene/PES composite membrane with excellent hydrophilicity and high flux. RSC Advances 2017, 7 (89), 56204–56210. 10.1039/C7RA10318B. [DOI] [Google Scholar]
  211. Zhang S.; Liao S.; Qi F.; Liu R.; Xiao T.; Hu J.; Li K.; Wang R.; Min Y. Direct deposition of two-dimensional MXene nanosheets on commercially available filter for fast and efficient dye removal. J. Hazardous Mater. 2020, 384, 121367. 10.1016/j.jhazmat.2019.121367. [DOI] [PubMed] [Google Scholar]
  212. Yousaf T.; Areeb A.; Murtaza M.; Munir A.; Khan Y.; Waseem A. Silane-grafted MXene (Ti3C2TX) membranes for enhanced water purification performance. ACS Omega 2022, 7, 19502. 10.1021/acsomega.2c01143. [DOI] [PMC free article] [PubMed] [Google Scholar]
  213. Pandey R. P.; Rasool K.; Madhavan V. E.; Aïssa B.; Gogotsi Y.; Mahmoud K. A. Ultrahigh-flux and fouling-resistant membranes based on layered silver/MXene (Ti3C2Tx) nanosheets. J. Mater. Chem. A 2018, 6 (8), 3522–3533. 10.1039/C7TA10888E. [DOI] [Google Scholar]
  214. He S.; Zhan Y.; Hu J.; Zhang G.; Zhao S.; Feng Q.; Yang W. Chemically stable two-dimensional MXene@UIO-66-(COOH)2 composite lamellar membrane for multi-component pollutant-oil-water emulsion separation. Composites Part B: Eng. 2020, 197, 108188. 10.1016/j.compositesb.2020.108188. [DOI] [Google Scholar]
  215. Long Q.; Zhao S.; Chen J.; Zhang Z.; Qi G.; Liu Z.-Q. Self-assembly enabled nano-intercalation for stable high-performance MXene membranes. J. Membr. Sci. 2021, 635, 119464. 10.1016/j.memsci.2021.119464. [DOI] [Google Scholar]
  216. Hu W.; Xie L.; Zeng H. Novel sodium alginate-assisted MXene nanosheets for ultrahigh rejection of multiple cations and dyes. J. Colloid Interface Sci. 2020, 568, 36–45. 10.1016/j.jcis.2020.02.028. [DOI] [PubMed] [Google Scholar]
  217. Tao M.-j.; Cheng S.-Q.; Han X.-L.; Yi F.; Li R.-H.; Rong Y.; Sun Y.; Liu Y. Alignment of MXene based membranes to enhance water purification. J. Membr. Sci. 2022, 662, 120965. 10.1016/j.memsci.2022.120965. [DOI] [Google Scholar]
  218. Yi M.; Heraly F.; Chang J.; Khorsand Kheirabad A.; Yuan J.; Wang Y.; Zhang M. A transport channel-regulated MXene membrane via organic phosphonic acids for efficient water permeation. Chem. Commun. 2021, 57 (51), 6245–6248. 10.1039/D1CC01464A. [DOI] [PubMed] [Google Scholar]
  219. Tong X.; Liu S.; Qu D.; Gao H.; Yan L.; Chen Y.; Crittenden J. Tannic acid-metal complex modified MXene membrane for contaminants removal from water. J. Membr. Sci. 2021, 622, 119042. 10.1016/j.memsci.2020.119042. [DOI] [Google Scholar]
  220. Pandey R. P.; Rasheed P. A.; Gomez T.; Azam R. S.; Mahmoud K. A. A fouling-resistant mixed-matrix nanofiltration membrane based on covalently cross-linked Ti3C2TX (MXene)/cellulose acetate. J. Membr. Sci. 2020, 607, 118139. 10.1016/j.memsci.2020.118139. [DOI] [Google Scholar]
  221. Krecker M. C.; Bukharina D.; Hatter C. B.; Gogotsi Y.; Tsukruk V. V. Bioencapsulated MXene flakes for enhanced stability and composite precursors. Adv. Funct. Mater. 2020, 30 (43), 2004554. 10.1002/adfm.202004554. [DOI] [Google Scholar]
  222. Han R.; Xie Y.; Ma X. Crosslinked P84 copolyimide/MXene mixed matrix membrane with excellent solvent resistance and permselectivity. Chinese J. Chem. Eng. 2019, 27 (4), 877–883. 10.1016/j.cjche.2018.10.005. [DOI] [Google Scholar]
  223. Lin Q.; Liu Y.; Zeng G.; Li X.; Wang B.; Cheng X.; Sengupta A.; Yang X.; Feng Z. Bionics inspired modified two-dimensional MXene composite membrane for high-throughput dye separation. J. Environ. Chem. Eng. 2021, 9 (4), 105711. 10.1016/j.jece.2021.105711. [DOI] [Google Scholar]
  224. Yi M.; Wang M.; Wang Y.; Wang Y.; Chang J.; Kheirabad A. K.; He H.; Yuan J.; Zhang M. Poly (ionic liquid)-armored MXene membrane: Interlayer engineering for facilitated water transport. Angew. Chem. 2022, 61, e202202515 10.1002/anie.202202515. [DOI] [PMC free article] [PubMed] [Google Scholar]
  225. Yao Y.-y.; Wang T.; Wu L.-g.; Chen H.-l. PES mixed-matrix membranes incorporating ZIF-8@ MXene nanocomposite for the efficient dye/salt separation. Desalination 2022, 543, 116116. 10.1016/j.desal.2022.116116. [DOI] [Google Scholar]
  226. Li Y.; Dai R.; Zhou H.; Li X.; Wang Z. Aramid nanofiber membranes reinforced by MXene nanosheets for recovery of dyes from textile wastewater. ACS Appl. Nano Mater. 2021, 4 (6), 6328–6336. 10.1021/acsanm.1c01217. [DOI] [Google Scholar]
  227. Liu T.; Liu X.; Graham N.; Yu W.; Sun K. Two-dimensional MXene incorporated graphene oxide composite membrane with enhanced water purification performance. J. Membr. Sci. 2020, 593, 117431. 10.1016/j.memsci.2019.117431. [DOI] [Google Scholar]
  228. Han R.; Wu P. High-performance graphene oxide nanofiltration membrane with continuous nanochannels prepared by the in situ oxidation of MXene. J. Mater. Chem. A 2019, 7 (11), 6475–6481. 10.1039/C9TA00137A. [DOI] [Google Scholar]
  229. Feng X.; Yu Z.; Long R.; Sun Y.; Wang M.; Li X.; Zeng G. Polydopamine intimate contacted two-dimensional/two-dimensional ultrathin nylon basement membrane supported RGO/PDA/MXene composite material for oil-water separation and dye removal. Sep. Purif. Technol. 2020, 247, 116945. 10.1016/j.seppur.2020.116945. [DOI] [Google Scholar]
  230. Zeng G.; Lin Q.; Wei K.; Liu Y.; Zheng S.; Zhan Y.; He S.; Patra T.; Chiao Y.-H. High-performing composite membrane based on dopamine-functionalized graphene oxide incorporated two-dimensional MXene nanosheets for water purification. J. Mater. Sci. 2021, 56 (11), 6814–6829. 10.1007/s10853-020-05746-5. [DOI] [Google Scholar]
  231. Ding M.; Xu H.; Chen W.; Kong Q.; Lin T.; Tao H.; Zhang K.; Liu Q.; Zhang K.; Xie Z. Construction of a hierarchical carbon nanotube/MXene membrane with distinct fusiform channels for efficient molecular separation. J. Mate. Chem. A 2020, 8 (43), 22666–22673. 10.1039/D0TA07354G. [DOI] [Google Scholar]
  232. Sun Y.; Xu D.; Li S.; Cui L.; Zhuang Y.; Xing W.; Jing W. Assembly of multidimensional MXene-carbon nanotube ultrathin membranes with an enhanced anti-swelling property for water purification. J. Membr. Sci. 2021, 623, 119075. 10.1016/j.memsci.2021.119075. [DOI] [Google Scholar]
  233. Gong X.; Zhang G.; Dong H.; Wang H.; Nie J.; Ma G. Self-assembled hierarchical heterogeneous MXene/COF membranes for efficient dye separations. J. Membr. Sci. 2022, 657, 120667. 10.1016/j.memsci.2022.120667. [DOI] [Google Scholar]
  234. Huang H.; Xu Y.; Lu Z.; Zhang A.; Zhang D.; Xue H.; Dong P.; Zhang J.; Goto T. Highly permeable and dye-rejective nanofiltration membranes of TiO2 and Bi2S3 double-embedded Ti3C2Tx with a visible-light-induced self-cleaning ability. J. Mate. Re. Technol. 2022, 18, 4156–4168. 10.1016/j.jmrt.2022.04.104. [DOI] [Google Scholar]
  235. Ren C. E.; Alhabeb M.; Byles B. W.; Zhao M.-Q.; Anasori B.; Pomerantseva E.; Mahmoud K. A.; Gogotsi Y. Voltage-gated ions sieving through 2D MXene Ti3C2T x membranes. ACS App. Nano Mater. 2018, 1 (7), 3644–3652. 10.1021/acsanm.8b00762. [DOI] [Google Scholar]
  236. Li J.; Xu C.; Long J.; Ding Z.; Yuan R.; Li Z. Lamellar MXene nanofiltration membranes for electrostatic modulation of molecular permeation: Implications for fine separation. ACS Appl. Nano Mater. 2022, 5, 7373. 10.1021/acsanm.2c01304. [DOI] [Google Scholar]
  237. Kim J. H.; Park G. S.; Kim Y.-J.; Choi E.; Kang J.; Kwon O.; Kim S. J.; Cho J. H.; Kim D. W. Large-area Ti3C2Tx-MXene coating: Toward industrial-scale fabrication and molecular separation. ACS Nano 2021, 15 (5), 8860–8869. 10.1021/acsnano.1c01448. [DOI] [PubMed] [Google Scholar]
  238. Xing Y.; Akonkwa G.; Liu Z.; Ye H.; Han K. Crumpled two-dimensional Ti3C2T x MXene lamellar membranes for solvent permeation and separation. ACS Appl. Nano Mater. 2020, 3 (2), 1526–1534. 10.1021/acsanm.9b02322. [DOI] [Google Scholar]
  239. Li S.; Gu W.; Sun Y.; Zou D.; Jing W. Perforative pore formation on nanoplates for 2D porous MXene membranes via H2O2 mild etching. Ceram. Int. 2021, 47 (21), 29930–29940. 10.1016/j.ceramint.2021.07.166. [DOI] [Google Scholar]
  240. Xiang J.; Wang X.; Ding M.; Tang X.; Zhang S.; Zhang X.; Xie Z. The role of lateral size of MXene nanosheets in membrane filtration of dyeing wastewater: Membrane characteristic and performance. Chemosphere 2022, 294, 133728. 10.1016/j.chemosphere.2022.133728. [DOI] [PubMed] [Google Scholar]
  241. Feng C.; Ou K.; Zhang Z.; Liu Y.; Huang Y.; Wang Z.; Lv Y.; Miao Y.-E; Wang Y.; Lan Q.; Liu T. Dual-layered covalent organic framework/MXene membranes with short paths for fast water treatment. J. Membr. Sci. 2022, 658, 120761. 10.1016/j.memsci.2022.120761. [DOI] [Google Scholar]
  242. Cai C.; Wang R.; Liu S.; Yan X.; Zhang L.; Wang M.; Tong Q.; Jiao T. Synthesis of self-assembled phytic acid-MXene nanocomposites via a facile hydrothermal approach with elevated dye adsorption capacities. Colloids Surf., A 2020, 589, 124468. 10.1016/j.colsurfa.2020.124468. [DOI] [Google Scholar]
  243. Zhang P.; Zhang Y.; Wang L.; Qiu K.; Tang X.; Gibson J. K.; Liu X.; Mei L.; An S.; Huang Z.; Ren P.; Wang Y.; Chai Z.; Shi W. Bioinspired macrocyclic molecule supported two-dimensional lamellar membrane with robust interlayer structure for high-efficiency nanofiltration. Adv. Sci. 2023, 10 (5), 2206516. 10.1002/advs.202206516. [DOI] [PMC free article] [PubMed] [Google Scholar]
  244. Kallem P.; Elashwah N.; Bharath G.; Hegab H. M.; Hasan S. W.; Banat F. Zwitterion-grafted 2D MXene (Ti3C2TX) nanocomposite membranes with improved water permeability and self-cleaning properties. ACS Appl. Nano Mater. 2023, 6, 607–621. 10.1021/acsanm.2c04722. [DOI] [Google Scholar]
  245. Jun B.-M.; Kim S.; Rho H.; Park C. M.; Yoon Y. Ultrasound-assisted Ti3C2Tx MXene adsorption of dyes: Removal performance and mechanism analyses via dynamic light scattering. Chemosphere 2020, 254, 126827. 10.1016/j.chemosphere.2020.126827. [DOI] [PubMed] [Google Scholar]
  246. Li K.; Zou G.; Jiao T.; Xing R.; Zhang L.; Zhou J.; Zhang Q.; Peng Q. Self-assembled MXene-based nanocomposites via layer-by-layer strategy for elevated adsorption capacities. Colloids Surf., A 2018, 553, 105–113. 10.1016/j.colsurfa.2018.05.044. [DOI] [Google Scholar]
  247. Lei Y.; Cui Y.; Huang Q.; Dou J.; Gan D.; Deng F.; Liu M.; Li X.; Zhang X.; Wei Y. Facile preparation of sulfonic groups functionalized Mxenes for efficient removal of methylene blue. Ceram. Int. 2019, 45 (14), 17653–17661. 10.1016/j.ceramint.2019.05.331. [DOI] [Google Scholar]
  248. Vakili M.; Cagnetta G.; Huang J.; Yu G.; Yuan J. Synthesis and regeneration of a MXene-based pollutant adsorbent by mechanochemical methods. Molecules 2019, 24 (13), 2478. 10.3390/molecules24132478. [DOI] [PMC free article] [PubMed] [Google Scholar]
  249. Wei Z.; Peigen Z.; Wubian T.; Xia Q.; Yamei Z.; ZhengMing S. Alkali treated Ti3C2Tx MXenes and their dye adsorption performance. Mater. Chem. Phys. 2018, 206, 270–276. 10.1016/j.matchemphys.2017.12.034. [DOI] [Google Scholar]
  250. Zhu Z.; Xiang M.; Shan L.; He T.; Zhang P. Effect of temperature on methylene blue removal with novel 2D-Magnetism titanium carbide. J. Solid State Chem. 2019, 280, 120989. 10.1016/j.jssc.2019.120989. [DOI] [Google Scholar]
  251. Zhang P.; Xiang M.; Liu H.; Yang C.; Deng S. Novel two-dimensional magnetic titanium carbide for methylene blue removal over a wide pH range: insight into removal performance and mechanism. ACS Appl. Mater. Interfaces 2019, 11 (27), 24027–24036. 10.1021/acsami.9b04222. [DOI] [PubMed] [Google Scholar]
  252. Jun B.-M.; Heo J.; Taheri-Qazvini N.; Park C. M.; Yoon Y. Adsorption of selected dyes on Ti3C2Tx MXene and Al-based metal-organic framework. Ceram. Int. 2020, 46 (3), 2960–2968. 10.1016/j.ceramint.2019.09.293. [DOI] [Google Scholar]
  253. Hao C.; Li G.; Wang G.; Chen W.; Wang S. Preparation of acrylic acid modified alkalized MXene adsorbent and study on its dye adsorption performance. Colloids Surfaces A: Physicochem. Eng. Aspects 2022, 632, 127730. 10.1016/j.colsurfa.2021.127730. [DOI] [Google Scholar]
  254. Wang L.; Tao W.; Yuan L.; Liu Z.; Huang Q.; Chai Z.; Gibson J. K.; Shi W. Rational control of the interlayer space inside two-dimensional titanium carbides for highly efficient uranium removal and imprisonment. Chem. Commun. 2017, 53 (89), 12084–12087. 10.1039/C7CC06740B. [DOI] [PubMed] [Google Scholar]
  255. Rethinasabapathy M.; Bhaskaran G.; Park B.; Shin J.-Y.; Kim W.-S.; Ryu J.; Huh Y. S. Iron oxide (Fe3O4)-laden titanium carbide (Ti3C2Tx) MXene stacks for the efficient sequestration of cationic dyes from aqueous solution. Chemosphere 2022, 286, 131679. 10.1016/j.chemosphere.2021.131679. [DOI] [PubMed] [Google Scholar]
  256. Luo S.; Wang R.; Yin J.; Jiao T.; Chen K.; Zou G.; Zhang L.; Zhou J.; Zhang L.; Peng Q. Preparation and dye degradation performances of self-assembled MXene-Co3O4 nanocomposites synthesized via solvothermal approach. ACS Omega 2019, 4 (2), 3946–3953. 10.1021/acsomega.9b00231. [DOI] [PMC free article] [PubMed] [Google Scholar]
  257. Gu C.; Weng W.; Lu C.; Tan P.; Jiang Y.; Zhang Q.; Liu X.; Sun L. Decorating MXene with tiny ZIF-8 nanoparticles: An effective approach to construct composites for water pollutant removal. Chinese J. Chem. Eng. 2022, 42, 42–48. 10.1016/j.cjche.2021.06.004. [DOI] [Google Scholar]
  258. Peng C.; Wei P.; Chen X.; Zhang Y.; Zhu F.; Cao Y.; Wang H.; Yu H.; Peng F. A hydrothermal etching route to synthesis of 2D MXene (Ti3C2, Nb2C): Enhanced exfoliation and improved adsorption performance. Ceram. Int. 2018, 44 (15), 18886–18893. 10.1016/j.ceramint.2018.07.124. [DOI] [Google Scholar]
  259. Zhang Z.-H.; Xu J.-Y.; Yang X.-L. MXene/sodium alginate gel beads for adsorption of methylene blue. Mater. Chem. Phys. 2021, 260, 124123. 10.1016/j.matchemphys.2020.124123. [DOI] [Google Scholar]
  260. Lei H.; Hao Z.; Chen K.; Chen Y.; Zhang J.; Hu Z.; Song Y.; Rao P.; Huang Q. Insight into adsorption performance and mechanism on efficient removal of methylene blue by accordion-like V2CTx MXene. J. Phys. Chem. Lett. 2020, 11 (11), 4253–4260. 10.1021/acs.jpclett.0c00973. [DOI] [PubMed] [Google Scholar]
  261. Yan Y.; Han H.; Dai Y.; Zhu H.; Liu W.; Tang X.; Gan W.; Li H. Nb2CT x MXene nanosheets for dye adsorption. ACS App. Nano Mater. 2021, 4 (11), 11763–11769. 10.1021/acsanm.1c02339. [DOI] [Google Scholar]
  262. Feng Y.; Wang H.; Xu J.; Du X.; Cheng X.; Du Z.; Wang H. Fabrication of MXene/PEI functionalized sodium alginate aerogel and its excellent adsorption behavior for Cr (VI) and Congo Red from aqueous solution. J. Hazardous Mater. 2021, 416, 125777. 10.1016/j.jhazmat.2021.125777. [DOI] [PubMed] [Google Scholar]
  263. Lim S.; Park H.; Kim J. H.; Yang J.; Kwak C.; Kim J.; Ryu S. Y.; Lee J. Polyelectrolyte-grafted Ti3C2-MXenes stable in extreme salinity aquatic conditions for remediation of contaminated subsurface environments. RSC Adv. 2020, 10 (43), 25966–25978. 10.1039/D0RA04348F. [DOI] [PMC free article] [PubMed] [Google Scholar]
  264. My Tran N.; Thanh Hoai Ta Q.; Sreedhar A.; Noh J.-S. Ti3C2Tx MXene playing as a strong methylene blue adsorbent in wastewater. Appl. Surf. Sci. 2021, 537, 148006. 10.1016/j.apsusc.2020.148006. [DOI] [Google Scholar]
  265. Kadhom M.; Kalash K.; Al-Furaiji M. Performance of 2D MXene as an adsorbent for malachite green removal. Chemosphere 2022, 290, 133256. 10.1016/j.chemosphere.2021.133256. [DOI] [PubMed] [Google Scholar]
  266. Sun B.; Dong X.; Li H.; Shang Y.; Zhang Y.; Hu F.; Gu S.; Wu Y.; Gao T.; Zhou G. Surface charge engineering for two-dimensional Ti2CTx MXene for highly efficient and selective removal of cationic dye from aqueous solution. Sep. Purif. Technol. 2021, 272, 118964. 10.1016/j.seppur.2021.118964. [DOI] [Google Scholar]
  267. Lim S.; Kim J. H.; Park H.; Kwak C.; Yang J.; Kim J.; Ryu S. Y.; Lee J. Role of electrostatic interactions in the adsorption of dye molecules by Ti3C2-MXenes. RSC Adv. 2021, 11 (11), 6201–6211. 10.1039/D0RA10876F. [DOI] [PMC free article] [PubMed] [Google Scholar]
  268. Zhang K.-N.; Wang C.-Z.; Lü Q.-F.; Chen M.-H. Enzymatic hydrolysis lignin functionalized Ti3C2Tx nanosheets for effective removal of MB and Cu2+ ions. Int. J. Biol. Macromol. 2022, 209, 680–691. 10.1016/j.ijbiomac.2022.04.036. [DOI] [PubMed] [Google Scholar]
  269. Abuhasel K.; Kchaou M.; Alquraish M.; Munusamy Y.; Jeng Y. T. Oily wastewater treatment: Overview of conventional and modern methods, challenges, and future opportunities. Water 2021, 13 (7), 980. 10.3390/w13070980. [DOI] [Google Scholar]
  270. Saththasivam J.; Wang K.; Yiming W.; Liu Z.; Mahmoud K. A. A flexible Ti3C2Tx (MXene)/paper membrane for efficient oil/water separation. RSC Adv. 2019, 9 (29), 16296–16304. 10.1039/C9RA02129A. [DOI] [PMC free article] [PubMed] [Google Scholar]
  271. Zhang H.; Wang Z.; Shen Y.; Mu P.; Wang Q.; Li J. Ultrathin 2D Ti3C2Tx MXene membrane for effective separation of oil-in-water emulsions in acidic, alkaline, and salty environment. J. Colloid Interface Sci. 2020, 561, 861–869. 10.1016/j.jcis.2019.11.069. [DOI] [PubMed] [Google Scholar]
  272. Zeng G.; Wei K.; Zhang H.; Zhang J.; Lin Q.; Cheng X.; Sengupta A.; Chiao Y.-H. Ultra-high oil-water separation membrane based on two-dimensional MXene (Ti3C2Tx) by co-incorporation of halloysite nanotubes and polydopamine. Appl. Clay Sci. 2021, 211, 106177. 10.1016/j.clay.2021.106177. [DOI] [Google Scholar]
  273. Li Z. K.; Liu Y.; Li L.; Wei Y.; Caro J.; Wang H. Ultra-thin titanium carbide (MXene) sheet membranes for high-efficient oil/water emulsions separation. J. Membrane Sci. 2019, 592, 117361. 10.1016/j.memsci.2019.117361. [DOI] [Google Scholar]
  274. Long X.; Zhao G.; Hu J.; Zheng Y.; Zhang J.; Zuo Y.; Jiao F. Cracked-earth-like titanium carbide MXene membranes with abundant hydroxyl groups for oil-in-water emulsion separation. J. Colloid Interface Sci. 2022, 607, 378–388. 10.1016/j.jcis.2021.08.175. [DOI] [PubMed] [Google Scholar]
  275. Huang Z.; Shen L.; Lin H.; Li B.; Chen C.; Xu Y.; Li R.; Zhang M.; Zhao D. Fabrication of fibrous MXene nanoribbons (MNRs) membrane with efficient performance for oil-water separation. J. Membr. Sci. 2022, 661, 120949. 10.1016/j.memsci.2022.120949. [DOI] [Google Scholar]
  276. Velioǧlu S.; Han L.; Chew J. W. Understanding membrane pore-wetting in the membrane distillation of oil emulsions via molecular dynamics simulations. J. Membr. Sci. 2018, 551, 76–84. 10.1016/j.memsci.2018.01.027. [DOI] [Google Scholar]
  277. Han L.; Tan Y. Z.; Netke T.; Fane A. G.; Chew J. W. Understanding oily wastewater treatment via membrane distillation. J. Membr. Sci. 2017, 539, 284–294. 10.1016/j.memsci.2017.06.012. [DOI] [Google Scholar]
  278. Liu Y.; Lin Q.; Zeng G.; Zhang L.; Zhou Y.; Sengupta A. Nature-inspired green method decorated MXene-based composite membrane for high-efficiency oil/water separation. Sep. Purif. Technol. 2022, 283, 120218. 10.1016/j.seppur.2021.120218. [DOI] [Google Scholar]
  279. Feng X.; Yu Z.; Long R.; Sun Y.; Wang M.; Li X.; Zeng G. Polydopamine intimate contacted two-dimensional/two-dimensional ultrathin nylon basement membrane supported RGO/PDA/MXene composite material for oil-water separation and dye removal. Sep. Purif. Technol. 2020, 247, 116945. 10.1016/j.seppur.2020.116945. [DOI] [Google Scholar]
  280. Zhang L.; Wei K.; Zeng G.; Lin Q.; Liu X.; Chen Y.; Sengupta A. High-efficient oil/water separation membrane based on MXene nanosheets by co-incorporation of APTES and amine functionalized carbon nanotubes. J. Environ. Chem. Eng. 2021, 9 (6), 106658. 10.1016/j.jece.2021.106658. [DOI] [Google Scholar]
  281. Ren R.-P.; Wang Z.; Ren J.; Lv Y.-K. Highly compressible polyimide/graphene aerogel for efficient oil/water separation. J. Mater. Sci. 2019, 54 (7), 5918–5926. 10.1007/s10853-018-03238-1. [DOI] [Google Scholar]
  282. Xue J.; Zhu L.; Zhu X.; Li H.; Ma C.; Yu S.; Sun D.; Xia F.; Xue Q. Tetradecylamine-MXene functionalized melamine sponge for effective oil/water separation and selective oil adsorption. Sep. Purif. Technol. 2021, 259, 118106. 10.1016/j.seppur.2020.118106. [DOI] [Google Scholar]
  283. Cai C.; Wei Z.; Huang Y.; Fu Y. Wood-inspired superelastic MXene aerogels with superior photothermal conversion and durable superhydrophobicity for clean-up of super-viscous crude oil. Chem. Eng. J. 2021, 421, 127772. 10.1016/j.cej.2020.127772. [DOI] [Google Scholar]
  284. Ma X.; Chen K.; Li S.; Gnanasekar P.; Zhong Y.; An Y.; Luo Q.; Huang Q.; Zhu J.; Chen J.; Yan N. Degradable Ti3C2Tx MXene nanosheets containing a lignin polyurethane photothermal foam (LPUF) for rapid crude oil cleanup. ACS Appl. Nano Mater. 2022, 5 (2), 2848–2858. 10.1021/acsanm.1c04556. [DOI] [Google Scholar]
  285. Du Y.; Si P.; Wei L.; Wang Y.; Tu Y.; Zuo G.; Yu B.; Zhang X.; Ye S. Demulsification of acidic oil-in-water emulsions driven by chitosan loaded Ti3C2Tx. Appl. Surf. Sci. 2019, 476, 878–885. 10.1016/j.apsusc.2019.01.106. [DOI] [Google Scholar]
  286. Sun A.; Zhan Y.; Feng Q.; Yang W.; Dong H.; Liu Y.; Chen X.; Chen Y. Assembly of MXene/ZnO heterojunction onto electrospun poly (arylene ether nitrile) fibrous membrane for favorable oil/water separation with high permeability and synergetic antifouling performance. J. Membr. Sci. 2022, 663, 120933. 10.1016/j.memsci.2022.120933. [DOI] [Google Scholar]
  287. Zhang Y.; Chen X.; Luo C.; Gu J.; Li M.; Chao M.; Chen X.; Chen T.; Yan L.; Wang X. Column-to-beam structure house inspired MXene-based integrated membrane with stable interlayer spacing for water purification. Adv. Funct. Mater. 2022, 32 (22), 2111660. 10.1002/adfm.202111660. [DOI] [Google Scholar]
  288. Wang N.-N.; Wang H.; Wang Y.-Y.; Wei Y.-H.; Si J.-Y.; Yuen A. C. Y.; Xie J.-S.; Yu B.; Zhu S.-E; Lu H.-D.; Yang W.; Chan Q. N.; Yeoh G.-H. Robust, lightweight, hydrophobic, and fire-retarded polyimide/MXene aerogels for effective oil/water separation. ACS Appl. Mater. Interfaces 2019, 11 (43), 40512–40523. 10.1021/acsami.9b14265. [DOI] [PubMed] [Google Scholar]
  289. Moghaddasi A.; Sobolčiak P.; Popelka A.; Krupa I. Separation of water/oil emulsions by an electrospun copolyamide mat covered with a 2D Ti3C2Tx MXene. Materials 2020, 13 (14), 3171. 10.3390/ma13143171. [DOI] [PMC free article] [PubMed] [Google Scholar]
  290. Liu Y.; Yu X.; Liu Y.; Xu Y.; Chang Z.; Wang D.; Li Q. Expanded polytetrafluoroethylene/MXene nanosheet composites with hydrophilicity and lipophilicity for purification of oil spills and wastewater. ACS Appl. Nano Mater. 2022, 5 (2), 2483–2491. 10.1021/acsanm.1c04166. [DOI] [Google Scholar]
  291. Yan X.; Yang C.; Ma C.; Tao H.; Cheng S.; Chen L.; Wang G.; Lin X.; Yao C. A novel janus membrane modified by MXene for enhanced anti-fouling and anti-wetting in direct contact membrane distillation. Chemosphere 2022, 307, 136114. 10.1016/j.chemosphere.2022.136114. [DOI] [PubMed] [Google Scholar]
  292. Shahzad A.; Nawaz M.; Moztahida M.; Jang J.; Tahir K.; Kim J.; Lim Y.; Vassiliadis V. S.; Woo S. H.; Lee D. S. Ti3C2Tx MXene core-shell spheres for ultrahigh removal of mercuric ions. Chem. Eng. J. 2019, 368, 400–408. 10.1016/j.cej.2019.02.160. [DOI] [Google Scholar]
  293. Dong X.; Wang Y.; Jia M.; Niu Z.; Cai J.; Yu X.; Ke X.; Yao J.; Zhang X. Sustainable and scalable in-situ synthesis of hydrochar-wrapped Ti3AlC2-derived nanofibers as adsorbents to remove heavy metals. Bioresource Technol. 2019, 282, 222–227. 10.1016/j.biortech.2019.03.010. [DOI] [PubMed] [Google Scholar]
  294. Yang X.; Liu Y.; Hu S.; Yu F.; He Z.; Zeng G.; Feng Z.; Sengupta A. Construction of Fe3O4@MXene composite nanofiltration membrane for heavy metal ions removal from wastewater. Polym. Adv. Technol. 2021, 32 (3), 1000–1010. 10.1002/pat.5148. [DOI] [Google Scholar]
  295. Karthikeyan P.; Ramkumar K.; Pandi K.; Fayyaz A.; Meenakshi S.; Park C. M. Effective removal of Cr (VI) and methyl orange from the aqueous environment using two-dimensional (2D) Ti3C2Tx MXene nanosheets. Ceram. Int. 2020, 47, 1. 10.1016/j.ceramint.2020.09.221. [DOI] [Google Scholar]
  296. Jin L.; Chai L.; Yang W.; Wang H.; Zhang L. Two-dimensional titanium carbides (Ti3C2Tx) functionalized by poly (m-phenylenediamine) for efficient adsorption and reduction of hexavalent chromium. International J. Environ. Res.and Public Health 2020, 17 (1), 167. 10.3390/ijerph17010167. [DOI] [PMC free article] [PubMed] [Google Scholar]
  297. Kong A.; Sun Y.; Peng M.; Gu H.; Fu Y.; Zhang J.; Li W. Amino-functionalized MXenes for efficient removal of Cr (VI). Colloids Surf., A 2021, 617, 126388. 10.1016/j.colsurfa.2021.126388. [DOI] [Google Scholar]
  298. Zou G.; Guo J.; Peng Q.; Zhou A.; Zhang Q.; Liu B. Synthesis of urchin-like rutile titania carbon nanocomposites by iron-facilitated phase transformation of MXene for environmental remediation. J. Mater. Chem. A 2016, 4 (2), 489–499. 10.1039/C5TA07343J. [DOI] [Google Scholar]
  299. He L.; Huang D.; He Z.; Yang X.; Yue G.; Zhu J.; Astruc D.; Zhao P. Nanoscale zero-valent iron intercalated 2D titanium carbides for removal of Cr (VI) in aqueous solution and the mechanistic aspect. J. Hazardous Mater. 2020, 388, 121761. 10.1016/j.jhazmat.2019.121761. [DOI] [PubMed] [Google Scholar]
  300. Xie X.; Chen C.; Zhang N.; Tang Z.-R.; Jiang J.; Xu Y.-J. Microstructure and surface control of MXene films for water purification. Nature Sustainability 2019, 2 (9), 856–862. 10.1038/s41893-019-0373-4. [DOI] [Google Scholar]
  301. Wang H.; Cui H.; Song X.; Xu R.; Wei N.; Tian J.; Niu H. Facile synthesis of heterojunction of MXenes/TiO2 nanoparticles towards enhanced hexavalent chromium removal. J. Colloid Interface Sci. 2020, 561, 46–57. 10.1016/j.jcis.2019.11.120. [DOI] [PubMed] [Google Scholar]
  302. Tang Y.; Yang C.; Que W. A novel two-dimensional accordion-like titanium carbide (MXene) for adsorption of Cr (VI) from aqueous solution. J. Adv. Dielectrics 2018, 08 (05), 1850035. 10.1142/S2010135X18500352. [DOI] [Google Scholar]
  303. Gan D.; Huang Q.; Dou J.; Huang H.; Chen J.; Liu M.; Wen Y.; Yang Z.; Zhang X.; Wei Y. Bioinspired functionalization of MXenes (Ti3C2Tx) with amino acids for efficient removal of heavy metal ions. Appl. Surf. Sci. 2020, 504, 144603. 10.1016/j.apsusc.2019.144603. [DOI] [Google Scholar]
  304. Elumalai S.; Yoshimura M.; Ogawa M. Simultaneous delamination and rutile formation on the surface of Ti3C2Tx MXene for copper adsorption. Chem.-An Asian J. 2020, 15 (7), 1044–1051. 10.1002/asia.202000090. [DOI] [PubMed] [Google Scholar]
  305. Shahzad A.; Rasool K.; Miran W.; Nawaz M.; Jang J.; Mahmoud K. A.; Lee D. S. Two-dimensional Ti3C2Tx MXene nanosheets for efficient copper removal from water. ACS Sustainable Chem. Eng. 2017, 5 (12), 11481–11488. 10.1021/acssuschemeng.7b02695. [DOI] [Google Scholar]
  306. Dong Y.; Sang D.; He C.; Sheng X.; Lei L. MXene/alginate composites for lead and copper ion removal from aqueous solutions. RSC Adv. 2019, 9 (50), 29015–29022. 10.1039/C9RA05251H. [DOI] [PMC free article] [PubMed] [Google Scholar]
  307. Zhang G.; Wang T.; Xu Z.; Liu M.; Shen C.; Meng Q. Synthesis of amino-functionalized Ti3C2Tx MXene by alkalization-grafting modification for efficient lead adsorption. Chem. Commun. 2020, 56 (76), 11283–11286. 10.1039/D0CC04265J. [DOI] [PubMed] [Google Scholar]
  308. Fu K.; Liu X.; Yu D.; Luo J.; Wang Z.; Crittenden J. C. Highly efficient and selective Hg (II) removal from water using multilayered Ti3C2Ox MXene via adsorption coupled with catalytic reduction mechanism. Environ. Sci. Technol. 2020, 54 (24), 16212–16220. 10.1021/acs.est.0c05532. [DOI] [PubMed] [Google Scholar]
  309. Shahzad A.; Jang J.; Lim S.-R.; Lee D. S. Unique selectivity and rapid uptake of molybdenum-disulfide-functionalized MXene nanocomposite for mercury adsorption. Environ. Res. 2020, 182, 109005. 10.1016/j.envres.2019.109005. [DOI] [PubMed] [Google Scholar]
  310. Shahzad A.; Rasool K.; Miran W.; Nawaz M.; Jang J.; Mahmoud K. A.; Lee D. S. Mercuric ion capturing by recoverable titanium carbide magnetic nanocomposite. J. Hazardous Mater. 2018, 344, 811–818. 10.1016/j.jhazmat.2017.11.026. [DOI] [PubMed] [Google Scholar]
  311. Feng X.; Yu Z.; Long R.; Li X.; Shao L.; Zeng H.; Zeng G.; Zuo Y. Self-assembling 2D/2D (MXene/LDH) materials achieve ultra-high adsorption of heavy metals Ni2+ through terminal group modification. Sep. Purif. Technol. 2020, 253, 117525. 10.1016/j.seppur.2020.117525. [DOI] [Google Scholar]
  312. Jun B.-M.; Her N.; Park C. M.; Yoon Y. Effective removal of Pb (ii) from synthetic wastewater using Ti3C2Tx MXene. Environ. Sci.: Water Res. Technol. 2020, 6 (1), 173–180. 10.1039/C9EW00625G. [DOI] [Google Scholar]
  313. 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. 10.1016/j.molliq.2019.111810. [DOI] [Google Scholar]
  314. Peng Q.; Guo J.; Zhang Q.; Xiang J.; Liu B.; Zhou A.; Liu R.; Tian Y. Unique lead adsorption behavior of activated hydroxyl group in two-dimensional titanium carbide. J. Am. Chem. Soc. 2014, 136 (11), 4113–4116. 10.1021/ja500506k. [DOI] [PubMed] [Google Scholar]
  315. Du Y.; Yu B.; Wei L.; Wang Y.; Zhang X.; Ye S. Efficient removal of Pb (II) by Ti3C2Tx powder modified with a silane coupling agent. J. Mater. Sci. 2019, 54 (20), 13283–13297. 10.1007/s10853-019-03814-z. [DOI] [Google Scholar]
  316. Gu P.; Xing J.; Wen T.; Zhang R.; Wang J.; Zhao G.; Hayat T.; Ai Y.; Lin Z.; Wang X. Experimental and theoretical calculation investigation on efficient Pb (II) adsorption on etched Ti3AlC2 nanofibers and nanosheets. Environ. Sci.: Nano 2018, 5 (4), 946–955. 10.1039/C8EN00029H. [DOI] [Google Scholar]
  317. Pandey R. P.; Rasool K.; Abdul Rasheed P.; Mahmoud K. A. Reductive sequestration of toxic bromate from drinking water using lamellar two-dimensional Ti3C2TX (MXene). ACS Sustainable Chem. Eng. 2018, 6 (6), 7910–7917. 10.1021/acssuschemeng.8b01147. [DOI] [Google Scholar]
  318. Guo X.; Zhang X.; Zhao S.; Huang Q.; Xue J. High adsorption capacity of heavy metals on two-dimensional MXenes: An ab initio study with molecular dynamics simulation. Phys. Chem. Chem. Phys. 2016, 18 (1), 228–233. 10.1039/C5CP06078H. [DOI] [PubMed] [Google Scholar]
  319. Yan Y.; An Q.; Xiao Z.; Zheng W.; Zhai S. Flexible core-shell/bead-like alginate@PEI with exceptional adsorption capacity, recycling performance toward batch and column sorption of Cr (VI). Chem. Eng. J. 2017, 313, 475–486. 10.1016/j.cej.2016.12.099. [DOI] [Google Scholar]
  320. Chen B.; Zhao X.; Liu Y.; Xu B.; Pan X. Highly stable and covalently functionalized magnetic nanoparticles by polyethyleneimine for Cr (VI) adsorption in aqueous solution. RSC Adv. 2015, 5 (2), 1398–1405. 10.1039/C4RA10602D. [DOI] [Google Scholar]
  321. Jun B.-M.; Park C. M.; Heo J.; Yoon Y. Adsorption of Ba2+ and Sr2+ on Ti3C2Tx MXene in model fracking wastewater. J. Environ. Management 2020, 256, 109940. 10.1016/j.jenvman.2019.109940. [DOI] [PubMed] [Google Scholar]
  322. Yang G.; Hu X.; Liang J.; Huang Q.; Dou J.; Tian J.; Deng F.; Liu M.; Zhang X.; Wei Y. Surface functionalization of MXene with chitosan through in-situ formation of polyimidazoles and its adsorption properties. J. Hazardous Mater. 2021, 419, 126220. 10.1016/j.jhazmat.2021.126220. [DOI] [PubMed] [Google Scholar]
  323. Shahzad A.; Rasool K.; Iqbal J.; Jang J.; Lim Y.; Kim B.; Oh J.-M.; Lee D. S. MXsorption of mercury: Exceptional reductive behavior of titanium carbide/carbonitride MXenes. Environ. Res. 2022, 205, 112532. 10.1016/j.envres.2021.112532. [DOI] [PubMed] [Google Scholar]
  324. Gu P.; Zhang S.; Zhang C.; Wang X.; Khan A.; Wen T.; Hu B.; Alsaedi A.; Hayat T.; Wang X. Two-dimensional MAX-derived titanate nanostructures for efficient removal of Pb (II). Dalton Transactions 2019, 48 (6), 2100–2107. 10.1039/C8DT04301A. [DOI] [PubMed] [Google Scholar]
  325. Fard A. K.; Mckay G.; Chamoun R.; Rhadfi T.; Preud’Homme H.; Atieh M. A. Barium removal from synthetic natural and produced water using MXene as two dimensional (2-D) nanosheet adsorbent. Chem. Eng. J. 2017, 317, 331–342. 10.1016/j.cej.2017.02.090. [DOI] [Google Scholar]
  326. Hu X.; Chen C.; Zhang D.; Xue Y. Kinetics, isotherm and chemical speciation analysis of Hg (II) adsorption over oxygen-containing MXene adsorbent. Chemosphere 2021, 278, 130206. 10.1016/j.chemosphere.2021.130206. [DOI] [PubMed] [Google Scholar]
  327. Mu W.; Du S.; Li X.; Yu Q.; Wei H.; Yang Y.; Peng S. Removal of radioactive palladium based on novel 2D titanium carbides. Chem. Eng. J. 2019, 358, 283–290. 10.1016/j.cej.2018.10.010. [DOI] [Google Scholar]
  328. Wang C.; Cheng R.; Hou P.-X.; Ma Y.; Majeed A.; Wang X.; Liu C. MXene-carbon nanotube hybrid membrane for robust recovery of Au from trace-level solution. ACS Appl. Mater. Interfaces 2020, 12 (38), 43032–43041. 10.1021/acsami.0c09310. [DOI] [PubMed] [Google Scholar]
  329. Qin Z.; Deng H.; Huang R.; Tong S. 3D MXene hybrid architectures for the cold-resistant, rapid and selective capture of precious metals from electronic waste and mineral. Chem. Eng. J. 2022, 428, 132493. 10.1016/j.cej.2021.132493. [DOI] [Google Scholar]
  330. Khan A. R.; Awan S. K.; Husnain S. M.; Abbas N.; Anjum D. H.; Abbas N.; Benaissa M.; Mirza C. R.; Mujtaba-ul-Hassan S.; Shahzad F. 3D flower like δ-MnO2/MXene nano-hybrids for the removal of hexavalent Cr from wastewater. Ceram. Int. 2021, 47, 25951. 10.1016/j.ceramint.2021.05.326. [DOI] [Google Scholar]
  331. Mu W.; Du S.; Yu Q.; Li X.; Wei H.; Yang Y. Improving barium ion adsorption on two-dimensional titanium carbide by surface modification. Dalton Transactions 2018, 47 (25), 8375–8381. 10.1039/C8DT00917A. [DOI] [PubMed] [Google Scholar]
  332. Khan A. R.; Husnain S. M.; Shahzad F.; Mujtaba-ul-Hassan S.; Mehmood M.; Ahmad J.; Mehran M. T.; Rahman S. Two-dimensional transition metal carbide (Ti3C2Tx) as an efficient adsorbent to remove cesium (Cs+). Dalton Transactions 2019, 48 (31), 11803–11812. 10.1039/C9DT01965K. [DOI] [PubMed] [Google Scholar]
  333. Liu P.; Hou J.; Zhang Y.; Li L.; Lu X.; Tang Z. Two-dimensional material membranes for critical separations. Inorg. Chem. Front. 2020, 7 (13), 2560–2581. 10.1039/D0QI00307G. [DOI] [Google Scholar]
  334. Shahzad A.; Moztahida M.; Tahir K.; Kim B.; Jeon H.; Ghani A. A.; Maile N.; Jang J.; Lee D. S. Highly effective prussian blue-coated MXene aerogel spheres for selective removal of cesium ions. J. Nucl. Mater. 2020, 539, 152277. 10.1016/j.jnucmat.2020.152277. [DOI] [Google Scholar]
  335. Zhang P.; Wang L.; Du K.; Wang S.; Huang Z.; Yuan L.; Li Z.; Wang H.; Zheng L.; Chai Z.; Shi W. Effective removal of U (VI) and Eu (III) by carboxyl functionalized MXene nanosheets. J. Hazardous Mater. 2020, 396, 122731. 10.1016/j.jhazmat.2020.122731. [DOI] [PubMed] [Google Scholar]
  336. Li S.; Wang L.; Peng J.; Zhai M.; Shi W. Efficient thorium (IV) removal by two-dimensional Ti2CTx MXene from aqueous solution. Chem. Eng. J. 2019, 366, 192–199. 10.1016/j.cej.2019.02.056. [DOI] [Google Scholar]
  337. Wang L.; Song H.; Yuan L.; Li Z.; Zhang Y.; Gibson J. K.; Zheng L.; Chai Z.; Shi W. Efficient U (VI) reduction and sequestration by Ti2CTx MXene. Environ. Sci. Technol. 2018, 52 (18), 10748–10756. 10.1021/acs.est.8b03711. [DOI] [PubMed] [Google Scholar]
  338. Zhang P.; Wang L.; Huang Z.; Yu J.; Li Z.; Deng H.; Yin T.; Yuan L.; Gibson J. K.; Mei L.; Zheng L.; Wang H.; Chai Z.; Shi W. Aryl diazonium-assisted amidoximation of MXene for boosting water stability and uranyl sequestration via electrochemical sorption. ACS Appl. Mater. Interfaces 2020, 12 (13), 15579–15587. 10.1021/acsami.0c00861. [DOI] [PubMed] [Google Scholar]
  339. Wang S.; Wang L.; Li Z.; Zhang P.; Du K.; Yuan L.; Ning S.; Wei Y.; Shi W. Highly efficient adsorption and immobilization of U (VI) from aqueous solution by alkalized MXene-supported nanoscale zero-valent iron. J. Hazardous Mater. 2021, 408, 124949. 10.1016/j.jhazmat.2020.124949. [DOI] [PubMed] [Google Scholar]
  340. Iwanade A.; Kasai N.; Hoshina H.; Ueki Y.; Saiki S.; Seko N. Hybrid grafted ion exchanger for decontamination of radioactive cesium in Fukushima Prefecture and other contaminated areas. J. Radioanalytical and Nuclear Chem. 2012, 293 (2), 703–709. 10.1007/s10967-012-1721-2. [DOI] [Google Scholar]
  341. Jun B.-M.; Jang M.; Park C. M.; Han J.; Yoon Y. Selective adsorption of Cs+ by MXene (Ti3C2Tx) from model low-level radioactive wastewater. Nuclear Eng. Technol. 2020, 52 (6), 1201–1207. 10.1016/j.net.2019.11.020. [DOI] [Google Scholar]
  342. Huang H.; Sha X.; Cui Y.; Sun S.; Huang H.; He Z.; Liu M.; Zhou N.; Zhang X.; Wei Y. Highly efficient removal of iodine ions using MXene-PDA-Ag2Ox composites synthesized by mussel-inspired chemistry. J. Colloid Interface Sci. 2020, 567, 190–201. 10.1016/j.jcis.2020.02.015. [DOI] [PubMed] [Google Scholar]
  343. Ibrahim A. G.; Rashad A. M.; Dincer I. Exergoeconomic analysis for cost optimization of a solar distillation system. Sol. Energy 2017, 151, 22–32. 10.1016/j.solener.2017.05.020. [DOI] [Google Scholar]
  344. He W.; Wang J. Feasibility study of energy storage by concentrating/desalinating water: Concentrated Water Energy Storage. Appl. nergy 2017, 185, 872–884. 10.1016/j.apenergy.2016.10.077. [DOI] [Google Scholar]
  345. Frank H.; Rahav E.; Bar-Zeev E. Short-term effects of SWRO desalination brine on benthic heterotrophic microbial communities. Desalination 2017, 417, 52–59. 10.1016/j.desal.2017.04.031. [DOI] [Google Scholar]
  346. Berdiyorov G. R.; Madjet M. E.; Mahmoud K. A. Ionic sieving through Ti3C2 (OH)2 MXene: First-principles calculations. Appl. Phys. Lett. 2016, 108 (11), 113110. 10.1063/1.4944393. [DOI] [Google Scholar]
  347. Ang E. Y.; Ng T. Y.; Yeo J.; Lin R.; Liu Z.; Geethalakshmi K. Investigations on different two-dimensional materials as slit membranes for enhanced desalination. J. Membr. Sci. 2020, 598, 117653. 10.1016/j.memsci.2019.117653. [DOI] [Google Scholar]
  348. Arshadi F.; Mohammad M.; Hosseini E.; Ahmadi H.; Asadnia M.; Orooji Y.; Korayem A. H.; Noorbakhsh A.; Razmjou A. The effect of D-spacing on the ion selectivity performance of MXene membrane. J. Membr. Sci. 2021, 639, 119752. 10.1016/j.memsci.2021.119752. [DOI] [Google Scholar]
  349. Jin Y.; Fan Y.; Meng X.; Li J.; Li C.; Sunarso J.; Yang N.; Meng B.; Zhang W. Modeling of hydrated cations transport through 2D MXene (Ti3C2Tx) membranes for water purification. J. Membr. Sci. 2021, 631, 119346. 10.1016/j.memsci.2021.119346. [DOI] [Google Scholar]
  350. Ma X.; Zhu X.; Huang C.; Fan J. Revealing the effects of terminal groups of MXene on the water desalination performance. J. Membr. Sci. 2022, 647, 120334. 10.1016/j.memsci.2022.120334. [DOI] [Google Scholar]
  351. Xu D.; Zhu X.; Luo X.; Guo Y.; Liu Y.; Yang L.; Tang X.; Li G.; Liang H. MXene nanosheet templated nanofiltration membranes toward ultrahigh water transport. Environ. Sci. Technol. 2021, 55 (2), 1270–1278. 10.1021/acs.est.0c06835. [DOI] [PubMed] [Google Scholar]
  352. Wang X.; Li Q.; Zhang J.; Huang H.; Wu S.; Yang Y. Novel thin-film reverse osmosis membrane with MXene Ti3C2Tx embedded in polyamide to enhance the water flux, anti-fouling and chlorine resistance for water desalination. J. Membr. Sci. 2020, 603, 118036. 10.1016/j.memsci.2020.118036. [DOI] [Google Scholar]
  353. Meidani K.; Cao Z.; Barati Farimani A. Titanium Carbide MXene for water desalination: A molecular dynamics study. ACS Appl. Nano Mater. 2021, 4, 6145. 10.1021/acsanm.1c00944. [DOI] [Google Scholar]
  354. Sun Y.; Li S.; Zhuang Y.; Liu G.; Xing W.; Jing W. Adjustable interlayer spacing of ultrathin MXene-derived membranes for ion rejection. J. Membr. Sci. 2019, 591, 117350. 10.1016/j.memsci.2019.117350. [DOI] [Google Scholar]
  355. Lu Z.; Wei Y.; Deng J.; Ding L.; Li Z.-K.; Wang H. Self-crosslinked MXene (Ti3C2Tx) membranes with good antiswelling property for monovalent metal ion exclusion. ACS Nano 2019, 13 (9), 10535–10544. 10.1021/acsnano.9b04612. [DOI] [PubMed] [Google Scholar]
  356. Yang G.; Xie Z.; Thornton A. W.; Doherty C. M.; Ding M.; Xu H.; Cran M.; Ng D.; Gray S. Ultrathin poly (vinyl alcohol)/MXene nanofilm composite membrane with facile intrusion-free construction for pervaporative separations. J. Membr. Sci. 2020, 614, 118490. 10.1016/j.memsci.2020.118490. [DOI] [Google Scholar]
  357. Zhu J.; Wang L.; Wang J.; Wang F.; Tian M.; Zheng S.; Shao N.; Wang L.; He M. Precisely tunable ion sieving with an Al13-Ti3C2Tx lamellar membrane by controlling interlayer spacing. ACS Nano 2020, 14, 15306. 10.1021/acsnano.0c05649. [DOI] [PubMed] [Google Scholar]
  358. Xue Q.; Zhang K. MXene nanocomposite nanofiltration membrane for low carbon and long-lasting desalination. J. Membr. Sci. 2021, 640, 119808. 10.1016/j.memsci.2021.119808. [DOI] [Google Scholar]
  359. Deng J.; Lu Z.; Ding L.; Li Z.-K.; Wei Y.; Caro J.; Wang H. Fast electrophoretic preparation of large-area two-dimensional Titanium Carbide membranes for ion sieving. Chem. Eng. J. 2021, 408, 127806. 10.1016/j.cej.2020.127806. [DOI] [Google Scholar]
  360. Xue Q.; Zhang K. The preparation of high-performance and stable MXene nanofiltration membranes with MXene embedded in the organic phase. Membranes 2022, 12 (1), 2. 10.3390/membranes12010002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  361. Shen Z.; Chen W.; Xu H.; Yang W.; Kong Q.; Wang A.; Ding M.; Shang J. Fabrication of a novel antifouling polysulfone membrane with in situ embedment of MXene nanosheets. Inter. J. Environ. Res. Public Health 2019, 16 (23), 4659. 10.3390/ijerph16234659. [DOI] [PMC free article] [PubMed] [Google Scholar]
  362. Xu M.; Zhao P.; Tang C. Y.; Yi X.; Wang X. Preparation of electrically enhanced forward osmosis (FO) membrane by two-dimensional MXenes for organic fouling mitigation. Chin. Chem. Lett. 2022, 33 (8), 3818–3822. 10.1016/j.cclet.2021.11.071. [DOI] [Google Scholar]
  363. Komlenic R. Rethinking the causes of membrane biofouling. Filtration Sep. 2010, 47 (5), 26–28. 10.1016/S0015-1882(10)70211-1. [DOI] [Google Scholar]
  364. Zha X.-J.; Zhao X.; Pu J.-H.; Tang L.-S.; Ke K.; Bao R.-Y.; Bai L.; Liu Z.-Y.; Yang M.-B.; Yang W. Flexible anti-biofouling MXene/cellulose fibrous membrane for sustainable solar-driven water purification. ACS Appl. Mater. Interfaces 2019, 11 (40), 36589–36597. 10.1021/acsami.9b10606. [DOI] [PubMed] [Google Scholar]
  365. Jastrzȩbska A. M.; Karwowska E.; Wojciechowski T.; Ziemkowska W.; Rozmysłowska A.; Chlubny L.; Olszyna A. The atomic structure of Ti2C and Ti3C2 MXenes is responsible for their antibacterial activity toward E. coli bacteria. J. Mater. Eng. Perfor.e 2019, 28 (3), 1272–1277. 10.1007/s11665-018-3223-z. [DOI] [Google Scholar]
  366. Seidi F.; Arabi Shamsabadi A.; Dadashi Firouzjaei M.; Elliott M.; Saeb M. R.; Huang Y.; Li C.; Xiao H.; Anasori B. MXenes antibacterial properties and applications: A review and perspective. Small 2023, 19, 2206716. 10.1002/smll.202206716. [DOI] [PubMed] [Google Scholar]
  367. Ding L.; Li L.; Liu Y.; Wu Y.; Lu Z.; Deng J.; Wei Y.; Caro J.; Wang H. Effective ion sieving with Ti3C2Tx MXene membranes for production of drinking water from seawater. Nature Sustainability 2020, 3 (4), 296–302. 10.1038/s41893-020-0474-0. [DOI] [Google Scholar]
  368. Zhao J.; Yang Y.; Yang C.; Tian Y.; Han Y.; Liu J.; Yin X.; Que W. A hydrophobic surface enabled salt-blocking 2D Ti3C2 MXene membrane for efficient and stable solar desalination. J. Mater. Chem. A 2018, 6 (33), 16196–16204. 10.1039/C8TA05569F. [DOI] [Google Scholar]
  369. Zhang Q.; Yi G.; Fu Z.; Yu H.; Chen S.; Quan X. Vertically aligned janus MXene-based aerogels for solar desalination with high efficiency and salt resistance. ACS Nano 2019, 13 (11), 13196–13207. 10.1021/acsnano.9b06180. [DOI] [PubMed] [Google Scholar]
  370. Wang Y.; Qi Q.; Fan J.; Wang W.; Yu D. Simple and robust MXene/carbon nanotubes/cotton fabrics for textile wastewater purification via solar-driven interfacial water evaporation. Sep. Purif. Technol. 2021, 254, 117615. 10.1016/j.seppur.2020.117615. [DOI] [Google Scholar]
  371. Zhang T.; Zhang L.-Z. Development of a MXene-based membrane with excellent anti-fouling for air humidification-dehumidification type desalination. J. Membr. Sci. 2022, 641, 119907. 10.1016/j.memsci.2021.119907. [DOI] [Google Scholar]
  372. Wu X.; Ding M.; Xu H.; Yang W.; Zhang K.; Tian H.; Wang H.; Xie Z. Scalable Ti3C2Tx MXene interlayered orward osmosis membranes for enhanced water purification and organic solvent recovery. ACS Nano 2020, 14 (7), 9125–9135. 10.1021/acsnano.0c04471. [DOI] [PubMed] [Google Scholar]
  373. Ding M.; Xu H.; Chen W.; Yang G.; Kong Q.; Ng D.; Lin T.; Xie Z. 2D laminar maleic acid-crosslinked MXene membrane with tunable nanochannels for efficient and stable pervaporation desalination. J. Membr. Sci. 2020, 600, 117871. 10.1016/j.memsci.2020.117871. [DOI] [Google Scholar]
  374. Wang J.; Zhang Z.; Zhu J.; Tian M.; Zheng S.; Wang F.; Wang X.; Wang L. Ion sieving by a two-dimensional Ti3C2Tx alginate lamellar membrane with stable interlayer spacing. Nat. Commun. 2020, 11 (1), 3540. 10.1038/s41467-020-17373-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
  375. Climate Change: The IPCC Scientific Assessment; Houghton J. T., Jenkins G. J., Ephraums J. J., Eds.; Cambridge University Press, 1990.
  376. Ming X.; Guo A.; Zhang Q.; Guo Z.; Yu F.; Hou B.; Wang Y.; Homewood K. P.; Wang X. 3D macroscopic graphene Oxide/MXene architectures for multifunctional water purification. Carbon 2020, 167, 285. 10.1016/j.carbon.2020.06.023. [DOI] [Google Scholar]
  377. Wu Q.; Su W.; Li Q.; Tao Y.; Li H. Enabling continuous and improved solar-driven atmospheric water harvesting with Ti3C2-incorporated metal-organic framework monoliths. ACS Appl. Mater. Interfaces 2021, 13 (32), 38906–38915. 10.1021/acsami.1c10536. [DOI] [PubMed] [Google Scholar]
  378. Zhao X.; Zha X.-J.; Pu J.-H.; Bai L.; Bao R.-Y.; Liu Z.-Y.; Yang M.-B.; Yang W. Macroporous three-dimensional MXene architectures for highly efficient solar steam generation. J. Mater. Chem. A 2019, 7 (17), 10446–10455. 10.1039/C9TA00176J. [DOI] [Google Scholar]
  379. Fan X.; Yang Y.; Shi X.; Liu Y.; Li H.; Liang J.; Chen Y. A MXene-based hierarchical design enabling highly efficient and stable solar-water desalination with good salt resistance. Adv. Funct. Mater. 2020, 30 (52), 2007110. 10.1002/adfm.202007110. [DOI] [Google Scholar]
  380. Shen X.; Xiong Y.; Hai R.; Yu F.; Ma J. All-MXene-based integrated membrane electrode constructed using Ti3C2Tx as an intercalating agent for high-performance desalination. Environ. Sci. Technol. 2020, 54 (7), 4554–4563. 10.1021/acs.est.9b05759. [DOI] [PubMed] [Google Scholar]
  381. Chen Z.; Xu X.; Ding Z.; Wang K.; Sun X.; Lu T.; Konarova M.; Eguchi M.; Shapter J. G.; Pan L.; Yamauchi Y. Ti3C2 MXenes-derived NaTi2(PO4)3/MXene nanohybrid for fast and efficient hybrid capacitive deionization performance. Chem. Eng. J. 2021, 407, 127148. 10.1016/j.cej.2020.127148. [DOI] [Google Scholar]
  382. Liu G.; Shen J.; Liu Q.; Liu G.; Xiong J.; Yang J.; Jin W. Ultrathin two-dimensional MXene membrane for pervaporation desalination. J. Membr. Scie. 2018, 548, 548–558. 10.1016/j.memsci.2017.11.065. [DOI] [Google Scholar]
  383. Liu G.; Guo Y.; Meng B.; Wang Z.; Liu G.; Jin W. Two-dimensional MXene hollow fiber membrane for divalent ions exclusion from water. Chinese J. Chem. Engi. 2022, 41, 260–266. 10.1016/j.cjche.2021.09.022. [DOI] [Google Scholar]
  384. Sun P.-F.; Yang Z.; Song X.; Lee J. H.; Tang C. Y.; Park H.-D., Interlayered forward osmosis membranes with Ti3C2Tx MXene and carbon nanotubes for enhanced municipal wastewater concentration. Environ. Sci. Technol. 2021.55 (19), 13219–13230. 10.1021/acs.est.1c01968 [DOI] [PubMed] [Google Scholar]
  385. Ren J.; Zhu Z.; Qiu Y.; Yu F.; Zhou T.; Ma J.; Zhao J. J. C. Enhanced adsorption performance of alginate/MXene/CoFe2O4 for antibiotic and heavy metal under rotating magnetic field. Chemosphere 2021, 284, 131284. 10.1016/j.chemosphere.2021.131284. [DOI] [PubMed] [Google Scholar]
  386. Miri-Jahromi A.; Didandeh M.; Shekarsokhan S. Capability of MXene 2D material as an amoxicillin, ampicillin, and cloxacillin adsorbent in wastewater. J. Mol. Liq. 2022, 351, 118545. 10.1016/j.molliq.2022.118545. [DOI] [Google Scholar]
  387. Gao H.; Chen Y.. Hybrid dimensional MXene/CNC framework-regulated nanofiltration membrane with high separation performance. ACS ES&T Water, 2023, 3, (7), , 1767–1777 10.1021/acsestwater.2c00231. [DOI] [Google Scholar]
  388. Rabb H.; Lee K.; Parikh C. R. Beyond kidney dialysis and transplantation: what’s on the horizon?. J. Clin. Invest. 2022, 132 (7), e159308 10.1172/JCI159308. [DOI] [PMC free article] [PubMed] [Google Scholar]
  389. Shuck C. E.; Gogotsi Y. Taking MXenes from the lab to commercial products. Chem. Eng. J. 2020, 401, 125786. 10.1016/j.cej.2020.125786. [DOI] [Google Scholar]
  390. Arifutzzaman A.; Musa I. N.; Aroua M. K.; Saidur R. MXene based activated carbon novel nano-sandwich for efficient CO2 adsorption in fixed-bed column. J. CO2 Utilization 2023, 68, 102353. 10.1016/j.jcou.2022.102353. [DOI] [Google Scholar]
  391. Chen L.; Dai X.; Feng W.; Chen Y. Biomedical applications of MXenes: From nanomedicine to biomaterials. Accounts of Mater. Res. 2022, 3 (8), 785–798. 10.1021/accountsmr.2c00025. [DOI] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

ao3c01182_si_001.pdf (233.2KB, pdf)

Articles from ACS Omega are provided here courtesy of American Chemical Society

RESOURCES