Next Article in Journal
Laboratory High-Contrast X-ray Microscopy of Copper Nanostructures Enabled by a Liquid-Metal-Jet X-ray Source
Previous Article in Journal
Feature-Assisted Machine Learning for Predicting Band Gaps of Binary Semiconductors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

MXene-Based Chemo-Sensors and Other Sensing Devices

by
Ilya Navitski
1,2,
Agne Ramanaviciute
2,
Simonas Ramanavicius
3,
Maksym Pogorielov
4,5,* and
Arunas Ramanavicius
2,*
1
Department of Nanotechnology, State Research Institute Center for Physical Sciences and Technology (FTMC), Sauletekio av. 3, LT-10257 Vilnius, Lithuania
2
Department of Physical Chemistry, Faculty of Chemistry and Geosciences, Institute of Chemistry, Vilnius University, Naugarduko 24, LT-03225 Vilnius, Lithuania
3
Department of Organic Chemistry, State Research Institute Center for Physical Sciences and Technology, Saulėtekio av. 3, LT-10257 Vilnius, Lithuania
4
Biomedical Research Centre, Sumy State University, 2, Kharkivska Str., 40007 Sumy, Ukraine
5
Institute of Atomic Physics and Spectroscopy, University of Latvia, 3 Jelgavas St., LV-1004 Riga, Latvia
*
Authors to whom correspondence should be addressed.
Nanomaterials 2024, 14(5), 447; https://doi.org/10.3390/nano14050447
Submission received: 10 January 2024 / Revised: 15 February 2024 / Accepted: 24 February 2024 / Published: 28 February 2024
(This article belongs to the Section 2D and Carbon Nanomaterials)

Abstract

:
MXenes have received worldwide attention across various scientific and technological fields since the first report of the synthesis of Ti3C2 nanostructures in 2011. The unique characteristics of MXenes, such as superior mechanical strength and flexibility, liquid-phase processability, tunable surface functionality, high electrical conductivity, and the ability to customize their properties, have led to the widespread development and exploration of their applications in energy storage, electronics, biomedicine, catalysis, and environmental technologies. The significant growth in publications related to MXenes over the past decade highlights the extensive research interest in this material. One area that has a great potential for improvement through the integration of MXenes is sensor design. Strain sensors, temperature sensors, pressure sensors, biosensors (both optical and electrochemical), gas sensors, and environmental pollution sensors targeted at volatile organic compounds (VOCs) could all gain numerous improvements from the inclusion of MXenes. This report delves into the current research landscape, exploring the advancements in MXene-based chemo-sensor technologies and examining potential future applications across diverse sensor types.

1. Introduction

The family of two-dimensional (2D) transition metal carbides, nitrides, and carbonitrides called MXenes has significantly grown since the first report of the synthesis of Ti3C2 in 2011 [1]. In recent years, MXenes have greatly impacted many fields, including energy storage [2,3], electronics [4,5], biomedicine, catalysis [6,7], environmental technologies, and many other technological areas. This widespread attention arises from the exceptional set of properties attributed to MXenes, which include superior mechanical strength and flexibility [8], liquid-phase processability [9], tunable surface functionality [10], high electrical conductivity [11], and the ability to customize their properties. This array of highly functional and desirable material qualities inevitably led to the production of thousands of publications during the previous ten years [12]. MXenes are characterized by the general chemical formula Mn+1XnTx, where M is one of the transition metals, n varies from 1 to 4 [13], X indicates C and/or N, and Tx denotes the surface terminations, which include –F, –O, and –OH as well as other chalcogens (–S, –Se, –Te), halogens (–Cl, –Br, –I), or imido (=NH) groups [14,15]. The suffix “ene” underscores their structural similarity to graphene and to other 2D materials [16]. Figure 1 shows the elements of the known MAX phases and MXenes.
One of the many reasons why MXenes are used in a variety of industries is the diverse chemistry associated with the element “M”. For instance, Vanadium (V)-based MXenes have a low ion diffusion barrier and some characteristics required for energy storage devices [17], Molybdenum (Mo)-based MXenes have great potential for application in the fields of electrocatalysis and thermoelectricity [18], and Niobium (Nb)-based MXenes exhibit magnetic phase change due to their diamagnetic nature [19]. However, Ti3C2Tx remains the most studied and commonly used MXene.
Etching the “A” element from the MAX phase, which is represented by the formula Mn+1AXn, is a common step in the MXene synthesis process [20,21]. HF etching is the most popular etching technique, which is applied for the preparation of MXenes from a particular MAX phase [22,23,24,25]. However, the primary issue with this approach is the corrosive and hazardous nature of HF, which raises concerns about potential harm to the environment and public health. To overcome this issue, scientists have recently developed a number of innovative etching methods. Alkali etching [26,27], molten salt etching [28,29], electrochemical etching [30,31], iodine etching [32] and UV-induced etching [33] are some of the many new approaches for synthesizing MXenes. After the etching, multilayered MXenes are created and can already be utilized in some applications (such as lubrication [34]). However, in many other cases, MXenes must be intercalated and delaminated in order to optimize the chemical characteristics of these materials. DMSO [35], bases such TBAOH [36] and isopropylamine [37], and a variety of cations (Li+ [13], Na+ [38], K+ and others [23]) are used as intercalants during the delamination of MXphase and the formation of 2D MXene sheets. Figure 2 shows a schematic depiction of Nb2CTz MXene synthesis, depicting all major steps necessary for MXene formation.
MAX phases are usually produced by combining laboratory-grade elemental powders (Ti, Al, and C powders for Ti3AlC2) and heating the mixture in an inert atmosphere to extreme temperatures; as such, this makes the production of the precursor relatively expensive compared to other two-dimensional substances (in comparison to obtaining graphene from graphite). Consequently, MXene production is comparatively expensive [40]. As a result, the developments in some areas may be hindered as it becomes economically unfeasible. Some research groups were able to synthesize MAX phases using inexpensive components (such as TiO2, Al from scrap metal, and C recovered from tires) and then utilize the obtained product to create MXenes. The comparison of properties between MXenes made from pure MAX phases and the ones from the inexpensive precursors exhibited similarities that are sufficient for certain applications. This shows that much more affordable MXene synthesis can be achieved and successfully applied as well [40,41]. However, for many other fields, especially ones that require highly pure MXenes, further improvements in synthesis are still needed.
MXenes are among the greatest materials that are used for sensing applications due to their various properties and the ability to tune them for users’ purposes by changing “M”, “Tx” or the number of layers [42]. The addition of MXenes to sensor designs can improve strain sensors, temperature sensors, pressure and gas sensors and chemical sensors. Therefore, the aim of this report is to examine the state of research and potential future applications of some MXene-based sensors. In this paper, we discuss the benefits that MXenes can provide to different types of sensors, examine the mechanisms or reactions behind the sensing, and highlight the most unique and interesting works of the past decade. The main focus will be placed on outlining fairly recent (2022–2023) and, in our opinion, the most exciting advancements in the aforementioned sensor types. Moreover, the review will focus not only on the most popular Ti3C2Tx sensors but also on other novel MXene applications. It must be noted that biosensing applications of MXenes will not be discussed due to the enormously wide scope of this particular topic.

2. MXene-Based Strain Sensors

The rapid advancement of flexible electronics-based devices in recent years has led to their widespread adoption across numerous industries. These devices, characterized by their bendable and conformable nature, have unlocked new possibilities and applications in fields such as sports, communications, medicine, and beyond. The versatility of flexible electronic devices has spurred innovation and opened new avenues for interdisciplinary research. As technology continues to advance, it is likely that flexible electronics will play an even more significant role in shaping the future of various industries [43]. For wearable applications like human motion detection, strain sensors need to meet several baseline requirements, including high stretchability, flexibility, sensitivity, and durability. To achieve this, scientists need to carefully design and coordinate different materials. One method is to insert zero-dimensional nanoparticles, such as metal nanoparticles [44,45], into an elastic substrate to create a conductive network, which can help in achieving high sensitivity. However, steady sensing may be hampered by the instability of the conductive network. On the other hand, one-dimensional nanowires like CNTs [46,47] and AgNWs [48] can be used to create a conductive network inside or on top of the elastic substrate. Because of the relatively stable conductive channels provided by this construction, resistance changes are less likely to occur due to deformation, yet this comes at a cost of decreased sensitivity [49]. Contrary to most other two-dimensional materials, such as graphene, MXenes have a large initial metallic conductivity. Since the crack propagation mechanism (Figure 3) dominates the sensing, MXene-based strain sensors exhibit increased resistance during stretching [50]. A narrow sensing range is the result of MXene sheets’ tendency to lose inter-sheet connections during the tensile process [51,52,53]. Combining conductive materials can enhance the overall conductivity and performance by taking advantage of different qualities of the materials.
One of the profound features of MXenes in strain sensor design is the ability to tune properties, such as the sensitivity and working area. Ti3C2Tx An MXene/carbon nanotube (CNT) composite was conceptually designed and included in strain sensors by Yichen Cai et al. CNT crossing and Ti3C2Tx nano-stacks were combined in a weaving architecture. The resulting strain sensor had an ultralow detection limit of 0.1% strain, tunable sensitivity (gauge factor up to 772.6), and customizable sensing range, mostly depending on the concentration ratio between CNT and MXene (30% to 130% strain) [52]. Wireless Ti3C2Tx MXene strain sensing devices were created by Haitao Yang et al. Within several user-specified high-strain working windows (from 130% to 900%), the wireless MXene sensor system may simultaneously achieve an ultrahigh sensitivity (with gauge factor up to 14,000) and reliable linearity (R2 = 0.99). The authors used a single database channel by the wireless system to track all of the exoskeleton actuations [55].
Due to the hydrophilicity of MXenes, based on surface terminations, they can be implemented in hydrogels. Figure 4 shows a schematic illustration of the mechanism of MXene-based hydrogel strain sensors. Hui Liao et al. immersed MXene nanocomposite hydrogel (MNH) into ethylene glycol to create an anti-freezing, self-healing, and conductive MNOH. According to the authors, MNOH may be incorporated into a wearable strain sensor with a reasonably wide strain range (up to 350% strain) and a high gauge factor of 44.85 under very low temperatures to monitor human biological activity [56].
In 2023, Chunqing Yang et al. created a sandwich-structured flexible strain sensor using MXene, polypyrrole, hydroxyethyl cellulose (MXene/PPy/HEC), and a flexible substrate made of PDMS. The signals of handwritten Chinese, Arabic, and English words measured by the sensor exhibited distinctive properties. Different handwritten characters were successfully recognized by the authors using machine learning techniques, with a recognition accuracy above 96% [57].
In conclusion, MXenes, especially Ti3C2Tx, have been successfully applied in strain sensor design due to their superior electrochemical properties, metallic conductivity, and hydrophilicity. They are able to provide high sensitivity to strain sensors at the cost of narrowing the working area. More examples of MXene usage in strain sensors can be found in Table 1.

3. MXene-Based Pressure Sensors

Flexible pressure sensors have become integral components in wearable electronic devices, offering a wide range of applications in areas such as health monitoring, human-machine interfaces, and robotics. In the past decade, numerous studies have focused on exploring various mechanisms employed in flexible pressure sensors, including piezoresistive, piezocapacitive, piezoelectric, and triboelectric principles. These mechanisms enable these sensors to detect and respond to pressure changes in different ways. The Piezoresistive mechanism involves changes in the electrical resistance as a response to mechanical deformation. Flexible pressure sensors using piezoresistive materials, such as certain polymers or carbon-based composites, can detect pressure by measuring variations in resistance [69]. The Piezocapacitive mechanism is mostly based on monitoring the capacitance changes, which are generated in response to pressure and are the basis of piezocapacitive sensors. These sensors typically consist of flexible materials with varying capacitance, and the pressure-induced deformations lead to changes in the capacitance that can be measured [70]. The Piezoelectric mechanism is realized in materials with piezoelectric properties to generate an electric charge in response to mechanical stress. Flexible pressure sensors based on the piezoelectric mechanism use such materials to convert pressure-induced deformations into electrical signals [71]. Triboelectric sensors generate electric charges through the triboelectric effect, which involves the transfer of electrons between materials in contact. Flexible pressure sensors based on triboelectric principles can detect pressure changes through the generation of triboelectric charges [72]. All these mechanisms offer different advantages and are suitable for various applications based on the specific requirements of the wearable device. The choice of the mechanism depends on factors such as the required sensitivity, response time, and the overall design of the sensor. The ongoing research and development in this field aims to improve the performance, sensitivity, and flexibility of these pressure sensors, contributing to the continued evolution of wearable electronic devices and their integration into diverse applications [73,74,75,76,77]. Piezoresistive pressure sensors stand out among the others due to the ease of device assembly and the comparatively low requirements for signal collection and readout [78]. The structure of all types of pressure sensors consists of flexible substrates, active materials, and conductive electrodes despite the fact that different sensing mechanisms are used. High sensitivity, a wide detection range, rapid response time, low detection limit, and good linearity are required for the perfect flexible pressure sensor [79,80,81]. However, because of the constrained sensing capabilities and intricate manufacturing process, the practical application of pressure sensors is still challenging and expensive. A brand-new frontier in the development of pressure sensors has been opened by the discovery of MXenes and especially Ti3C2Tx-based ones. The surface of Ti3C2Tx, in contrast to other 2D materials like graphene and black phosphorus, has a great number of attached exchangeable functional groups, which gives them exceptional water dispersion and plasticity. They may also be mixed with other materials to create a variety of multifunctional arrangements and micro-structures. Ti3C2Tx nanosheets are noteworthy due to their advantageous mechanical properties, high conductivity, and pressure-adjustable layer spacing. These features offer a solid foundation for the design of microstructures, which are required for pressure sensors, tactile sensors, and the formation of force-sensing layers [82]. Figure 5 shows the mechanism behind MXene-based pressure sensors.
An MXene/cotton fabric (MCF)-based pressure sensor was developed in 2020 by Yanjun Zheng et al., utilizing the flexibility and three-dimensional porous structure of cotton fabric as well as the unique sandwich construction of the sensor. It displayed advanced stability, long-term durability, high sensitivity (5.30 kPa−1 in the pressure range of 0–1.30 kPa), a wide sensing range (0–160 kPa), and a quick response/recovery time (50 ms/20 ms) [83]. A flexible, extremely sensitive, and waterproof sponge pressure sensor was described by Zhenyuan Xu et al. SiO2, MXene, and NH2-CNTs were put together to create the sensor on a melamine sponge core. Good sensitivity (10.8 kPa−1), a quick response and recovery time (40 ms and 60 ms, respectively), a broad detection range (30 kPa), and a remarkably low detection limit (4.6 Pa) were all features of this waterproof sponge pressure sensor [84]. Lin Wang et al. created an MXene/ANFs composite aerogel through a controlled vacuum filtering procedure followed by freeze-drying. The aerogel sensor exhibited a broad detection range (2.0–80.0% compression strain), good sensitivity (128 kPa−1), and a low detection limit (100 Pa) [85]. In 2022, Tingting Yin et al. developed MXene/polyaniline (PANI) foam utilizing a steam-induced foaming technique to produce a 3D porous structure. This foam-based sensor demonstrated exceptional wear resistance (10,000 cycles), quick response and recovery times (106/95 ms), and remarkable sensitivity (690.91 kPa−1) [86]. In 2022, using electrospinning technology, Xiyao Fu et al. created a pressure sensor using the MXene/ZIF-67/polyacrylonitrile (PAN) nanofiber film. The film-based device exhibited a broad operating range (0–100 kPa), good sensitivity (62.8 kPa−1), robust mechanical stability (over 10,000 cycles), and a quick response/recovery time (10/8 ms) [81].
Sound, which is essentially a ‘vibration’ of molecules, may travel through a gaseous/liquid/solid medium and can be transmitted as mechanical waves based on the variation of pressure and motion [87]. Exploring communication systems that use other modes of signals instead of sound is important as various human infirmities can restrict or impede the capacity to talk [88]. Compared to layered graphene or other materials, MXene nanoflakes with considerably varied interlayer spacing may produce a higher mechanical response, making them an appropriate mechanical sensing material for developing piezoresistive devices for sound registration [89]. Yangyang Pei et al. used foamed MXenes for their sound sensor. This method enabled this group of scientists to tackle the MXenes’ self-stacking problem, a common limitation that restricts material applications, by reducing the total energy. The Ti3C2Tx-based sensor demonstrated exceptional durability over 5000 cycles, a low detection limit of 1 Pa, a quick reaction time of 132 ms, and an impressive sensitivity of 102.89 kPa−1 for pressures less than 1 kPa. The sensor demonstrated the ability to perceive air pressure waves as an eardrum, distinguishing between diverse natural sounds and identifying human voices [90]. Guang-Yang Gou and others reported on an ultrasensitive intelligent artificial eardrum based on MXene and PDMS-PE. This artificial eardrum demonstrated an unprecedented sensitivity of 62 kPa−1 and a very low detection limit of 0.1 Pa. The accompanying machine-learning algorithm for real-time voice classification developed by the authors showcases a high accuracy of recognition of 96.4 and 95% [91].
The aforementioned examples of MXenes in pressure sensing applications highlight how its unique properties contribute to significant advancement in this field. High conductivity, hydrophilicity, abundance of changeable surface functional groups, customizable layer spacing and several other key characteristics of MXenes make this material particularly well-suited for pressure sensor development [92]. Most importantly, MXenes typically exhibit high electrical conductivity, which is essential for efficient signal transmission in pressure sensors. This property allows for accurate and rapid detection of pressure changes [69]. Moreover, MXenes are hydrophilic, meaning they have a strong affinity for water molecules. This property can be advantageous in pressure sensors designed to operate in humid or wet conditions as it helps maintain sensor performance in such environments [84]. MXenes’ surfaces contain an abundance of changeable functional surface groups that can be modified or altered. This provides researchers and engineers with the flexibility to tailor the surface chemistry of MXene-based pressure sensors, optimizing their sensitivity and selectivity to specific stimuli [93]. The layered structure of MXenes allows for the adjustment of interlayer spacings. Customizable layer-spacing-based tunability is valuable in pressure sensors as it enables customization to enhance the sensitivity and responsiveness to pressure variations [94].
Aerogels, hydrogels, and foams with a highly organized porous microstructure are included in the force-sensitive layers, which helps to increase the sensitivity and pressure response range. The pressure sensor’s response time and sensitivity can be increased by improving the contact between the force-sensitive layer and the electrode layer in fiber composites and fabrics made from MXenes. As a result, there is a lot of development potential for pressure sensors based on MXenes. More examples of MXenes used in pressure sensors are presented in Table 2.
In conclusion, MXenes already play a part in the current advancements in pressure sensing technologies as they exhibit highly desired and useful qualities for sensor design including high conductivity, hydrophilicity, abundance of changeable surface functional groups, and customizable layer spacing. These inherent characteristics of MXenes contribute to its versatility in pressure sensing applications, making it a promising material for further development of advanced and high-performance pressure sensors. The inclusion of MXenes could benefit a diverse range of pressure sensors including those used in touchscreens, wearable devices, and structural health monitoring. Ongoing research is likely to further explore and optimize MXene-based pressure sensors for enhanced performance and expanded applications.

4. MXene-Based Temperature Sensors

Temperature sensors are used in various industries: from kitchen appliances to patient monitoring and reaction temperature monitoring. The sensing abilities of most temperature sensors are based on the variation in electrical properties of materials, which are sensitive to temperature. The sensing mechanisms behind MXene-based temperature and strain sensors are quite similar and are dominated by the so-called ‘crack mechanism’. Such temperature sensors consist of a temperature-sensitive substrate that expands with increasing temperature with an MXene-based layer deposited on the top. When the temperature rises, the thermal expansion of temperature-sensitive material places tensile stress on the MXene-based layer, causing it to displace and ‘crack’, which results in an increase in electrical resistance; the resistance gradually returns to its initial value when the temperature-sensitive substrate shrinks (Figure 6) [100].
Lianjia Zhao et al. developed a flexible temperature sensor using sodium alginate hydrogel as a heat-sensitive coating layer and Ti3C2Tx MXene as a conducting skeleton. With a sensitivity up to 3244% °C−1, the sensor displays an accurate and consistent temperature response over the whole range (from −20 to 100 °C). The sensitivity of the obtained sensor was higher compared to a metal oxide sensor, CNT sensor and PEDOT:PSS/GO sensor. It is worth noting that the linear range of the sensor was lower compared to that of the CNT sensor [101]. Zeyi Wang et al. created a Fe2+/Ti2CTx/CA-PAM hydrogel-based frost- and dehydration-resistant temperature sensor. The sensor demonstrated good linearity (R2 = 0.998), sensitivity (−1.07% °C−1) across a wide operating range (−10 to 60 °C), high resolution (0.1 °C), and good repeatability. Compared to most other fire detection and warning sensor systems (based on nanoparticles, graphene oxide and metal oxides), the developed sensor demonstrated a greater reproducibility of fire detection and the benefit of the absence of a need for external power to function. The sensor’s potential in healthcare monitoring, electronic skins, and intelligent robotics was also demonstrated by its integration into a wireless system for body temperature monitoring [102]. By dip-coating Ti3C2Tx MXene material on a polyester fiber substrate and depositing silver nanoparticles, Hailian Liu et al. were able to create a high-performance wearable sensor. The fiber sensor had great temperature sensitivity (1.0436% °C−1–3.4938% °C−1), in addition to responding well to stimuli of temperature, pressure/strain signals [103]. Ti3C2Tx nanoparticle–lamella hybrid networks were incorporated into PDMS substrates by Zherui Cao et al. to create a flexible temperature sensor. The sensor had a high sensitivity (up to 986% °C−1) and a broad response range (up to 140 °C), quick response time—below 7.0 s, high accuracy (0.1 °C), and good dependability and durability of >100 cycles, which could meet the needs of e-skin for temperature monitoring [100]. Using graphene and Ti3C2Tx MXene nanoinks, Mortaza Saeidi-Javash et al. presented an aerosol-based flexible bimodal sensor. The printed temperature sensor exhibited a competitive ’thermo-power’ output of 53.6 V/°C with extremely high precision and stability. It should be noted that even after 1000 bend cycles, the printed MXene-based sensor still exhibited outstanding flexibility with barely any degradations [104].
To create a photothermal optical sensor (PHOS), Yan Zuo et al. added Ti3C2Tx into the Mach–Zehnder interferometer (MZI) framework. An efficiency as high as 0.19 π·mW−1·mm−1 under 980 nm laser pumping was attained as a result of the Ti3C2Tx’s effective photothermal conversion, and an even higher efficiency was seen when exposed to red light. The designed scheme’s response time was measured to be 23.4 s [105]. Si Chen et al. have demonstrated an all-fiber temperature sensor based on MXene V2C. When the temperature range was ~25–70 °C, the corresponding transmission light intensity variation was linear, with a maximum normalized sensing efficiency of 2.21 dB·°C−1·mm−1. Figure 7 shows experimental setup for measuring temperature, used in the latter experiment. The runway structure used in the sensor device significantly enhanced the interaction length between light and the V2C, thereby improving the overall sensing efficiency of the MXene-based sensor. Compared to other fiber sensors based on graphene or metal oxides, this sensor has superior sensitivity but a reduced linear range [106]. An MXene V2C integrated runway-type micro-fiber knot resonator (MKR)-based temperature sensor was presented by Qing Wu et al. in 2023. A maximum normalized temperature sensing efficiency of 1.65 dB °C−1·mm−1 was attained. The highest sensing efficiency (~0.33 dB·°C−1) was 2.7 times greater than the naked runway-type MKR’s (~0.09 dB·°C−1) [107].
MXene-based temperature sensors encounter specific challenges that warrant attention. Firstly, despite exhibiting superior qualities compared to some alternative materials, MXene-based sensors face accuracy issues when utilized for temperature measurements. Secondly, the current manufacturing methods for temperature sensors, which often involve distinct processing steps and intricate fabrication designs, may compromise the comfort of wearable devices. Overcoming these challenges is crucial for optimizing the performance and applicability of MXene-based temperature sensors in various applications [108]. In general, compared to other materials, MXenes usually provide advanced sensitivity but lower stability and smaller linear range.

5. MXene-Based Humidity Sensors

Monitoring and regulating humidity have become increasingly crucial in various industries, such as agriculture, textile technology, and food storage. The significance of controlling humidity levels stems from its direct impact on product quality, preservation, and overall operational efficiency within these sectors. As a result, precise humidity management is now integral to maintaining optimal conditions for crop growth, ensuring the quality of textiles, and preserving the freshness and safety of food products during storage and transportation. The evolving importance of humidity control underscores its pivotal role in enhancing performance and outcomes across diverse industrial applications [109,110,111]. For these applications, it is crucial to detect relative humidity (RH) with good accuracy, sensitivity, and reliability. Ti3C2Tx MXene possesses terminal groups that provide ample hydrophilic active sites for water adsorption and intercalation [112,113], in contrast to semiconductor metal oxides where water-molecule-dependent surface conductivity dominates the sensing mechanism [114,115]. The low electrical conductivity of water and the lengthening of the layer-to-layer distance will result in an increase in Ti3C2Tx resistance following the intercalation of water molecules in the interlayers between 2D MXene sheets [113,116]. Thus, Ti3C2Tx MXene provides a sensing mechanism that is suitable for RH sensing [117].
Layer-by-layer assembly was used by Hyosung An et al. to create MXene/polyelectrolyte (poly(diallyldimethylammonium chloride)) multilayer films, which displayed extremely quick response (110 ms) and recovery (220 ms) times in the RH range of 20–40%. The sensor was successfully employed to monitor human respiration in real time, though the authors noted that direct contact with the electrolyte can lead to skin irritation. It was also shown how increasing the space between sheets increased the tunable resistance by intercalating water molecules into multilayers [118]. Yang Lu et al. described the fabrication of an impedance-type humidity sensor based on Ti3C2Tx/g-C3N4 nanomaterials. They investigated the process of humidity sensing using complex impedance spectroscopy, finding that the major conductive ion is H3O+ at RH between 11% and 43%. Physisorbed water layers began to grow on top of the initial chemisorbed water layer as the Grotthuss chain reaction (H2O + H3O+ → H3O+ + H2O) governed charge transportation at this stage, where the major charge carriers H+ move freely. As a result, the impendence dropped as the humidity increased, which formed the groundwork for the created sensor. At a range of 11–97% RH, the humidity sensor demonstrated exceptional repeatability, fast response and recovery times (4.8 and 8.9 s, respectively), and neglectable hysteresis. The response was significantly stronger than that of most other impedimetric sensors based on other materials (namely, Ag or LiCl). The authors showed how it might be used to detect water evaporation and monitor human breath [119]. Mimi Han et al. assembled MXene nanosheets into a layered TOCNF/MXene nanocomposite film utilizing vacuum-assisted filtration and (TEMPO)-oxidized cellulose nanofibers (TOCNFs) as a template. The swelling and contraction of channels between MXene interlayers caused by adsorbed H2O and the swelling of TOCNFs were both components of the sensor’s humidity-sensing mechanism. The TOCNF/MXene sensor displayed exceptional bending and folding durability (up to 50 cycles), long-term stability, and a maximum response (−∆I/I0) of 90% under 97% RH. The authors demonstrated the sensor applications in smart wearable electronics by dynamically monitoring human breath, skin, and fingertip humidity [120].
By fluoride doping, Runlong Li et al. improved the MXene quartz crystal microbalance-based humidity sensor. They demonstrated that the resulting sensor had 12.8 Hz/% RH sensitivity compared to 10.2% for the fluoride-free sensor that was made using the same procedure. The fluoride-doped sensor showed a rapid response time of 6 s and 2 s, a maximum humidity hysteresis of 1.16% RH, high short-term repeatability, long-term stability, and excellent selectivity in the range of 11.3–97.3% RH [121]. In 2017, Eric S. Muckley and colleagues developed a material for humidity detection by intercalating K and Mg between the MXene layers. They were able to demonstrate that the d-spacing was initially larger for Ti3C2-K (11.5Å versus 10.1 Å in Ti3C2-Mg) using neutron scattering. After hydration, Ti3C2-Mg had a greater d-spacing (14.6 Å against 12.4 Å in Ti3C2-K) as the smaller size and higher charge of Mg2+ allow for a stronger interaction with H2O, resulting in a larger hydrated radius of Mg2+ (~4.28 Å versus ~3.31 Å). The greater radius causes a bigger hydrated d-spacing in Ti3C2-Mg after hydration. The emerged Mg(H2O)6 cluster was shown to be the best candidate for the pillaring effect as a result of creating a stiff structure. Relative humidity (RH) detection thresholds of ~0.8% RH were found in Ti3C2-K and Ti3C2-Mg films, and they also demonstrated a monotonic RH response in the 0–85% RH range with ~3% resolution [122].
One of the other methods of improving sensitivity is alkaline treatment. Organ-like alkalized Ti3C2Tx sensors were created by Zijie Yang et al. to monitor NH3 and humidity at room temperature. The device based on alkalized Ti3C2Tx had opposite response signals with improved NH3 and humidity detecting capabilities compared to not alkalized Ti3C2Tx. The Ti3C2Tx altered carrier type of conductivity following oxygen functionalization was the cause of the response signal’s change in direction [16].

6. MXene-Based Gas Sensors

The fast industrialization of society has created numerous issues related to volatile pollutants and greenhouse gases, posing threats to both the environment and human health. Detecting and quantifying the presence of harmful gases in a timely and precise manner are essential both for mitigating the negative effects of the pollutants and for providing accurate data for environmental research. MXenes are currently attracting significant attention in the field of gas sensing due to their exceptional qualities, such as their high metallic conductivity, abundant and tunable functional groups (-OH, -O, or -F), porous structure, and quick electron transfer ability [123,124].
The fundamental gas sensing process of pure MXenes’ is based on the adsorption and desorption of gas from the sensing layer and the ensuing change in the concentration of free charge carriers [125]. The alteration in the electrical conductivity of MXenes arises from the reaction of target gases (as demonstrated in Figure 8 for NH3) with surface defects and termination groups:
2NH3 + 3O → N2 + 3H2O + 3e
NH3 + OH → NH2 + H2O + e
Figure 8 show General gas sensing mechanism of pristine MXenes for NH3.
Despite their unique properties, pristine bare MXenes face certain limitations that hinder their effectiveness as sensing materials sensitive towards a broad range of gases. Challenges during the development of gas sensors include a lack of specificity to certain gas groups (excluding hydrogen-binding gases like acetone and NH3 as research shows MXenes have high selectivity towards it, compared to most other 2D materials [127,128]) and molecular structures, a relatively low bandgap, and a mechanical stability that is not always sufficient. These limitations can impact the sensitivity and selectivity of MXene-based gas sensors, making them less suitable for comprehensive gas sensing applications. Researchers and engineers are actively exploring various strategies to overcome these challenges. Surface functionalization, hybridization with other materials, and structural modifications are some of the approaches being investigated to enhance the performance and detection spectrum of MXene-based gas sensors [129,130]. Synergistic approaches with other materials, such as metal oxides, conducting or non-conducting polymers, 2D carbon nanomaterials, chalcogenides, and other surface functionalizing materials, might improve the properties of pure MXene sensors [16,122]. Depending on the characteristics of the components, the synergistic and reverse enhancing effect in MXene composites directly affects the sensing mechanism. For example, the basic mechanism for MXene composites containing metal oxides is quite similar to that of conventional metal oxide semiconductor-based sensors, in which the gas analyte’s effective absorption or desorption on the surface of the sensing material results in a change in the device’s resistance. The interfacial interactions between the two contributing materials and heterojunction production are related to the gas-sensing process of MXene-metal oxide composites [131].

6.1. NH3 Sensing by MXene-Based Sensors

Ammonia (NH3) is an important raw material in many domains. However, NH3 is a hazardous, corrosive, and colorless odor gas, and even at low concentrations can be dangerous to human health [132,133,134]. Therefore, quick and accurate monitoring of NH3 gas is crucial to ensure safety in industrial settings [135]. Xiao and co-workers revealed the advantages of using MXenes for NH3 sensing using first-principle simulation. The results showed that NH3 has a strong affinity to MXene substrates [136].
Seon Joon Kim and colleagues used interfacial self-assembly to create ultrathin Ti3C2 MXene films. The film was able to detect a concentration of 5 ppm NH3, and its sensitivity was 0.46% [137]. Eunji Lee et al. assembled Ti3C2Tx MXene sheets on flexible polyimide platforms using a straightforward solution-based technique. The constructed sensors responded strongly to ammonia (100 ppm with a sensitivity of 0.21%) [127]. Meng Wu et al. created Ti3C2 MXene-based sensors that had a varying-resistance-based response to numerous gases, namely CH4, H2S, methanol, ethanol, acetone, NH3, and NO. The sensors had good selectivity for NH3, with a 6.13% response, LOD of 10 ppm, and good linearity in the 10–700 ppm range [128].
Despite promising results, low sensitivity, long recovery time, and resistance drift after exposure to NH3 are still major issues for room temperature sensing of NH3 using pristine Ti3C2Tx MXene sensors [16,127,128]. Additional nanomaterials like graphene, carbon nanotubes (CNTs), reduced graphene oxide (rGO), and graphene oxide (GO) can be applied to promote the reaction with NH3 gas molecule [138]. By using a scalable wet-spinning technique, Sang Hoon Lee et al. created Ti3C2Tx MXene/graphene hybrid fibers. Excellent mechanical and electrical qualities made these fibers ideal for use in flexible wearable gas sensors. According to the authors, in comparison to pure MXene and graphene, the NH3 sensing response was considerably improved (∆R/R0 = 6.77%) in the created sensor [129].
Metal oxides are frequently chosen as sensing film materials for various sensing devices [139,140]. Metal oxides are also applied to further improve the NH3 sensing properties of resistive-based gas sensors because they can prepare a high amount of charge carriers [138]. TiO2- [141], nonstecheometric-TiO2- [142,143] and WO3-based [144,145,146] gas sensors are very promising for NH3 detection. In2O3/Ti3C2Tx MXene composites were created by Miao Liu et al. and displayed exceptional gas sensing qualities, including a high response (60.6%) to 5 ppm NH3 at RT, rapid response/recovery time (3/2 s), excellent selectivity, linearity, and good stability in the 5–100 ppm range. Compared to pristine MXene sensors and other metal oxides sensors, the designed sensor had better sensitivity and higher response to NH3 [147]. A hybrid heterojunction structure was created by Miao Liu et al. using the α-Fe2O3/Ti3C2Tx MXene composite material, which has a high specific surface area and a lot of functional groups. For 5 ppm NH3, the Fe2O3/Ti3C2Tx MXene sensor showed a response value of 18.3% and quick response/recovery times sub 2.5 s [135]. A ZnO/MXene hybrid surface acoustic wave (SAW) sensor was designed by Kedhareswara Sairam Pasupuleti et al. The sensor increased selectivity with an ultralow detection limit (89.41 ppb), displayed quick response/recovery time (92/104 s), long-term stability, and strong sensitivity under UV activation [148]. To improve the NH3 sensing capabilities of α-Fe2O, Miao Liu et al. created a hybrid structure of three components: Au, α-Fe2O3, and Ti3C2Tx MXene. The Au/α-Fe2O3/Ti3C2Tx MXene sensor displayed an excellent NH3 sensing performance, and its response value to NH3 (1 ppm) reached 16.9% at RT. Compared to other sensors (based on pure MXene or metal oxides), the obtained sensor had better sensitivity and response/recovery times [149].
The need for wearable NH3 sensors capable of real-time monitoring on diverse surfaces, including skin and clothing, is rather urgent. Polymer-based sensors have gained attention as one of the most promising options for addressing this requirement [150]. Polymer materials offer versatility and compatibility with a variety of substrates [151], making them well-suited for wearable sensor applications [152,153]. Polymer-based sensors can be designed to be flexible, lightweight, and adaptable, allowing for comfortable integration into wearable devices [140,154]. Additionally, polymers can exhibit sensitivity to specific gases, including NH3, and their properties can be tailored to enhance selectivity and performance [155,156]. The development of wearable NH3 sensors using polymer materials holds great potential for applications in various fields, including agriculture, environmental monitoring, and personal health. Ongoing research aims to optimize the design and performance of these sensors, paving the way for effective and versatile real-time monitoring solutions [157]. PEDOT:PSS/MXene composites were made by Ling Jin et al. and used in the fabrication of a dip-coated gas sensor. The composite sensor demonstrated a gas response of 36.6% against 100 ppm of NH3, with the response and recovery times being 116 s and 40 s, respectively. Additionally, it showed superior NH3 sensing performance at RT compared to pure PEDOT:PSS- and MXene-based sensors [158]. A portable self-powered NH3 sensor device was designed by Xingwei Wang et al. by implementing a bi-functional material made of 2D PANI/V2C MXene nanosheets. The authors discovered that the PANI/V2C nanosheets perform much better than pure PANI and MXene nanosheets in terms of response value (∆R/R0 = 14.9%), stability, and response/recovery time (9 s and 9 s) [159]. Zr3C2O2-based MXene is also a suitable material for room temperature NH3 sensors with excellent sensitivity and great selectivity, according to Xiumei Li et al. They proposed that a monolayer of Zr3C2O2 can be used as an NH3 detector since the NH3 molecule can be weakly chemically adsorbed with an adsorption energy of −0.597 eV, noticeable charge transfers (0.117 e), and a quick recovery time (1.05 ms) [160]. Such weak adsorption is a technological advantage because it enables easy and fast regeneration of the sensor. In conclusion, the discussed research results for MXene-based ammonia sensors showcase impressive applicability and potential for further development.

6.2. Determination of Volatile Organic Compounds (VOCs)

A wide range of sectors, including industrial safety control, environmental monitoring, and personal health management, have benefited greatly from the invention of gas sensors that are capable of accurately detecting volatile organic compounds (VOCs). These sensors play a crucial role in ensuring safety in industrial settings by monitoring the concentrations of VOCs and provide the necessary tools for environmental monitoring such as air quality evaluation. As VOC exposure can negatively impact personal health, sensors integrated into personal devices or wearable technology and able to provide real-time information about the surrounding air could ensure that the users are informed about the possible dangers of the environment they are in. Additionally, detection of VOCs emitted from the individual’s breath can also aid in early detection of certain illnesses and provide various health insights. Moreover, such sensors would be beneficial in residential and commercial spaces to monitor the indoor air quality and help in maintaining a comfortable and safe environment for the users of the space. Due to the abovementioned reasons, rapid advancements in VOC sensor technologies are important to our society as they contribute to improved safety, environmental sustainability, and overall human well-being in various areas [161,162,163]. Acetone sensors are gaining popularity due to their promising uses in self-diagnosis and health monitoring [164,165]. As an example, the acetone concentration in human breath is recognized as a crucial indicator for the identification of diabetes [166]. W18O49/Ti3C2Tx composites were made by Shibin Sun et al. by combining Ti3C2Tx MXene sheets with W18O49 nanorods. The obtained W18O49/Ti3C2Tx-based acetone sensor exhibited an extremely low detection limit of 170 ppb, a quick response/recovery times (5.6/6 s to 170 ppb of acetone), and a high response of 11.6 to 20 ppm of acetone. The sensor also demonstrated nearly ideal selectivity towards acetone. However, the main problem with this parameter was the high response to ammonia, though authors claim that due to slower response and recovery rates to NH3, this should not be an issue. According to the authors, the distribution of W18O49 on the Ti3C2Tx surface, the removal of the fluorine terminational groups from the MXene, and the synergistic interactions between the W18O49 and the Ti3C2Tx provided the combined sensors superior acetone-sensing properties in comparison to W18O49 and Ti3C2Tx sensors [167]. The 3D Ti3C2Tx MXene/rGO/CuO aerogel sensor developed by Miao Liu et al. responded to 100 ppm acetone with a 52.09% rate at room temperature and demonstrated a quick response/recovery time (6.5 s/7.5 s) as well as high repeatability and selectivity. According to the scientists, the results are due to the highly interconnected porous structure and the combined effects of MXene, rGO, and CuO. Compared to pristine MXene and rGO/In2O3 sensors, the obtained sensor had a superior response/recovery time and higher response [168]. An urchin-like V2CTx/V2O5 MXene hybrid sensor was developed by Sanjit Manohar Majhi et al., and it demonstrated an increased response (S% = 11.9) toward 15 ppm acetone, a low limit of detection (250 ppb), and a quick response/recovery time (115s/180s). The selectivity test showed higher response to acetone in comparison to CO, H2, CO2, H2S and C2H4 (47% versus <15% towards other gases). According to the authors, the potential production of H-bonds in multilayer V2C MXenes, the synergistic effect of the created V2C/V2O5 composite, and the strong charge carrier transport at the V2O5 and V2C MXene interface can be applied for improving the sensing properties [169].
Gaseous toluene (C6H5CH3) is a dangerous VOC that is created during the process of tanning leather and other materials. Even at rather low concentrations, it is deadly to humans and very harmful to the environment [170,171]. To decrease the risk to one’s health from exposure to toluene gas, it is crucial to detect it early and monitor its concentration both indoors and outdoors. CuO/Ti3C2Tx MXene hybrids were created by Angga Hermawan et al., who also attributed the increased toluene gas sensing response of 11.4, selectivity, response (270 s) and recovery times (10 s) to the strong metallic phase conductivity of Ti3C2Tx. However, the developed sensor had some significant issues with selectivity. The response signal to toluene was not much higher compared to ethanol and acetone signals (~11% versus 7%, respectively) [172]. A toluene sensor based on the MXene Mo2CTx was proposed by Wenzhe Guo et al. It had an LOD of 220 ppb, an adequate selectivity for toluene against other VOCs (2.75% response compared to 1% of benzene and even lower responses towards other analytes), and a good sensitivity of 0.0366 Ω/ppm (140 ppm for toluene) [173].
Formaldehyde (HCHO) has been regarded as the primary contributor to sick building syndrome [174,175]. HCHO can cause the irritation of mucous membranes, skin, eyes and throat, cause shortness of breath, headaches, nausea, and possibly lead to cancer. This substance is one of the most harmful and carcinogenic VOC gases in urban and indoor environments [176,177,178]. Therefore, it has become necessary to create high-performance portable HCHO sensors. Co3O4 and ZnO/Ti3C2Tx MXene nanowire arrays were used to create the composite-based formaldehyde sensor that was proposed by Dongzhi Zhang and colleagues. The MXene/Co3O4 composite sensor demonstrated selectivity (~9.5% compared to the highest of ~1.75% towards acetone) and outstanding response characteristics (9.2 to 10 ppm HCHO, with an LOD of 0.01 ppm). The selectivity to HCHO was also improved compared to pure MXene and Co3O4 sensors. The synergistic interfacial interactions between MXene and Co3O4 were thought to be the cause of the improved responsiveness to HCHO [179]. To design a formaldehyde sensor, Gaoqiang Niu et al. applied a pre-oxidation technique to change SnO2/Ti3C2Tx MXene into SnO2/TiO2/Ti3C2Tx MXene nanocomposites. Due to the synergistic interaction between SnO2 nanosheets and TiO2/Ti3C2Tx MXene, the sensor demonstrated a high response rate (38.6 at 160 °C to 20 ppm with LOD being 50 ppb), strong selectivity (38.4 compared to the highest 16.0 towards ethanol), and reproducibility. When the selectivity of the sensor was compared to a pure SnO2 sensor, the results showed clear improvement of the signal (38.4 compared to <20%) [180].
Another VOC, methanol, is used as an organic solvent in numerous sectors, including the pharmaceutical, chemical, biomedical, and automotive industries. However, it is extremely combustible and has a low explosive limit of 6% in air. When a person is exposed to more than 200 ppm of methanol for 8 h, it can lead to intoxication [181]. Additionally, unlike other alcohols, methanol oxidizes slowly and builds up in the human body, which can result in poisoning. Therefore, it is crucial to design a methanol sensor that has good sensitivity and selectivity and can operate at ambient temperatures [182]. In2O3 nanocubes/Ti3C2Tx MXene nanocomposites showed strong responsiveness (29.6% to 5 ppm) and noticeable selectivity to methanol against other gases (30% towards methanol versus <18% towards trimethylamine, acetone, xylene and other gases) at room temperature, as reported by Miao Liu et al. According to the authors, the sensor was characterized by quick reaction and recovery times (6.5 and 3.5 s, respectively) to 5 ppm methanol at ambient temperature [183].
Ethanol is employed extensively in the pharmaceutical, food manufacturing, and biomedical industries as a key chemical ingredient [184,185]. Ethanol also poses a serious risk to the safety of production and transportation due to its combustible and explosive nature (explosion range of 3.3% to 19%) [186]. Additionally, long-term exposure to an atmosphere even with low levels of ethanol (25 ppm) can lead to serious negative health effects, including headaches, liver damage, and neurasthenia [187,188]. Therefore, designing and creating sensing materials for monitoring low-concentration ethanol gas at low temperatures is greatly import. Shuai Zhang et al. developed MoO3/Ti3C2Tx MXene nanocomposites employing a straightforward hydrothermal synthesis technique for efficient monitoring of low ethanol concentrations. The sensor responded well to low-concentration ethanol gas at 100 °C (Ra/Rg = 1.61/1 ppm) and recovered quickly (10 s/49 s). Outstanding selectivity for ethanol was obtained, with the response being five times higher compared to other gases, like methanol and acetone [189]. In order to create resistance-type gas sensors, Chen Wang et al. produced SnO2/MXene composites that were brushed onto a MEMS platform. The composition significantly increased the sensitivity of SnO2-based gas sensor, with ethanol gas having the maximum sensitivity. The composite additionally accelerated the sensor’s response and recovery times (14s and 26s, respectively) and reached a 5.0 response value to 10 ppm ethanol at its ideal temperature of 230 °C, which was twice as high as the response value of a pure SnO2 sensor. Though the selectivity of the sensor was improved after adding MXenes, the response towards acetone is still comparable to ethanol (~4 towards acetone compared to ~5 towards ethanol). The improvement in the composite’s gas-sensing characteristics is attributed, according to the authors, to its Schottky barrier and the metal-like conductivity of Ti3C2Tx [190].

6.3. Other Gas Sensors Based on MXenes

There are other gases that can be monitored by MXene-based gas sensors. One of them is H2, a gas with a significant energy capacity that is renewable and sustainable. It is commonly used in the metallurgy, fuel cell, and automotive industries [191,192]. However, H2 gas is highly explosive in air at concentrations of more than 4% and has a low ignition energy. Additionally, due to its small molecule size and high diffusion coefficient, it leaks during usage, storage, and transit [193,194]. Because H2 gas is odorless, our sense of smell system is unable to detect it; hence, the creation of high-response H2 sensors is essential to guaranty safety. A unique Pt = Pd/Ti3C2Tx hydrogen sensor was developed by Lei Wang et al. using a simple hydrothermal chemical reduction process. The resulting sensor had a response of 24.6% (200 ppm H2 at RT), a response/recovery time of 6/8 s, an LOD of 200 ppm, and strong selectivity (24.6% for hydrogen compared to the highest towards ethanol 5.2%) in addition to good linearity, long-term stability, and repeatability [195].
The main sources of nitrogen dioxide (NO2) are industrial combustion, automotive exhaust, and coal mining. According to research, even very small concentrations of NO2 (~1 ppm) can harm the respiratory system and eyes to varying degrees. Furthermore, persistent exposure to higher NO2 concentrations may possibly be fatal to humans [196]. Fuqiang Guo et al. effectively created Ti3C2Tx/CuO nanocomposites using an eco-friendly and straightforward hydrothermal process for NO2 sensing. The response of the author’s nanocomposite sensor to NO2 (56.99%) was five times greater than that of the pure Ti3C2Tx sensor. The improved gas sensing capability was attributed to higher oxygen vacancies and adsorbed oxygen in comparison to Ti3C2Tx as well as hybrid heterojunctions between CuO and Ti3C2Tx. The Ti3C2Tx/CuO sensor demonstrated extremely quick response and recovery times (16.6/31.3 s to 20 ppm NO2), exceptional selectivity to NO2 (with the response being ~57% towards NO2 versus 31.09% towards NO and 12.63% towards ammonia), and long-term stability [197].
Other researchers have demonstrated that other gases, such as NOx [198,199], hexanal [200] and others [201,202], can also be determined by gas sensors based on MXenes. It can be predicted that due to the unique characteristics of MXenes, such as tunable terminal groups, and metallic conductivity, they have been successfully applied in gas sensor design. Compared to other materials, pristine MXenes can offer high sensitivity with mediocre selectivity (except for hydrogen bonding gases) and a lower linear range. Combining MXenes with other more selective or more stable materials (especially with metal oxides or metal nanoparticles) usually results in much better results.

7. Chemical Sensors

The socio-economic changes induced by industrialization in the middle of the 19th and 20th centuries have considerably improved the quality of life of many people. However, such advancements have also resulted in pollution and contamination of water, soil, air, etc. Naturally, MXene nanomaterials have drawn a lot of interest in the field of pesticide sensing, heavy metal sensing, etc., because of their exceptional electrical and optical properties. Efficient functionalization of MXenes is made possible by the surface termination groups, which is essential for electrochemical sensing [203] as they give MXenes a superb reducing ability. Qualities of MXenes make them suitable for improvement of the functionality of sensors and biosensors, providing higher sensitivity, selectivity to some analytes, and strong reproducibility but shorter detection ranges compared to other materials. Though MXenes have remarkable electrocatalytic activity, one of their main flaws is that they are unstable in the anodic potential window and exhibit an irreversible oxidation peak at roughly 0.43 V. Thus, combining MXenes with other components is essential for further enhancement of sensory abilities [204].

7.1. Nitrite Sensors

Nitrite (NO2) is one of the most prevalent nitrogen-containing ions, which is widely used in the chemical and food industries [205,206]. Detection of nitrite is important as it poses serious dangers for human health. The nitrite oxidizes hemoglobin into methemoglobin, reducing the blood’s ability to carry oxygen [207], and it can also react to form N-nitrosamine, which can cause stomach cancer [208]. Tan Wang et al. created a novel nanocomposite of AuNPs/Ti3C2Tx/ERGO for the sensitive electrochemical detection of nitrite. AuNPs/MXene/ERGO exhibited enhanced catalysis for nitrite oxidation compared to ERGO, MXENE/ERGO, and AuNPs/ERGO due to the synergy of these components. The developed sensor’s overall electrocatalytic oxidation of nitrite followed the mechanism:
NO2 + H2O → NO3 + 2H+ + 2e
Figure 9 shows the detection scheme using the obtained sensor. Linear ranges of obtained sensors were 0.5 to 80 µM and 80 to 780 µM, respectively, with an LOD of 0.15 µM and 0.015 µM and a sensitivity of 340.14 and 977.89 µA·mM−1cm−2 with great selectivity to nitrite (the sensor did not react to different interference salts, like KCl and NaNO3, even when their concentrations were much higher) [209].
The synergetic catalytic effect of the AuNPs/Ti3C2Tx-PDDA nitrite sensor was designed by Yuhuan Wang et al. The Ti3C2Tx improved electron transport, and Ti3C2Tx-PDDA composites gave AuNPs more places to attach during the electrodeposition process. The AuNPs/Ti3C2Tx-PDDA/GCE sensor demonstrated exceptional linear ranges in the ranges of 0.1–2490 µM and 2490–13490 µM, high sensitivity of 250 µA·mM−1·cm−2, and great selectivity with an LOD of 0.059 µM. The sensor had a superior working range length and detection limit compared to a combination of AuNPs with graphene or CNT [210]. In order to combine the great electrocatalytic activity of AuNPs with carbon quantum dots (CQDs), MXene’s substantial specific surface area, and the good electrical conductivity of AuNPs with MXene, Xiwen Feng et al. created a AuNPs@CQDs-Ti3C2Tx MXene nitrite sensor, where MXene was used as the immobilization matrix. The sensor had a detection limit of 0.078 µM and a linear detection range of 1 µM to 3200 µM with great sensitivity and selectivity, as the sensor did not respond towards nitrates, chlorides and sulphates. However, it is worth noting that working range and detection limit were inferior to the AuNPs/Ti3C2Tx-PDDA sensor [211]. A novel electrochemical sensing platform for nitrite detection was developed by Jinghao Zhuang et al. It was based on a Ti3C2Tx MXene sheet and a three-dimensional urchin-like Co(VO3)2(H2O)4 (CoVO). The following was proposed as the mechanism for the electrochemical reduction of nitrite on CoVO/Ti3C2Tx electrodes:
Co2+ ↔ Co3+ + e
2Co3+ + NO2 + H2O → 2Co2+ + NO3 + 2H+
The CoVO/Ti3C2Tx-based composite displayed a significantly improved electrochemical signal with good linearity in the range of 0.5–50 µM and 50–2000 µM with an LOD of 0.1 µM [212].

7.2. Hydrazine Sensors

Hydrazine and its derivatives are widely used in fuel cells, photography, insecticides, dyes, and pharmaceuticals [213]. Hydrazine is a known carcinogen and exposure to it can cause temporary blindness, skin allergies, carcinogenesis, neurotoxicity, and organ damage, especially to the liver and kidneys [214]. Additionally, industrial waste and waterways contain it, making it a potentially widespread environmental pollutant [213]. Therefore, creating efficient methods to detect hydrazine at low concentrations is essential from an industrial, medical, and environmental perspective. The idea of using few-layered (FL) Cr2CTx MXene nanosheets as a receptor integrated into a microresonator to detect hydrazine at the ppm level was proposed by Bhuvaneswari Soundiraraju et al. [215]. The same author designed a FL-Cr2CTx modified electrode for hydrazine detection. With an LOD of 6.60 × 10−7 M (at S/N = 3), the electrode demonstrated a high sensitivity of 1.810 µA·µM−1 towards hydrazine. The LOD for hydrazine was also more than 10 times lower than that achieved by electrodes modified by Ti3C2Tx and Nb2CTx-based pristine MXenes. The authors also presented a possible mechanism for detecting hydrazine by the sensor:
2 Cr2COx (surface) + N2H4 → 2Cr2COx−1 + N2 + 2 H2O
The developed sensor was also employed to quantitatively analyze hydrazine liquid propellant spiked in water with satisfying results [216]. Yanqing Yao et al. intercalated a Ti3C2Tx-based MXene as a conductive platform, which dramatically increased the conductivity of ZIF-8. The obtained sensor demonstrated excellent sensing performance for hydrazine, with amazing selectivity (with no reaction to nitrates, chlorides, ethanol, DMF and other possible interferences), a 10–7700 µM linear range, and an LOD of 5.1 µM, thanks to the synergistic effect of MXene/ZIF-8 nanocomposite. Compared to other sensors, the MXene/ZIF-8 sensor had a superior linear range, but it lacked a sufficient LOD [217]. Sundus Gul et al. used Nb2C MXene as a direct electrode material for electrochemical hydrazine sensing. They also added erbium (Er) to several Nb2C MXene samples and compared the results with non-doped ones. The pure MXene and Er-doped samples were found to have sensitivity values of 169 µAm−1·M−1·cm−2 and 276 µAm−1·M−1·cm−2 and LOD values of 10.08 mM and 67 µM, respectively, demonstrating that the Er-doped MXene is significantly more sensitive than the unmodified one. According to the authors both mentioned MXenes are potential candidates for hydrazine sensing [218].

7.3. Sensing of Phenol and Its Derivatives

The US Environmental Protection Agency and the European Union have identified phenolic compounds as environmental contaminants [219]. Phenolic compounds are common waste products of many chemical facilities. Bisphenols are frequently employed in a range of industries, along with benzenediols and aminophenols. Many of these compounds, even in very low quantities, are harmful to people, plants, animals, and biological and ecological ecosystems [220]. Thus, developing sensors to monitor the aforementioned substances is important to ensure the health of society and the environment. Lingxia Wu et al. described the application of Ti3C2 in phenol sensing [221]. According to Ling Lei et al., Ti3C2Tx MXene nanosheets are significantly more active for the oxidation of 4-chlorophenol (4-CP) and 4-nitrophenol (4-NP) than electrodes made from graphite. The linear ranges for 4-CP and 4-NP detection were 0.1 to 20.0 µM and 0.5 to 25.0 µM with detection limits of 0.062 µM and 0.11 µM, respectively. Though the sensitivity and LOD were significantly better than that of other electrodes, the linear range and stability were still lower in comparison to them. The sensor created by the authors was successfully applied to detect 4-CP and 4-NP in wastewater samples with promising results [222]. The exfoliated Ti3AlC2 was applied in sensor design, without oxidation, and the results demonstrated improved sensitivity of 16.35 µA·µM−1·cm−2 and a lower detection limit of 42 nM/L for 4-NP. With an excellent linear sensing range, MXene sensor electrodes were capable of detecting 4-NP in a broad concentration range from 500 nM to 100 M. The selectivity was also high, as the electrode did not react to phenol and its derivatives, like chlorophenol. The characteristics of the sensor were far superior in comparison to graphene-based sensors. The possible mechanism behind the 4-NP oxidation by Ti3C2Tx was also described as: Nanomaterials 14 00447 i001
The sensors have also successfully shown reliability and repeatability in continuous monitoring of 4-NP in tap water samples with low interferences [223]. A new heterostructure made of Nb2CTx and a Zn-Co-NC-based nanocage was proposed by Runmin Huang et al. The performance of the Nb2CTx/Zn-Co-NC-based sensor in the detection 4-NP was advanced by the large specific surface area and conductivity of MXene and the enhanced catalytic activity of the Nb2CTx/Zn-Co-NC-based nanocages. The linear range of the sensor was 1 µM to 500 µM, the LOD was 0.070 µM, and the sensitivity was 4.65 µA·µM−1·cm−2. The repeatability of the sensor was also studied, and it was shown that the relative standard deviation (RSD) was less than 3.48% [224].
The P. Abdul Rasheed group used a platinum nanoparticle/Ti3C2Tx nanocomposite to measure another phenol derivative—Bisphenol A—and to improve the electrochemical performance and stability of MXene used for the modification of electrodes. The sensor’s mechanism involves the oxidation of Bisphenol A [225]: Nanomaterials 14 00447 i002
The proposed sensor had an extremely low LOD (32 nM) in a linear range of 0.05–5 µM, and it was further demonstrated to be successful in detecting Bisphenol A in samples of fresh milk and drinking water [225]. For efficient catechol and hydroquinone detection, Runmin Huang et al. decorated alkalization-intercalated Ti3C2 with N-doped porous carbon produced from MOFs. This strategy successfully stopped the Ti3C2 sheets from stacking and allowed benzenediol detection, which was based on hydrogen-bond interaction, due to the abundance of C-N and -OH functional groups. The sensor was characterized by a broad linear range of 0.5–150 µM with low detection limits of 4.8 nM (S/N = 3) for hydroquinone and 3.1 nM (S/N = 3) for catechol. The sensing mechanism revolved around oxidation of hydroquinone and catechol into quinone and 1,2-benzoquinone, respectively. The selectivity of the obtained sensor was high, with an RSD less than 6.6 from a large list of possible interferences (from ions to other phenols like bisphenol A and resorcinol). Compared to sensors from other materials, the obtained one had an adequate linear range and one of the lowest detection limits. Catechol and hydroquinone were also detected by the authors using the acquired sensor in industrial effluent with good recoveries [226].

7.4. Pesticide Sensors

Although pesticide residues pose a serious risk to human health, pesticides play a significant role in agricultural fields; hence, it is necessary to develop sensitive and quick methods for pesticide monitoring. For the purpose of determining the pesticide carbendazim (CBZ), Xiaolong Tu et al. developed a nanoarchitecture of Ti3C2Tx/CNHs (carbon nanohorns)/-CD-MOFs (cyclodextrin–metal–organic frameworks). Due to the synergistic combination of the high electronic conductivity and better catalytic activity from MXene/CNHs and the plentiful electrocatalytic active sites from -CD-MOFs, the MXene/CNHs/-CD-MOFs demonstrated boosted catalytic activity toward CBZ oxidation. The mechanism behind CBZ sensing involved oxidation of the compound on the sensor: Nanomaterials 14 00447 i003
A large linear range from 3.0 nM to 10.0 µM and a low LOD of 1.0 nM (S/N = 3) were both determined for the MXene/CNHs/-CD-MOF electrode. The produced sensor was then used in tomato samples, exhibiting high selectivity, repeatability, and long-term stability [227]. Yu Xie et al. used MXene/ERGO composite for CBZ detection to enhance the Ti3C2Tx sensing capability. The separated Ti3C2Tx layers and particles were tightly connected by the ERGO conductive networks, thus improving the electrochemical reactivity of the electrode. The electrochemical sensor was characterized by a low LOD of 0.67 nM and a large linear range of 2.0 nM to 10.0 µM and was used to measure CBZ in food products [228]. Prussian blue (PB) and Ti3C2Tx hybrid composites for H2O2 and malathion sensing were created by Ying He et al. In the sensor, Ti3C2Tx served as a reducing agent, and the multilayer 2D nanostructure efficiently prevented PB nanoparticles from aggregating. With a detection limit of 1.3 × 10−16 M, good linearity of malathion was obtained in the range of 1.0 × 10−15 to 1.0 × 10−9 M [229]. Fengnian Zhao et al. created Au-Pd bimetallic nanoparticles through a self-reduction process using ultrathin MXene nanosheets as the reducing agent and support. These nanoparticles advanced the immobilization of acetylcholinesterase and the determination of paraoxon. The MXene/Au-Pd biosensor with an LOD of 1.75 ng·L−1 showed a good linear relationship with paraoxon concentration from 0.1 to 1000 µg·L−1, and promising results with cucumber samples were obtained [230]. A new acetylcholinesterase biosensor based on Ti3C2Tx nanosheets and chitosan was developed by Liya Zhou et al., and it was found to be suitable for the detection of organophosphate pesticides. The biosensor performed well in terms of malathion detection, with a linearity in the range of 1 × 10−14–1 × 10−8 M and an LOD of 0.3 × 10−14 M [231].

8. Optical Sensing

Optical sensing is a promising approach for the detection of chemical compounds because of its high accuracy, strong anti-interference ability, and the non-invasive nature of the measurement [105]. Ti3C2Tx-based MXene has a variety of intriguing characteristics that can benefit from optical sensing [232,233,234]. Ti3C2Tx has strong wavelength dependence in its optical absorption zone (from the visible to the short-wave infrared) [235]. In addition, Ti3C2Tx has a high intrinsic photothermal conversion efficiency in the infrared irradiation range, allowing the absorbed irradiation power to produce heat effectively [236]. The photoelectric properties of MXenes can also be easily controlled by surface modification [237]. Wide-band optical absorption and good optical modulation are the result of the direct band gap structure of MXene, which may be altered by changing surface functional groups or the number of layers [238,239].
Electrochemiluminescence (ECL) is a sensitive analytical technique that combines the benefits of electrochemistry and luminescence [240,241,242]. Electrochemically produced intermediates in ECL go through a highly exergonic process to create an excited state, which eventually relaxes to a lower-level state with the emission of light. ECL offers great sensitivity, low cost, and a wide dynamic range; therefore, it can be integrated within simple, easily controllable, and portable analytical instrumentation [243,244]. A prominent biomarker for the diagnosis of chronic myelogenous leukemia is the oncogenic BCR-ABL fusion gene, a distinctive gene of the Ph chromosome [245]. Heterojunction based on CeO2/Ti3C2Tx was applied to an ECL biosensor for the detection of BCR-ABL. The synthesized heterojunction with strong dispersibility and a large specific surface area synergistically increased the ECL signal of the S2O82−/O2 system and improved the immobilization of biomolecules. The sensor demonstrated a large working concentration range from 1 fM to 100 pM [246].
The process of photoelectrochemistry, or PEC, is defined as the interaction of optical and electrochemical processes wherein an applied light causes electron excitation after charge transfer via photoexcited material [247,248]. The PEC technique uses light as the detecting signal and photocurrent excitation, which is the opposite to ECL [249]. The advantages of sensing based on PEC are the low background signal, easy setup, and excellent sensitivity [250,251]. The PEC-SERS dual-mode biosensor was developed by Fanglei Liu et al. to detect Staphylococcus aureus with high accuracy and sensitivity. The method made use of the superior PEC and SERS modes of carbon nitride nanosheet (C3N4)/Ti3C2Tx-AuNPs. According to experimental findings, PEC and SERS work well together to amplify Staphylococcus aureus detection, with detection ranges of 5–108 CFU/mL and an LOD of 0.70 CFU/mL for PEC, and 10–108 CFU/mL and 1.35 CFU/mL for SERS, respectively [252].
The measurement of fluorescence intensity is mostly applied during the action of fluorescence sensors. Typically, these sensors yield low LOD and high sensitivity. Fungal toxins are very dangerous substances due to their mutagenicity, teratogenicity, carcinogenicity, and other characteristics [253,254]. Therefore, the development of efficient methods for toxin detection is crucial, as the potential contamination of cereals, peanuts, and maize puts human and animal health at risk [255]. A unique fluorescent biosensor for Aflatoxin B1 (AFB1) detection based on the CRISPR/Cas12a system deposited on Ti3C2Tx was created by Zhihui Wu et al. When AFB1 was present, the aptamer’s preferential binding to it released the activator, which in turn triggered Cas12a’s trans-cleavage activity, which indiscriminately cleaved ssDNA on MXenes and restored the fluorescence signal. With a detection range of 0.001 to 80 ng·mL−1 and a detection limit of 0.92 pg·mL−1, the fluorescent biosensor was highly versatile [256].
Optical-calorimetry-based techniques are mostly either quantitative or semiquantitative due to their moderate LODs [257], with principal advantages being their inexpensive construction costs and ease of use. Yapeng Li et al. manufactured Ti3C2/CuS nanocomposites and used them in a sensor for the detection of cholesterol. When utilized as a peroxidase to catalyze the reaction of 3,3,5,5-tetramethylbenzidine (TMB) in the presence of H2O2, MXene-based nanocomposites demonstrated peroxidase-like activity and caused a change in color to blue. The cholesterol determination using nanocomposites exhibited a linear range between 10 and 100 μM and LOD of 1.9 μM. Results of the same research illustrated that nanocomposites based on MXene-Ti3C2/CuS can be applied in clinical medicine for the detection of H2O2 and cholesterol [258].
Surface plasmon resonance (SPR) is a process in which photons of incident light at a specific angle excite the electrons in the metallic surface layer, which then travel parallel to the metal surface [259,260,261]. The refractive index, or RI, of the substance close to the metal surface determines the specific angle that initiates SPR when there is a constant light source frequency and a thin metal surface. Various analytes can be detected by this technique since a slight alteration in the sensing medium’s reflective index will prevent SPR from occurring from a certain angle [262]. Thiol-modified Nb2C quantum dots (Nb2C-SH QDs) have been developed by Rongyuan Chen et al. for use in a novel label-free SPR aptasensor for the detection of SARS-CoV2 N-gene. Nb2C-SH QDs improved the SPR response in addition to having a high bioaffinity for the aptamer. The SPR aptasensor demonstrated a low LOD of 4.9 pg·mL−1 sensitivity in the concentration range of 0.05 to 100 ng·mL−1. Additionally, a qualitative examination of N-gene from various sources, such as human serum, seafood, and seawater, was demonstrated using the aptasensor [263].
Raman scattering spectroscopy has distinct advantages over other spectroscopy techniques, including narrow peaks in multicomponent analysis and frequency shifts that offer “fingerprint” information about the chemical structure of an analyte [264]. Nevertheless, the limited sensitivity of Raman spectroscopy renders it useless for the identification of clinical indicators. Surface-enhanced Raman scattering (SERS) occurs when molecules are adsorbed on or near rough metal surfaces or metal-based nanostructures and exhibit an amplification in their Raman signal [262]. To combine the benefits of MXene and metal nanoparticles for SERS, Rongyang Liu et al. created a hybrid biosensor, Ti3C2Tx-AgNPs. Dopamine and adenine monitoring demonstrated the SERS sensor’s capacity to identify biological molecules; the LOD for adenine was 10−8 M, while the LOD for dopamine was 5 × 10−8 M. The biosensor’s capacity for detection in serum samples was further demonstrated [265].

9. Conclusions and Outlook

MXenes offer a myriad of advantages that have led to the development of various sensors, including strain, pressure, gas, electrochemical, optical, humidity, and multifunctional sensors. However, there are challenges and issues that need to be addressed for the optimal utilization of MXenes in sensor design. Some of the notable issues include specificity and selectivity, integration within other materials, stability and durability, cost-effectiveness, advanced manufacturing techniques, and difficulties in standardization. One of the challenging issues that needs to be resolved before effectively utilizing MXenes in sensor design is enhancing the specificity and selectivity of MXene-based sensors for certain target analytes. Additionally, achieving high sensitivity without compromising on the ability to discriminate between different substances is crucial for the reliability of the sensors. While MXenes exhibit excellent properties, combining them with other materials enables the achievement of the desired sensor characteristics. However, ensuring compatibility and efficient integration with other components is still a key consideration for practical sensor designs. The long-term stability and durability of MXene-based sensors, especially under harsh environmental conditions or repeated usage, need further improvement. Ensuring that the sensors maintain their performance over extended periods is essential for practical applications. The production cost of MXenes can be a limiting factor in large-scale applications. Research efforts are ongoing to develop cost-effective methods for MXene synthesis and processing to make them more accessible for widespread sensor deployment; however, at the current time, MXenes remain very expensive materials. Current manufacturing methods for MXene-based sensors involve complex processes. Therefore, streamlining and optimizing these techniques are essential for mass production and commercialization of MXene-based sensors. Further establishing standardized protocols for MXene-based sensor fabrication and testing is crucial for ensuring consistency and comparability across different research studies and applications. The environmental impact of MXene production and disposal is still under consideration. Therefore, a high number of sustainable and eco-friendly approaches for MXene synthesis and integration into sensors are being explored. It is still currently difficult to synthesize MXene-based sensing materials in a controlled manner. For instance, producing MXene nanosheets with specific parameters or surface terminations (currently, a lot of methods result in an abundance of -F groups, which can be damaging to human health and the environment) is still challenging. To this day, the most popular pathways to synthesize MXenes require using or producing hazardous chemicals. Secondly, the chemical stability of MXenes (especially in aqueous solutions) also needs further research and improvements, as the oxidation of MXenes (especially, Ti3C2Tx) hinders the progress in many applications (for example, biosensors). Finally, more different materials that belong to the MXene family could be applied in sensor design as V-, Mo-, Zr-, Nb-based MXenes and many others can provide a lot of benefits for certain types of sensors, compared to the most commonly used Ti3C2Tx-based MXene. Addressing these challenges will contribute to the continued advancement and widespread adoption of MXene-based sensors in various industries. Ongoing research and collaborative efforts are key to overcoming these issues and unlocking the full potential of MXenes in sensor design. Therefore, it is expected that a great number of MXene-based sensors, such as strain, pressure, gas, electrochemical, optical, humidity, and even multifunctional sensors, will be developed by taking advantage of their features, which includes their great electrical conductivity, tunable surface functional groups, a great surface area, superior electrochemical and electrocatalytic capabilities, and remarkable optical properties.

Author Contributions

Conceptualization, I.N., A.R. (Arunas Ramanavicius) and M.P.; methodology, I.N. and S.R.; software, I.N.; validation, I.N., A.R. (Agne Ramanaviciute) and A.R. (Arunas Ramanavicius); formal analysis, I.N. and A.R. (Agne Ramanaviciute); investigation, I.N.; resources, A.R. (Arunas Ramanavicius) and M.P.; data curation, A.R. (Arunas Ramanavicius) and S.R.; writing—original draft preparation, I.N.; writing—review and editing, I.N., A.R. (Agne Ramanaviciute) and A.R. (Arunas Ramanavicius); visualization, I.N.; supervision, A.R. (Agne Ramanaviciute) and S.R.; project administration, A.R. (Arunas Ramanavicius) and S.R.; funding acquisition, A.R. (Agne Ramanaviciute) and S.R. All authors have read and agreed to the published version of the manuscript.

Funding

This project received funding from the Research Council of Lithuania (LMTLT), agreement No S-PD-22-155. A.R. (Arunas Ramanavicius), M.P. and (Agne Ramanaviciute) are thankful to the Horizon Europe MSCA-2021-SE-01 projects (ARGO #101086441 and ESCULAPE #101131147).

Data Availability Statement

No new data were created or analyzed in this study. Data sharing is not applicable to this article.

Conflicts of Interest

The authors declare no conflicts of interest.

Abbreviations

VOC volatile organic compound
LOD limit of detection
CNT carbon nanotube
NW nanowire
PPy polypyrrole
HEC hydroxyethyl cellulose
MWCNTs multi-walled carbon nanotubes
TPU thermoplastic polyurethane
PDMS polydimethylsiloxane
PANI polyaniline
PANIF polyaniline fiber
P(VDF-TrFE) poly(vinylidenefluoride-co-trifluoroethylene)
PVA polyvinyl alcohol
CMC sodium carboxymethylcellulose
TA tannic acid
PVA-CA catechol-functionalized poly(vinyl alcohol)
PAA polyacrylic acid
PEDOT poly(3,4-ethylenedioxythiophene)
PSS poly(styrene-sulfonate)
ANFs aramid nanofibers
ZIF-67 zeolitic imidazolate framework-67
PAN polyacrylonitrile
PE polyethylene
CTAB hexadecyl trimethyl ammonium bromide
CMF carbon sponge
PPNs polypyrrole nanospheres
TPUEM thermoplastic polyurethane electrospinning membrane
GO graphene oxide
rGO reduced graphene oxide
PS polystyrene
BC bacterial cellulose
CA κ-carrageenan
PAM polyacrylamide
PHOS photothermal optical sensor
MZI Mach–Zehnder interferometer
MKR micro-fiber knot resonator
RH relative humidity
TOCNFs (TEMPO)-oxidized cellulose nanofibers
SAW surface acoustic wave
MEMS micro-electromechanical systems
NPs nanoparticles
ERGO electrochemically reduced graphene oxide
PDDA poly(dimethyl diallyl ammonium chloride)
GCE glassy carbon electrode
CQDs carbon quantum dots
FL-Cr2CTx few-layered two-dimensional chromium carbide
ZIF-8 zeolitic imidazolate framework-8
CC carbon cloth
4-CP 4-chlorophenol
4-NP 4-nitrophenol
NC nanocage
RSD relative standard deviation
MOF metal–organic framework
CBZ carbendazim
CNHs carbon nanohorns
CD-MOFs cyclodextrin–metal–organic frameworks
PB Prussian blue
ECL electrochemiluminescence
PEC photoelectrochemistry
CFU Colony-forming unit
CRISPR clustered regularly interspaced short palindromic repeats
TMB 3,3,5,5-tetramethylbenzidine
SPR surface plasmon resonance
SERS surface-enhanced Raman scattering

References

  1. Naguib, M.; Kurtoglu, M.; Presser, V.; Lu, J.; Niu, J.; Heon, M.; Hultman, L.; Gogotsi, Y.; Barsoum, M.W. Two-Dimensional Nanocrystals Produced by Exfoliation of Ti3AlC2. Adv. Mater. 2011, 23, 4248–4253. [Google Scholar] [CrossRef] [PubMed]
  2. Sun, N.; Zhu, Q.; Anasori, B.; Zhang, P.; Liu, H.; Gogotsi, Y.; Xu, B. MXene-Bonded Flexible Hard Carbon Film as Anode for Stable Na/K-Ion Storage. Adv. Funct. Mater. 2019, 29, 1906282. [Google Scholar] [CrossRef]
  3. Du, Y.-T.; Kan, X.; Yang, F.; Gan, L.-Y.; Schwingenschlögl, U. MXene/Graphene Heterostructures as High-Performance Electrodes for Li-Ion Batteries. ACS Appl. Mater. Interfaces 2018, 10, 32867–32873. [Google Scholar] [CrossRef]
  4. Zhou, Q.; Qian, K.; Fang, J.; Miao, M.; Cao, S.; Feng, X. UV-Light Modulated Ti3C2Tx MXene/g-C3N4 Heterojunction Film for Electromagnetic Interference Shielding. Compos. Part. A Appl. Sci. Manuf. 2020, 134, 105899. [Google Scholar] [CrossRef]
  5. Hantanasirisakul, K.; Gogotsi, Y. Electronic and Optical Properties of 2D Transition Metal Carbides and Nitrides (MXenes). Adv. Mater. 2018, 30, 1804779. [Google Scholar] [CrossRef]
  6. Wen, Y.; Yang, J.; Zou, H.; Fan, Y.; Li, J.; Kuang, Y.; Liu, W.; Zhang, K.; Xiong, L. 2D TiVCTx Layered Nanosheets Grown on Nickel Foam as Highly Efficient Electrocatalysts for the Hydrogen Evolution Reaction. RSC Adv. 2022, 12, 23584–23594. [Google Scholar] [CrossRef]
  7. Peng, J.; Chen, X.; Ong, W.-J.; Zhao, X.; Li, N. Surface and Heterointerface Engineering of 2D MXenes and Their Nanocomposites: Insights into Electro- and Photocatalysis. Chem 2019, 5, 18–50. [Google Scholar] [CrossRef]
  8. Lipatov, A.; Alhabeb, M.; Lu, H.; Zhao, S.; Loes, M.J.; Vorobeva, N.S.; Dall’Agnese, Y.; Gao, Y.; Gruverman, A.; Gogotsi, Y.; et al. Electrical and Elastic Properties of Individual Single-Layer Nb4C3Tx MXene Flakes. Adv. Electron. Mater. 2020, 6, 1901382. [Google Scholar] [CrossRef]
  9. Abdolhosseinzadeh, S.; Jiang, X.; Zhang, H.; Qiu, J.; Zhang, C. Perspectives on Solution Processing of Two-Dimensional MXenes. Mater. Today 2021, 48, 214–240. [Google Scholar] [CrossRef]
  10. Kamysbayev, V.; Filatov, A.S.; Hu, H.; Rui, X.; Lagunas, F.; Wang, D.; Klie, R.F.; Talapin, D.V. Covalent Surface Modifications and Superconductivity of Two-Dimensional Metal Carbide MXenes. Science 2020, 369, 979–983. [Google Scholar] [CrossRef]
  11. Shayesteh Zeraati, A.; Mirkhani, S.A.; Sun, P.; Naguib, M.; Braun, P.V.; Sundararaj, U. Improved Synthesis of Ti3C2Tx MXenes Resulting in Exceptional Electrical Conductivity, High Synthesis Yield, and Enhanced Capacitance. Nanoscale 2021, 13, 3572–3580. [Google Scholar] [CrossRef]
  12. Saravanan, P.; Rajeswari, S.; Kumar, J.A.; Rajasimman, M.; Rajamohan, N. Bibliometric Analysis and Recent Trends on MXene Research—A Comprehensive Review. Chemosphere 2022, 286, 131873. [Google Scholar] [CrossRef]
  13. Naguib, M.; Come, J.; Dyatkin, B.; Presser, V.; Taberna, P.-L.; Simon, P.; Barsoum, M.W.; Gogotsi, Y. MXene: A Promising Transition Metal Carbide Anode for Lithium-Ion Batteries. Electrochem. Commun. 2012, 16, 61–64. [Google Scholar] [CrossRef]
  14. Lim, K.R.G.; Shekhirev, M.; Wyatt, B.C.; Anasori, B.; Gogotsi, Y.; Seh, Z.W. Fundamentals of MXene Synthesis. Nat. Synth. 2022, 1, 601–614. [Google Scholar] [CrossRef]
  15. Wei, Y.; Zhang, P.; Soomro, R.A.; Zhu, Q.; Xu, B. Advances in the Synthesis of 2D MXenes. Adv. Mater. 2021, 33, 2103148. [Google Scholar] [CrossRef] [PubMed]
  16. Yang, Z.; Liu, A.; Wang, C.; Liu, F.; He, J.; Li, S.; Wang, J.; You, R.; Yan, X.; Sun, P.; et al. Improvement of Gas and Humidity Sensing Properties of Organ-like MXene by Alkaline Treatment. ACS Sens. 2019, 4, 1261–1269. [Google Scholar] [CrossRef] [PubMed]
  17. Caffrey, N.M. Effect of Mixed Surface Terminations on the Structural and Electrochemical Properties of Two-Dimensional Ti3C2T2 and V2CT2 MXenes Multilayers. Nanoscale 2018, 10, 13520–13530. [Google Scholar] [CrossRef]
  18. Seh, Z.W.; Fredrickson, K.D.; Anasori, B.; Kibsgaard, J.; Strickler, A.L.; Lukatskaya, M.R.; Gogotsi, Y.; Jaramillo, T.F.; Vojvodic, A. Two-Dimensional Molybdenum Carbide (MXene) as an Efficient Electrocatalyst for Hydrogen Evolution. ACS Energy Lett. 2016, 1, 589–594. [Google Scholar] [CrossRef]
  19. Rasheed, P.A.; Pandey, R.P.; Banat, F.; Hasan, S.W. Recent Advances in Niobium MXenes: Synthesis, Properties, and Emerging Applications. Matter 2022, 5, 546–572. [Google Scholar] [CrossRef]
  20. Jiang, L.; Zhou, D.; Yang, J.; Zhou, S.; Wang, H.; Yuan, X.; Liang, J.; Li, X.; Chen, Y.; Li, H. 2D Single- and Few-Layered MXenes: Synthesis, Applications and Perspectives. J. Mater. Chem. A Mater. 2022, 10, 13651–13672. [Google Scholar] [CrossRef]
  21. Pogorielov, M.; Smyrnova, K.; Kyrylenko, S.; Gogotsi, O.; Zahorodna, V.; Pogrebnjak, A. MXenes—A New Class of Two-Dimensional Materials: Structure, Properties and Potential Applications. Nanomaterials 2021, 11, 3412. [Google Scholar] [CrossRef]
  22. Ghidiu, M.; Naguib, M.; Shi, C.; Mashtalir, O.; Pan, L.M.; Zhang, B.; Yang, J.; Gogotsi, Y.; Billinge, S.J.L.; Barsoum, M.W. Synthesis and Characterization of Two-Dimensional Nb4C3 (MXene). Chem. Commun. 2014, 50, 9517–9520. [Google Scholar] [CrossRef]
  23. Lukatskaya, M.R.; Mashtalir, O.; Ren, C.E.; Dall’Agnese, Y.; Rozier, P.; Taberna, P.L.; Naguib, M.; Simon, P.; Barsoum, M.W.; Gogotsi, Y. Cation Intercalation and High Volumetric Capacitance of Two-Dimensional Titanium Carbide. Science 2013, 341, 1502–1505. [Google Scholar] [CrossRef]
  24. Mashtalir, O.; Naguib, M.; Dyatkin, B.; Gogotsi, Y.; Barsoum, M.W. Kinetics of Aluminum Extraction from Ti3AlC2 in Hydrofluoric Acid. Mater. Chem. Phys. 2013, 139, 147–152. [Google Scholar] [CrossRef]
  25. Halim, J.; Kota, S.; Lukatskaya, M.R.; Naguib, M.; Zhao, M.-Q.; Moon, E.J.; Pitock, J.; Nanda, J.; May, S.J.; Gogotsi, Y.; et al. Synthesis and Characterization of 2D Molybdenum Carbide (MXene). Adv. Funct. Mater. 2016, 26, 3118–3127. [Google Scholar] [CrossRef]
  26. Li, T.; Yao, L.; Liu, Q.; Gu, J.; Luo, R.; Li, J.; Yan, X.; Wang, W.; Liu, P.; Chen, B.; et al. Fluorine-Free Synthesis of High-Purity Ti3C2Tx (T=OH, O) via Alkali Treatment. Angew. Chem. Int. Ed. 2018, 57, 6115–6119. [Google Scholar] [CrossRef] [PubMed]
  27. Li, G.; Tan, L.; Zhang, Y.; Wu, B.; Li, L. Highly Efficiently Delaminated Single-Layered MXene Nanosheets with Large Lateral Size. Langmuir 2017, 33, 9000–9006. [Google Scholar] [CrossRef] [PubMed]
  28. Li, Y.; Shao, H.; Lin, Z.; Lu, J.; Liu, L.; Duployer, B.; Persson, P.O.Å.; Eklund, P.; Hultman, L.; Li, M.; et al. A General Lewis Acidic Etching Route for Preparing MXenes with Enhanced Electrochemical Performance in Non-Aqueous Electrolyte. Nat. Mater. 2020, 19, 894–899. [Google Scholar] [CrossRef] [PubMed]
  29. Li, M.; Lu, J.; Luo, K.; Li, Y.; Chang, K.; Chen, K.; Zhou, J.; Rosen, J.; Hultman, L.; Eklund, P.; et al. Element Replacement Approach by Reaction with Lewis Acidic Molten Salts to Synthesize Nanolaminated MAX Phases and MXenes. J. Am. Chem. Soc. 2019, 141, 4730–4737. [Google Scholar] [CrossRef] [PubMed]
  30. Pang, S.-Y.; Wong, Y.-T.; Yuan, S.; Liu, Y.; Tsang, M.-K.; Yang, Z.; Huang, H.; Wong, W.-T.; Hao, J. Universal Strategy for HF-Free Facile and Rapid Synthesis of Two-Dimensional MXenes as Multifunctional Energy Materials. J. Am. Chem. Soc. 2019, 141, 9610–9616. [Google Scholar] [CrossRef] [PubMed]
  31. Yang, S.; Zhang, P.; Wang, F.; Ricciardulli, A.G.; Lohe, M.R.; Blom, P.W.M.; Feng, X. Fluoride-Free Synthesis of Two-Dimensional Titanium Carbide (MXene) Using A Binary Aqueous System. Angew. Chem. Int. Ed. 2018, 57, 15491–15495. [Google Scholar] [CrossRef]
  32. Shi, H.; Zhang, P.; Liu, Z.; Park, S.; Lohe, M.R.; Wu, Y.; Shaygan Nia, A.; Yang, S.; Feng, X. Ambient-Stable Two-Dimensional Titanium Carbide (MXene) Enabled by Iodine Etching. Angew. Chem. Int. Ed. 2021, 60, 8689–8693. [Google Scholar] [CrossRef] [PubMed]
  33. 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. Sustain. Mater. Technol. 2020, 25, e00156. [Google Scholar] [CrossRef]
  34. Marian, M.; Song, G.C.; Wang, B.; Fuenzalida, V.M.; Krauß, S.; Merle, B.; Tremmel, S.; Wartzack, S.; Yu, J.; Rosenkranz, A. Effective Usage of 2D MXene Nanosheets as Solid Lubricant—Influence of Contact Pressure and Relative Humidity. Appl. Surf. Sci. 2020, 531, 147311. [Google Scholar] [CrossRef]
  35. Wang, B.; Zhou, A.; Liu, F.; Cao, J.; Wang, L.; Hu, Q. Carbon Dioxide Adsorption of Two-Dimensional Carbide MXenes. J. Adv. Ceram. 2018, 7, 237–245. [Google Scholar] [CrossRef]
  36. 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, 9353–9358. [Google Scholar] [CrossRef] [PubMed]
  37. 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, 3501–3506. [Google Scholar] [CrossRef]
  38. Wang, H.; Zhang, J.; Wu, Y.; Huang, H.; Li, G.; Zhang, X.; Wang, Z. Surface Modified MXene Ti3C2 Multilayers by Aryl Diazonium Salts Leading to Large-Scale Delamination. Appl. Surf. Sci. 2016, 384, 287–293. [Google Scholar] [CrossRef]
  39. Pazniak, H.; Plugin, I.A.; Sheverdyaeva, P.M.; Rapenne, L.; Varezhnikov, A.S.; Agresti, A.; Pescetelli, S.; Moras, P.; Kostin, K.B.; Gorokhovsky, A.V.; et al. Alcohol Vapor Sensor Based on Quasi-2D Nb2O5 Derived from Oxidized Nb2CTz MXenes. Sensors 2024, 24, 38. [Google Scholar] [CrossRef]
  40. Jolly, S.; Paranthaman, M.P.; Naguib, M. Synthesis of Ti3C2Tz MXene from Low-Cost and Environmentally Friendly Precursors. Mater. Today Adv. 2021, 10, 100139. [Google Scholar] [CrossRef]
  41. Li, C.; Kota, S.; Hu, C.; Barsoum, M. On the Synthesis of Low-Cost, Titanium-Based MXenes. J. Ceram. Sci. Technol. 2016, 7, 7–10. [Google Scholar] [CrossRef]
  42. Ramanavicius, S.; Ramanavicius, A. Progress and Insights in the Application of MXenes as New 2D Nano-Materials Suitable for Biosensors and Biofuel Cell Design. Int. J. Mol. Sci. 2020, 21, 9224. [Google Scholar] [CrossRef]
  43. Guo, B.; He, S.; Yao, M.; Tan, Z.; Li, X.; Liu, M.; Yu, C.; Liang, L.; Zhao, Z.; Guo, Z.; et al. MXene-Containing Anisotropic Hydrogels Strain Sensors with Enhanced Sensing Performance for Human Motion Monitoring and Wireless Transmission. Chem. Eng. J. 2023, 461, 142099. [Google Scholar] [CrossRef]
  44. Yi, J.; Xianyu, Y. Gold Nanomaterials-Implemented Wearable Sensors for Healthcare Applications. Adv. Funct. Mater. 2022, 32, 2113012. [Google Scholar] [CrossRef]
  45. Wu, L.; Wang, L.; Guo, Z.; Luo, J.; Xue, H.; Gao, J. Durable and Multifunctional Superhydrophobic Coatings with Excellent Joule Heating and Electromagnetic Interference Shielding Performance for Flexible Sensing Electronics. ACS Appl. Mater. Interfaces 2019, 11, 34338–34347. [Google Scholar] [CrossRef] [PubMed]
  46. Tang, J.; Wu, Y.; Ma, S.; Yan, T.; Pan, Z. Flexible Strain Sensor Based on CNT/TPU Composite Nanofiber Yarn for Smart Sports Bandage. Compos. B Eng. 2022, 232, 109605. [Google Scholar] [CrossRef]
  47. Duan, Q.; Lan, B.; Lv, Y. Highly Dispersed, Adhesive Carbon Nanotube Ink for Strain and Pressure Sensors. ACS Appl. Mater. Interfaces 2022, 14, 1973–1982. [Google Scholar] [CrossRef]
  48. Zheng, X.; Cao, W.; Hong, X.; Zou, L.; Liu, Z.; Wang, P.; Li, C. Versatile Electronic Textile Enabled by a Mixed-Dimensional Assembly Strategy. Small 2023, 19, 2208134. [Google Scholar] [CrossRef]
  49. Li, X.; Zhao, B.; Qin, W.; Yang, M.; Feng, L.; Gu, C.; Tian, Z.; Liu, H.; Guo, X.; Zhang, Y.; et al. MXene Ti3C2Tx, EGaIn, and Carbon Nanotube Composites on Polyurethane Substrates for Strain Sensing, Electromagnetic Interference Shielding, and Joule Heating. ACS Appl. Nano Mater. 2023, 6, 14459–14468. [Google Scholar] [CrossRef]
  50. Pu, J.-H.; Zhao, X.; Zha, X.-J.; Bai, L.; Ke, K.; Bao, R.-Y.; Liu, Z.-Y.; Yang, M.-B.; Yang, W. Multilayer Structured AgNW/WPU-MXene Fiber Strain Sensors with Ultrahigh Sensitivity and a Wide Operating Range for Wearable Monitoring and Healthcare. J. Mater. Chem. A Mater. 2019, 7, 15913–15923. [Google Scholar] [CrossRef]
  51. Yang, Y.; Cao, Z.; He, P.; Shi, L.; Ding, G.; Wang, R.; Sun, J. Ti3C2Tx MXene-Graphene Composite Films for Wearable Strain Sensors Featured with High Sensitivity and Large Range of Linear Response. Nano Energy 2019, 66, 104134. [Google Scholar] [CrossRef]
  52. Cai, Y.; Shen, J.; Ge, G.; Zhang, Y.; Jin, W.; Huang, W.; Shao, J.; Yang, J.; Dong, X. Stretchable Ti3C2Tx MXene/Carbon Nanotube Composite Based Strain Sensor with Ultrahigh Sensitivity and Tunable Sensing Range. ACS Nano 2018, 12, 56–62. [Google Scholar] [CrossRef]
  53. Xu, X.; Chen, Y.; He, P.; Wang, S.; Ling, K.; Liu, L.; Lei, P.; Huang, X.; Zhao, H.; Cao, J.; et al. Wearable CNT/Ti3C2Tx MXene/PDMS Composite Strain Sensor with Enhanced Stability for Real-Time Human Healthcare Monitoring. Nano Res. 2021, 14, 2875–2883. [Google Scholar] [CrossRef]
  54. Cao, J.; Jiang, Y.; Li, X.; Yuan, X.; Zhang, J.; He, Q.; Ye, F.; Luo, G.; Guo, S.; Zhang, Y.; et al. A Flexible and Stretchable MXene/Waterborne Polyurethane Composite-Coated Fiber Strain Sensor for Wearable Motion and Healthcare Monitoring. Sensors 2024, 24, 271. [Google Scholar] [CrossRef]
  55. Yang, H.; Xiao, X.; Li, Z.; Li, K.; Cheng, N.; Li, S.; Low, J.H.; Jing, L.; Fu, X.; Achavananthadith, S.; et al. Wireless Ti3C2Tx MXene Strain Sensor with Ultrahigh Sensitivity and Designated Working Windows for Soft Exoskeletons. ACS Nano 2020, 14, 11860–11875. [Google Scholar] [CrossRef]
  56. Liao, H.; Guo, X.; Wan, P.; Yu, G. Conductive MXene Nanocomposite Organohydrogel for Flexible, Healable, Low-Temperature Tolerant Strain Sensors. Adv. Funct. Mater. 2019, 29, 1904507. [Google Scholar] [CrossRef]
  57. Yang, C.; Zhang, D.; Wang, D.; Luan, H.; Chen, X.; Yan, W. In Situ Polymerized MXene/Polypyrrole/Hydroxyethyl Cellulose-Based Flexible Strain Sensor Enabled by Machine Learning for Handwriting Recognition. ACS Appl. Mater. Interfaces 2023, 15, 5811–5821. [Google Scholar] [CrossRef] [PubMed]
  58. Wang, H.; Zhou, R.; Li, D.; Zhang, L.; Ren, G.; Wang, L.; Liu, J.; Wang, D.; Tang, Z.; Lu, G.; et al. High-Performance Foam-Shaped Strain Sensor Based on Carbon Nanotubes and Ti3C2Tx MXene for the Monitoring of Human Activities. ACS Nano 2021, 15, 9690–9700. [Google Scholar] [CrossRef] [PubMed]
  59. Yang, K.; Yin, F.; Xia, D.; Peng, H.; Yang, J.; Yuan, W. A Highly Flexible and Multifunctional Strain Sensor Based on a Network-Structured MXene/Polyurethane Mat with Ultra-High Sensitivity and a Broad Sensing Range. Nanoscale 2019, 11, 9949–9957. [Google Scholar] [CrossRef] [PubMed]
  60. Chao, M.; Wang, Y.; Ma, D.; Wu, X.; Zhang, W.; Zhang, L.; Wan, P. Wearable MXene Nanocomposites-Based Strain Sensor with Tile-like Stacked Hierarchical Microstructure for Broad-Range Ultrasensitive Sensing. Nano Energy 2020, 78, 105187. [Google Scholar] [CrossRef]
  61. Bu, Y.; Shen, T.; Yang, W.; Yang, S.; Zhao, Y.; Liu, H.; Zheng, Y.; Liu, C.; Shen, C. Ultrasensitive Strain Sensor Based on Superhydrophobic Microcracked Conductive Ti3C2Tx MXene/Paper for Human-Motion Monitoring and E-Skin. Sci. Bull. 2021, 66, 1849–1857. [Google Scholar] [CrossRef] [PubMed]
  62. Fu, X.; Li, L.; Chen, S.; Xu, H.; Li, J.; Shulga, V.; Han, W. Knitted Ti3C2Tx MXene Based Fiber Strain Sensor for Human–Computer Interaction. J. Colloid. Interface Sci. 2021, 604, 643–649. [Google Scholar] [CrossRef] [PubMed]
  63. Kong, D.; El-Bahy, Z.M.; Algadi, H.; Li, T.; El-Bahy, S.M.; Nassan, M.A.; Li, J.; Faheim, A.A.; Li, A.; Xu, C.; et al. Highly Sensitive Strain Sensors with Wide Operation Range from Strong MXene-Composited Polyvinyl Alcohol/Sodium Carboxymethylcellulose Double Network Hydrogel. Adv. Compos. Hybrid. Mater. 2022, 5, 1976–1987. [Google Scholar] [CrossRef]
  64. Chae, A.; Murali, G.; Lee, S.-Y.; Gwak, J.; Kim, S.J.; Jeong, Y.J.; Kang, H.; Park, S.; Lee, A.S.; Koh, D.-Y.; et al. Highly Oxidation-Resistant and Self-Healable MXene-Based Hydrogels for Wearable Strain Sensor. Adv. Funct. Mater. 2023, 33, 2213382. [Google Scholar] [CrossRef]
  65. Peng, W.; Pan, X.; Liu, X.; Gao, Y.; Lu, T.; Li, J.; Xu, M.; Pan, L. A Moisture Self-Regenerative, Ultra-Low Temperature Anti-Freezing and Self-Adhesive Polyvinyl Alcohol/Polyacrylamide/CaCl2/MXene Ionotronics Hydrogel for Bionic Skin Strain Sensor. J. Colloid. Interface Sci. 2023, 634, 782–792. [Google Scholar] [CrossRef] [PubMed]
  66. Dong, L.; Zhou, X.; Zheng, S.; Luo, Z.; Nie, Y.; Feng, X.; Zhu, J.; Wang, Z.; Lu, X.; Mu, L. Liquid Metal @ Mxene Spring Supports Ionic Gel with Excellent Mechanical Properties for High-Sensitivity Wearable Strain Sensor. Chem. Eng. J. 2023, 458, 141370. [Google Scholar] [CrossRef]
  67. Liang, J.; He, J.; Xin, Y.; Gao, W.; Zeng, G.; He, X. MXene Reinforced PAA/PEDOT:PSS/MXene Conductive Hydrogel for Highly Sensitive Strain Sensors. Macromol. Mater. Eng. 2023, 308, 2200519. [Google Scholar] [CrossRef]
  68. Qin, W.; Geng, J.; Lin, C.; Li, G.; Peng, H.; Xue, Y.; Zhou, B.; Liu, G. Flexible Multifunctional TPU Strain Sensors with Improved Sensitivity and Wide Sensing Range Based on MXene/AgNWs. J. Mater. Sci. Mater. Electron. 2023, 34, 564. [Google Scholar] [CrossRef]
  69. Seesaard, T.; Wongchoosuk, C. Flexible and Stretchable Pressure Sensors: From Basic Principles to State-of-the-Art Applications. Micromachines 2023, 14, 1638. [Google Scholar] [CrossRef]
  70. Ma, Z.; Zhang, Y.; Zhang, K.; Deng, H.; Fu, Q. Recent Progress in Flexible Capacitive Sensors: Structures and Properties. Nano Mater. Sci. 2023, 5, 265–277. [Google Scholar] [CrossRef]
  71. Wu, Y.; Ma, Y.; Zheng, H.; Ramakrishna, S. Piezoelectric Materials for Flexible and Wearable Electronics: A Review. Mater. Des. 2021, 211, 110164. [Google Scholar] [CrossRef]
  72. Garcia, C.; Trendafilova, I.; Guzman de Villoria, R.; Sanchez del Rio, J. Self-Powered Pressure Sensor Based on the Triboelectric Effect and Its Analysis Using Dynamic Mechanical Analysis. Nano Energy 2018, 50, 401–409. [Google Scholar] [CrossRef]
  73. Trung, T.Q.; Lee, N.-E. Flexible and Stretchable Physical Sensor Integrated Platforms for Wearable Human-Activity Monitoringand Personal Healthcare. Adv. Mater. 2016, 28, 4338–4372. [Google Scholar] [CrossRef] [PubMed]
  74. Liang, G.; Ruan, Z.; Liu, Z.; Li, H.; Wang, Z.; Tang, Z.; Mo, F.; Yang, Q.; Ma, L.; Wang, D.; et al. Toward Multifunctional and Wearable Smart Skins with Energy-Harvesting, Touch-Sensing, and Exteroception-Visualizing Capabilities by an All-Polymer Design. Adv. Electron. Mater. 2019, 5, 1900553. [Google Scholar] [CrossRef]
  75. Fiorillo, A.S.; Critello, C.D.; Pullano, S.A. Theory, Technology and Applications of Piezoresistive Sensors: A Review. Sens. Actuators A Phys. 2018, 281, 156–175. [Google Scholar] [CrossRef]
  76. Li, H.; Wu, K.; Xu, Z.; Wang, Z.; Meng, Y.; Li, L. Ultrahigh-Sensitivity Piezoresistive Pressure Sensors for Detection of Tiny Pressure. ACS Appl. Mater. Interfaces 2018, 10, 20826–20834. [Google Scholar] [CrossRef]
  77. Zhang, Y.; Wang, L.; Zhao, L.; Wang, K.; Zheng, Y.; Yuan, Z.; Wang, D.; Fu, X.; Shen, G.; Han, W. Flexible Self-Powered Integrated Sensing System with 3D Periodic Ordered Black Phosphorus@MXene Thin-Films. Adv. Mater. 2021, 33, 2007890. [Google Scholar] [CrossRef] [PubMed]
  78. Zhao, Y.; Liu, Y.; Li, Y.; Hao, Q. Development and Application of Resistance Strain Force Sensors. Sensors 2020, 20, 5826. [Google Scholar] [CrossRef]
  79. Wu, Z.; Wei, L.; Tang, S.; Xiong, Y.; Qin, X.; Luo, J.; Fang, J.; Wang, X. Recent Progress in Ti3C2Tx MXene-Based Flexible Pressure Sensors. ACS Nano 2021, 15, 18880–18894. [Google Scholar] [CrossRef]
  80. Zou, Y.; Chen, Z.; Guo, X.; Peng, Z.; Yu, C.; Zhong, W. Mechanically Robust and Elastic Graphene/Aramid Nanofiber/Polyaniline Nanotube Aerogels for Pressure Sensors. ACS Appl. Mater. Interfaces 2022, 14, 17858–17868. [Google Scholar] [CrossRef]
  81. Fu, X.; Li, J.; Li, D.; Zhao, L.; Yuan, Z.; Shulga, V.; Han, W.; Wang, L. MXene/ZIF-67/PAN Nanofiber Film for Ultra-Sensitive Pressure Sensors. ACS Appl. Mater. Interfaces 2022, 14, 12367–12374. [Google Scholar] [CrossRef] [PubMed]
  82. Lei, D.; Liu, N.; Su, T.; Zhang, Q.; Wang, L.; Ren, Z.; Gao, Y. Roles of MXene in Pressure Sensing: Preparation, Composite Structure Design, and Mechanism. Adv. Mater. 2022, 34, 2110608. [Google Scholar] [CrossRef] [PubMed]
  83. Zheng, Y.; Yin, R.; Zhao, Y.; Liu, H.; Zhang, D.; Shi, X.; Zhang, B.; Liu, C.; Shen, C. Conductive MXene/Cotton Fabric Based Pressure Sensor with Both High Sensitivity and Wide Sensing Range for Human Motion Detection and E-Skin. Chem. Eng. J. 2021, 420, 127720. [Google Scholar] [CrossRef]
  84. Xu, Z.; Zhang, D.; Li, Z.; Du, C.; Yang, Y.; Zhang, B.; Zhao, W. Waterproof Flexible Pressure Sensors Based on Electrostatic Self-Assembled MXene/NH2-CNTs for Motion Monitoring and Electronic Skin. ACS Appl. Mater. Interfaces 2023, 15, 32569–32579. [Google Scholar] [CrossRef] [PubMed]
  85. Wang, L.; Zhang, M.; Yang, B.; Tan, J.; Ding, X. Highly Compressible, Thermally Stable, Light-Weight, and Robust Aramid Nanofibers/Ti3AlC2 MXene Composite Aerogel for Sensitive Pressure Sensor. ACS Nano 2020, 14, 10633–10647. [Google Scholar] [CrossRef]
  86. Yin, T.; Cheng, Y.; Hou, Y.; Sun, L.; Ma, Y.; Su, J.; Zhang, Z.; Liu, N.; Li, L.; Gao, Y. 3D Porous Structure in MXene/PANI Foam for a High-Performance Flexible Pressure Sensor. Small 2022, 18, 2204806. [Google Scholar] [CrossRef]
  87. Li, G.; Sun, F.; Zhao, S.; Xu, R.; Wang, H.; Qu, L.; Tian, M. Autonomous Electroluminescent Textile for Visual Interaction and Environmental Warning. Nano Lett. 2023, 23, 8436–8444. [Google Scholar] [CrossRef]
  88. Guo, D.; Mu, C.; Liu, Q.; Wang, B.; Xiang, J.; Nie, A.; Zhai, K.; Shu, Y.; Xue, T.; Wen, F.; et al. Aramid Nanofiber/Polypyrrole Composite Films for Broadband EMI Shielding, Wearable Electronics, Joule Heating, and Photothermal Conversion. ACS Appl. Nano Mater. 2023, 6, 15108–15118. [Google Scholar] [CrossRef]
  89. Ma, Y.; Liu, N.; Li, L.; Hu, X.; Zou, Z.; Wang, J.; Luo, S.; Gao, Y. A Highly Flexible and Sensitive Piezoresistive Sensor Based on MXene with Greatly Changed Interlayer Distances. Nat. Commun. 2017, 8, 1207. [Google Scholar] [CrossRef]
  90. Pei, Y.; Wang, K.; Hui, Z.; Pan, H.; Zhou, J.; Sun, G. Blowing up Ti3C2TX MXene Membrane for Robust Sound Detection. Appl. Phys. Lett. 2023, 122, 083501. [Google Scholar] [CrossRef]
  91. Gou, G.-Y.; Li, X.-S.; Jian, J.-M.; Tian, H.; Wu, F.; Ren, J.; Geng, X.-S.; Xu, J.-D.; Qiao, Y.-C.; Yan, Z.-Y.; et al. Two-Stage Amplification of an Ultrasensitive MXene-Based Intelligent Artificial Eardrum. Sci. Adv. 2024, 8, eabn2156. [Google Scholar] [CrossRef]
  92. Bilibana, M.P. Electrochemical Properties of MXenes and Applications. Adv. Sens. Energy Mater. 2023, 2, 100080. [Google Scholar] [CrossRef]
  93. Hajian, S.; Maddipatla, D.; Narakathu, B.B.; Atashbar, M.Z. MXene-Based Flexible Sensors: A Review. Front. Sens. 2022, 3, 1006749. [Google Scholar] [CrossRef]
  94. Qi, Z.; Zhang, T.; Zhang, X.-D.; Xu, Q.; Cao, K.; Chen, R. MXene-Based Flexible Pressure Sensor with Piezoresistive Properties Significantly Enhanced by Atomic Layer Infiltration. Nano Mater. Sci. 2023, 5, 439–446. [Google Scholar] [CrossRef]
  95. Li, C.; Gu, H.; Ji, Z.; Zhang, L.; Xu, R.; Gao, J.; Jiang, Y. Hierarchically Structured MXene Nanosheets on Carbon Sponges with a Synergistic Effect of Electrostatic Adsorption and Capillary Action for Highly Sensitive Pressure Sensors. ACS Appl. Nano Mater. 2023, 6, 13482–13491. [Google Scholar] [CrossRef]
  96. Cheng, H.; Yang, C.; Chu, J.; Zhou, H.; Wang, C. Multifunctional Ti3C2Tx MXene/Nanospheres/Ti3C2Tx MXene/Thermoplastic Polyurethane Electrospinning Membrane Inspired by Bean Pod Structure for EMI Shielding and Pressure Sensing. Sens. Actuators A Phys. 2023, 353, 114226. [Google Scholar] [CrossRef]
  97. Xie, Y.; Cheng, Y.; Ma, Y.; Wang, J.; Zou, J.; Wu, H.; Yue, Y.; Li, B.; Gao, Y.; Zhang, X.; et al. 3D MXene-Based Flexible Network for High-Performance Pressure Sensor with a Wide Temperature Range. Adv. Sci. 2023, 10, 2205303. [Google Scholar] [CrossRef]
  98. Li, L.; Cheng, Y.; Cao, H.; Liang, Z.; Liu, Z.; Yan, S.; Li, L.; Jia, S.; Wang, J.; Gao, Y. MXene/RGO/PS Spheres Multiple Physical Networks as High-Performance Pressure Sensor. Nano Energy 2022, 95, 106986. [Google Scholar] [CrossRef]
  99. Yang, J.; Li, H.; Cheng, J.; He, T.; Li, J.; Wang, B. Nanocellulose Intercalation to Boost the Performance of MXene Pressure Sensor for Human Interactive Monitoring. J. Mater. Sci. 2021, 56, 13859–13873. [Google Scholar] [CrossRef]
  100. Cao, Z.; Yang, Y.; Zheng, Y.; Wu, W.; Xu, F.; Wang, R.; Sun, J. Highly Flexible and Sensitive Temperature Sensors Based on Ti3C2Tx (MXene) for Electronic Skin. J. Mater. Chem. A Mater. 2019, 7, 25314–25323. [Google Scholar] [CrossRef]
  101. Zhao, L.; Fu, X.; Xu, H.; Zheng, Y.; Han, W.; Wang, L. Tissue-Like Sodium Alginate-Coated 2D MXene-Based Flexible Temperature Sensors for Full-Range Temperature Monitoring. Adv. Mater. Technol. 2022, 7, 2101740. [Google Scholar] [CrossRef]
  102. Wang, Z.; Zou, X.; Yang, Z.; Wang, J.; Chen, Q.; Pei, C.; Wang, Z.; Dong, M.; Zhang, J.; Li, H.; et al. Highly Sensitive Temperature Detection Based on a Frost- and Dehydration-Resistive Ion-Doped Hydrogel-MXene Composite. ACS Appl. Mater. Interfaces 2023, 15, 35525–35533. [Google Scholar] [CrossRef]
  103. Liu, H.; Yang, N.; Zhang, Q.; Wang, F.; Yan, X.; Zhang, X.; Cheng, T. A High-Performance Wearable Ag/Ti3C2Tx MXene-Based Fiber Sensor for Temperature Sensing, Pressure Sensing, and Human Motion Detection. IEEE Trans. Instrum. Meas. 2022, 71, 9512608. [Google Scholar] [CrossRef]
  104. Saeidi-Javash, M.; Du, Y.; Zeng, M.; Wyatt, B.C.; Zhang, B.; Kempf, N.; Anasori, B.; Zhang, Y. All-Printed MXene–Graphene Nanosheet-Based Bimodal Sensors for Simultaneous Strain and Temperature Sensing. ACS Appl. Electron. Mater. 2021, 3, 2341–2348. [Google Scholar] [CrossRef]
  105. Zuo, Y.; Gao, Y.; Qin, S.; Wang, Z.; Zhou, D.; Li, Z.; Yu, Y.; Shao, M.; Zhang, X. Broadband Multi-Wavelength Optical Sensing Based on Photothermal Effect of 2D MXene Films. Nanophotonics 2020, 9, 123–131. [Google Scholar] [CrossRef]
  106. Chen, S.; Ran, J.; Zheng, T.; Wu, Q. Ultracompact MXene V2C-Improved Temperature Sensor by a Runway-Type Microfiber Knot Resonator. Nanomaterials 2023, 13, 2354. [Google Scholar] [CrossRef] [PubMed]
  107. Wu, Q.; Ran, J.; Zheng, T.; Wu, H.; Liao, Y.; Wang, F.; Chen, S. MXene V2C-Coated Runway-Type Microfiber Knot Resonator for an All-Optical Temperature Sensor. RSC Adv. 2023, 13, 19366–19372. [Google Scholar] [CrossRef] [PubMed]
  108. Yuan, X.; Zhang, M.; Wu, Y. MXene-Based Sensors for Detecting Human Physiological Information. Sens. Mater. 2020, 32, 4047–4065. [Google Scholar] [CrossRef]
  109. Tai, Y.; Lubineau, G. Human-Finger Electronics Based on Opposing Humidity-Resistance Responses in Carbon Nanofilms. Small 2017, 13, 1603486. [Google Scholar] [CrossRef] [PubMed]
  110. Wu, J.; Wu, Z.; Xu, H.; Wu, Q.; Liu, C.; Yang, B.-R.; Gui, X.; Xie, X.; Tao, K.; Shen, Y.; et al. An Intrinsically Stretchable Humidity Sensor Based on Anti-Drying, Self-Healing and Transparent Organohydrogels. Mater. Horiz. 2019, 6, 595–603. [Google Scholar] [CrossRef]
  111. Wu, J.; Sun, Y.-M.; Wu, Z.; Li, X.; Wang, N.; Tao, K.; Wang, G.P. Carbon Nanocoil-Based Fast-Response and Flexible Humidity Sensor for Multifunctional Applications. ACS Appl. Mater. Interfaces 2019, 11, 4242–4251. [Google Scholar] [CrossRef] [PubMed]
  112. Li, J.; Yuan, X.; Lin, C.; Yang, Y.; Xu, L.; Du, X.; Xie, J.; Lin, J.; Sun, J. Achieving High Pseudocapacitance of 2D Titanium Carbide (MXene) by Cation Intercalation and Surface Modification. Adv. Energy Mater. 2017, 7, 1602725. [Google Scholar] [CrossRef]
  113. Célérier, S.; Hurand, S.; Garnero, C.; Morisset, S.; Benchakar, M.; Habrioux, A.; Chartier, P.; Mauchamp, V.; Findling, N.; Lanson, B.; et al. Hydration of Ti3C2Tx MXene: An Interstratification Process with Major Implications on Physical Properties. Chem. Mater. 2019, 31, 454–461. [Google Scholar] [CrossRef]
  114. Dai, J.; Zhang, T.; Qi, R.; Zhao, H.; Fei, T.; Lu, G. LiCl Loaded Cross-Linked Polymer Composites by Click Reaction for Humidity Sensing. Sens. Actuators B Chem. 2017, 253, 361–367. [Google Scholar] [CrossRef]
  115. Zhang, Z.; Huang, J.; Dong, B.; Yuan, Q.; He, Y.; Wolfbeis, O.S. Rational Tailoring of ZnSnO3/TiO2 Heterojunctions with Bioinspired Surface Wettability for High-Performance Humidity Nanosensors. Nanoscale 2015, 7, 4149–4155. [Google Scholar] [CrossRef] [PubMed]
  116. Muckley, E.S.; Naguib, M.; Ivanov, I.N. Multi-Modal, Ultrasensitive, Wide-Range Humidity Sensing with Ti3C2 Film. Nanoscale 2018, 10, 21689–21695. [Google Scholar] [CrossRef]
  117. Wu, J.; Lu, P.; Dai, J.; Zheng, C.; Zhang, T.; Yu, W.W.; Zhang, Y. High Performance Humidity Sensing Property of Ti3C2Tx MXene-Derived Ti3C2Tx/K2Ti4O9 Composites. Sens. Actuators B Chem. 2021, 326, 128969. [Google Scholar] [CrossRef]
  118. An, H.; Habib, T.; Shah, S.; Gao, H.; Patel, A.; Echols, I.; Zhao, X.; Radovic, M.; Green, M.J.; Lutkenhaus, J.L. Water Sorption in MXene/Polyelectrolyte Multilayers for Ultrafast Humidity Sensing. ACS Appl. Nano Mater. 2019, 2, 948–955. [Google Scholar] [CrossRef]
  119. Lu, Y.; Wang, M.-Y.; Wang, D.-Y.; Sun, Y.-H.; Liu, Z.-H.; Gao, R.-K.; Yu, L.-D.; Zhang, D.-Z. Flexible Impedance Sensor Based on Ti3C2Tx MXene and Graphitic Carbon Nitride Nanohybrid for Humidity-Sensing Application with Ultrahigh Response. Rare Met. 2023, 42, 2204–2213. [Google Scholar] [CrossRef]
  120. Han, M.; Shen, W. Nacre-Inspired Cellulose Nanofiber/MXene Flexible Composite Film with Mechanical Robustness for Humidity Sensing. Carbohydr. Polym. 2022, 298, 120109. [Google Scholar] [CrossRef]
  121. Li, R.; Fan, Y.; Ma, Z.; Zhang, D.; Liu, Y.; Xu, J. Controllable Preparation of Ultrathin MXene Nanosheets and Their Excellent QCM Humidity Sensing Properties Enhanced by Fluoride Doping. Microchim. Acta 2021, 188, 81. [Google Scholar] [CrossRef]
  122. Muckley, E.S.; Naguib, M.; Wang, H.-W.; Vlcek, L.; Osti, N.C.; Sacci, R.L.; Sang, X.; Unocic, R.R.; Xie, Y.; Tyagi, M.; et al. Multimodality of Structural, Electrical, and Gravimetric Responses of Intercalated MXenes to Water. ACS Nano 2017, 11, 11118–11126. [Google Scholar] [CrossRef]
  123. Li, N.; Jiang, Y.; Xiao, Y.; Meng, B.; Xing, C.; Zhang, H.; Peng, Z. A Fully Inkjet-Printed Transparent Humidity Sensor Based on a Ti3C2/Ag Hybrid for Touchless Sensing of Finger Motion. Nanoscale 2019, 11, 21522–21531. [Google Scholar] [CrossRef]
  124. Sun, B.; Lv, H.; Liu, Z.; Wang, J.; Bai, X.; Zhang, Y.; Chen, J.; Kan, K.; Shi, K. Co3O4@PEI/Ti3C2Tx MXene Nanocomposites for a Highly Sensitive NOx Gas Sensor with a Low Detection Limit. J. Mater. Chem. A Mater. 2021, 9, 6335–6344. [Google Scholar] [CrossRef]
  125. Kim, S.J.; Koh, H.-J.; Ren, C.E.; Kwon, O.; Maleski, K.; Cho, S.-Y.; Anasori, B.; Kim, C.-K.; Choi, Y.-K.; Kim, J.; et al. Metallic Ti3C2Tx MXene Gas Sensors with Ultrahigh Signal-to-Noise Ratio. ACS Nano 2018, 12, 986–993. [Google Scholar] [CrossRef]
  126. John, R.A.B.; Vijayan, K.; Septiani, N.L.W.; Hardiansyah, A.; Kumar, A.R.; Yuliarto, B.; Hermawan, A. Gas-Sensing Mechanisms and Performances of MXenes and MXene-Based Heterostructures. Sensors 2023, 23, 8674. [Google Scholar] [CrossRef]
  127. Lee, E.; VahidMohammadi, A.; Prorok, B.C.; Yoon, Y.S.; Beidaghi, M.; Kim, D.-J. Room Temperature Gas Sensing of Two-Dimensional Titanium Carbide (MXene). ACS Appl. Mater. Interfaces 2017, 9, 37184–37190. [Google Scholar] [CrossRef] [PubMed]
  128. Wu, M.; He, M.; Hu, Q.; Wu, Q.; Sun, G.; Xie, L.; Zhang, Z.; Zhu, Z.; Zhou, A. Ti3C2 MXene-Based Sensors with High Selectivity for NH3 Detection at Room Temperature. ACS Sens. 2019, 4, 2763–2770. [Google Scholar] [CrossRef]
  129. Lee, S.H.; Eom, W.; Shin, H.; Ambade, R.B.; Bang, J.H.; Kim, H.W.; Han, T.H. Room-Temperature, Highly Durable Ti3C2Tx MXene/Graphene Hybrid Fibers for NH3 Gas Sensing. ACS Appl. Mater. Interfaces 2020, 12, 10434–10442. [Google Scholar] [CrossRef] [PubMed]
  130. Shuvo, S.N.; Ulloa Gomez, A.M.; Mishra, A.; Chen, W.Y.; Dongare, A.M.; Stanciu, L.A. Sulfur-Doped Titanium Carbide MXenes for Room-Temperature Gas Sensing. ACS Sens. 2020, 5, 2915–2924. [Google Scholar] [CrossRef] [PubMed]
  131. Bhardwaj, R.; Hazra, A. MXene-Based Gas Sensors. J. Mater. Chem. C Mater. 2021, 9, 15735–15754. [Google Scholar] [CrossRef]
  132. Ismail, A.H.; Mohd Yahya, N.A.; Mahdi, M.A.; Yaacob, M.H.; Sulaiman, Y. Gasochromic Response of Optical Sensing Platform Integrated with Polyaniline and Poly(3,4-Ethylenedioxythiophene) Exposed to NH3 Gas. Polymer 2020, 192, 122313. [Google Scholar] [CrossRef]
  133. Ahamad, T.; Naushad, M.; Alshheri, S.M. Fabrication of Highly Porous N/S Doped Carbon Embedded with CuO/CuS Nanoparticles for NH3 Gas Sensing. Mater. Lett. 2020, 268, 127515. [Google Scholar] [CrossRef]
  134. Li, P.; Wang, B.; Qin, C.; Han, C.; Sun, L.; Wang, Y. Band-Gap-Tunable CeO2 Nanoparticles for Room-Temperature NH3 Gas Sensors. Ceram. Int. 2020, 46, 19232–19240. [Google Scholar] [CrossRef]
  135. Liu, M.; Ji, J.; Song, P.; Wang, J.; Wang, Q. Sensing Performance of α-Fe2O3/Ti3C2Tx MXene Nanocomposites to NH3 at Room Temperature. J. Alloys Compd. 2022, 898, 162812. [Google Scholar] [CrossRef]
  136. Xiao, B.; Li, Y.-C.; Yu, X.F.; Cheng, J. MXenes: Reusable Materials for NH3 Sensor or Capturer by Controlling the Charge Injection. Sens. Actuators B Chem. 2016, 235, 103–109. [Google Scholar] [CrossRef]
  137. Kim, S.J.; Choi, J.; Maleski, K.; Hantanasirisakul, K.; Jung, H.-T.; Gogotsi, Y.; Ahn, C.W. Interfacial Assembly of Ultrathin, Functional MXene Films. ACS Appl. Mater. Interfaces 2019, 11, 32320–32327. [Google Scholar] [CrossRef]
  138. Seekaew, Y.; Kamlue, S.; Wongchoosuk, C. Room-Temperature Ammonia Gas Sensor Based on Ti3C2Tx MXene/Graphene Oxide/CuO/ZnO Nanocomposite. ACS Appl. Nano Mater. 2023, 6, 9008–9020. [Google Scholar] [CrossRef]
  139. Gabriunaite, I.; Valiuniene, A.; Ramanavicius, S.; Ramanavicius, A. Biosensors Based on Bio-Functionalized Semiconducting Metal Oxides. Crit. Rev. Anal. Chem. 2022, 1–16. [Google Scholar] [CrossRef]
  140. Celiesiute, R.; Ramanaviciene, A.; Gicevicius, M.; Ramanavicius, A. Electrochromic Sensors Based on Conducting Polymers, Metal Oxides, and Coordination Complexes. Crit. Rev. Anal. Chem. 2019, 49, 195–208. [Google Scholar] [CrossRef]
  141. Ramanavicius, S.; Jagminas, A.; Ramanavicius, A. Gas Sensors Based on Titanium Oxides (Review). Coatings 2022, 12, 699. [Google Scholar] [CrossRef]
  142. Ramanavicius, S.; Ramanavicius, A. Insights in the Application of Stoichiometric and Non-Stoichiometric Titanium Oxides for the Design of Sensors for the Determination of Gases and VOCs (TiO2−x and TinO2n−1 vs. TiO2). Sensors 2020, 20, 6833. [Google Scholar] [CrossRef]
  143. Ramanavicius, S.; Tereshchenko, A.; Karpicz, R.; Ratautaite, V.; Bubniene, U.; Maneikis, A.; Jagminas, A.; Ramanavicius, A. TiO2−x/TiO2-Structure Based ‘Self-Heated’ Sensor for the Determination of Some Reducing Gases. Sensors 2020, 20, 74. [Google Scholar] [CrossRef] [PubMed]
  144. Petruleviciene, M.; Savickaja, I.; Juodkazyte, J.; Ramanavicius, A. Investigation of WO3 and BiVO4 Photoanodes for Photoelectrochemical Sensing of Xylene, Toluene and Methanol. Chemosensors 2023, 11, 552. [Google Scholar] [CrossRef]
  145. Ramanavičius, S.; Petrulevičienė, M.; Juodkazytė, J.; Grigucevičienė, A.; Ramanavičius, A. Selectivity of Tungsten Oxide Synthesized by Sol-Gel Method Towards Some Volatile Organic Compounds and Gaseous Materials in a Broad Range of Temperatures. Materials 2020, 13, 523. [Google Scholar] [CrossRef] [PubMed]
  146. Petruleviciene, M.; Juodkazyte, J.; Parvin, M.; Tereshchenko, A.; Ramanavicius, S.; Karpicz, R.; Samukaite-Bubniene, U.; Ramanavicius, A. Tuning the Photo-Luminescence Properties of WO3 Layers by the Adjustment of Layer Formation Conditions. Materials 2020, 13, 2814. [Google Scholar] [CrossRef] [PubMed]
  147. Liu, M.; Wang, J.; Song, P.; Ji, J.; Wang, Q. Metal-Organic Frameworks-Derived In2O3 Microtubes/Ti3C2Tx MXene Composites for NH3 Detection at Room Temperature. Sens. Actuators B Chem. 2022, 361, 131755. [Google Scholar] [CrossRef]
  148. Pasupuleti, K.S.; Thomas, A.M.; Vidyasagar, D.; Rao, V.N.; Yoon, S.-G.; Kim, Y.-H.; Kim, S.-G.; Kim, M.-D. ZnO@Ti3C2Tx MXene Hybrid Composite-Based Schottky-Barrier-Coated SAW Sensor for Effective Detection of Sub-Ppb-Level NH3 at Room Temperature under UV Illumination. ACS Mater. Lett. 2023, 5, 2739–2746. [Google Scholar] [CrossRef]
  149. Liu, M.; Sun, R.-Y.; Ding, Y.-L.; Wang, Q.; Song, P. Au/α-Fe2O3/Ti3C2Tx MXene Nanosheet Heterojunctions for High-Performance NH3 Gas Detection at Room Temperature. ACS Appl. Nano Mater. 2023, 6, 11856–11867. [Google Scholar] [CrossRef]
  150. Turemis, M.; Zappi, D.; Giardi, M.T.; Basile, G.; Ramanaviciene, A.; Kapralovs, A.; Ramanavicius, A.; Viter, R. ZnO/Polyaniline Composite Based Photoluminescence Sensor for the Determination of Acetic Acid Vapor. Talanta 2020, 211, 120658. [Google Scholar] [CrossRef]
  151. Ramanavicius, S.; Ramanavicius, A. Conducting Polymers in the Design of Biosensors and Biofuel Cells. Polymers 2021, 13, 49. [Google Scholar] [CrossRef]
  152. Ramanavicius, S.; Ramanavicius, A. Charge Transfer and Biocompatibility Aspects in Conducting Polymer-Based Enzymatic Biosensors and Biofuel Cells. Nanomaterials 2021, 11, 371. [Google Scholar] [CrossRef] [PubMed]
  153. Gicevicius, M.; Celiesiute, R.; Kucinski, J.; Ramanaviciene, A.; Bagdziunas, G.; Ramanavicius, A. Analytical Evaluation of Optical PH-Sensitivity of Polyaniline Layer Electrochemically Deposited on ITO Electrode. J. Electrochem. Soc. 2018, 165, H903. [Google Scholar] [CrossRef]
  154. Ramanavicius, S.; Jagminas, A.; Ramanavicius, A. Advances in Molecularly Imprinted Polymers Based Affinity Sensors (Review). Polymers 2021, 13, 974. [Google Scholar] [CrossRef] [PubMed]
  155. Gicevicius, M.; Kucinski, J.; Ramanaviciene, A.; Ramanavicius, A. Tuning the Optical PH Sensing Properties of Polyaniline-Based Layer by Electrochemical Copolymerization of Aniline with o-Phenylenediamine. Dye. Pigment. 2019, 170, 107457. [Google Scholar] [CrossRef]
  156. Ramanavicius, S.; Samukaite-Bubniene, U.; Ratautaite, V.; Bechelany, M.; Ramanavicius, A. Electrochemical Molecularly Imprinted Polymer Based Sensors for Pharmaceutical and Biomedical Applications (Review). J. Pharm. Biomed. Anal. 2022, 215, 114739. [Google Scholar] [CrossRef]
  157. Ramanavicius, S.; Ramanavicius, A. Development of Molecularly Imprinted Polymer Based Phase Boundaries for Sensors Design (Review). Adv. Colloid. Interface Sci. 2022, 305, 102693. [Google Scholar] [CrossRef]
  158. Jin, L.; Wu, C.; Wei, K.; He, L.; Gao, H.; Zhang, H.; Zhang, K.; Asiri, A.M.; Alamry, K.A.; Yang, L.; et al. Polymeric Ti3C2Tx MXene Composites for Room Temperature Ammonia Sensing. ACS Appl. Nano Mater. 2020, 3, 12071–12079. [Google Scholar] [CrossRef]
  159. Wang, X.; Zhang, D.; Zhang, H.; Gong, L.; Yang, Y.; Zhao, W.; Yu, S.; Yin, Y.; Sun, D. In Situ Polymerized Polyaniline/MXene (V2C) as Building Blocks of Supercapacitor and Ammonia Sensor Self-Powered by Electromagnetic-Triboelectric Hybrid Generator. Nano Energy 2021, 88, 106242. [Google Scholar] [CrossRef]
  160. Li, X.; Wang, H.; Li, H.; Chen, Y.; Ni, Y.; Xia, Y. Zr3C2O2 MXene as Promising Candidate for NH3 Sensor with High Sensitivity and Selectivity at Room Temperature. Appl. Surf. Sci. 2023, 624, 157125. [Google Scholar] [CrossRef]
  161. Gao, X.; Zhang, T. An Overview: Facet-Dependent Metal Oxide Semiconductor Gas Sensors. Sens. Actuators B Chem. 2018, 277, 604–633. [Google Scholar] [CrossRef]
  162. Zhang, D.; Wu, Z.; Zong, X. Flexible and Highly Sensitive H2S Gas Sensor Based on In-Situ Polymerized SNO2/RGO/PANI Ternary Nanocomposite with Application in Halitosis Diagnosis. Sens. Actuators B Chem. 2019, 289, 32–41. [Google Scholar] [CrossRef]
  163. Gui, Y.; Yang, L.; Tian, K.; Zhang, H.; Fang, S. P-Type Co3O4 Nanoarrays Decorated on the Surface of n-Type Flower-like WO3 Nanosheets for High-Performance Gas Sensing. Sens. Actuators B Chem. 2019, 288, 104–112. [Google Scholar] [CrossRef]
  164. Xu, Y.; Ma, T.; Zheng, L.; Zhao, Y.; Liu, X.; Zhang, J. Heterostructures of Hematite-Sensitized W18O49 Hollow Spheres for Improved Acetone Detection with Ultralow Detection Limit. Sens. Actuators B Chem. 2019, 288, 432–441. [Google Scholar] [CrossRef]
  165. Zhang, R.; Zhou, T.; Wang, L.; Zhang, T. Metal–Organic Frameworks-Derived Hierarchical Co3O4 Structures as Efficient Sensing Materials for Acetone Detection. ACS Appl. Mater. Interfaces 2018, 10, 9765–9773. [Google Scholar] [CrossRef]
  166. Wang, Z.; Wang, C. Is Breath Acetone a Biomarker of Diabetes? A Historical Review on Breath Acetone Measurements. J. Breath. Res. 2013, 7, 037109. [Google Scholar] [CrossRef] [PubMed]
  167. Sun, S.; Wang, M.; Chang, X.; Jiang, Y.; Zhang, D.; Wang, D.; Zhang, Y.; Lei, Y. W18O49/Ti3C2Tx Mxene Nanocomposites for Highly Sensitive Acetone Gas Sensor with Low Detection Limit. Sens. Actuators B Chem. 2020, 304, 127274. [Google Scholar] [CrossRef]
  168. Liu, M.; Wang, Z.; Song, P.; Yang, Z.; Wang, Q. Flexible MXene/RGO/CuO Hybrid Aerogels for High Performance Acetone Sensing at Room Temperature. Sens. Actuators B Chem. 2021, 340, 129946. [Google Scholar] [CrossRef]
  169. Majhi, S.M.; Ali, A.; Greish, Y.E.; El-Maghraby, H.F.; Mahmoud, S.T. V2CTX MXene-Based Hybrid Sensor with High Selectivity and Ppb-Level Detection for Acetone at Room Temperature. Sci. Rep. 2023, 13, 3114. [Google Scholar] [CrossRef] [PubMed]
  170. Zhao, X.; Zhang, G.; Wang, J.; Yuan, T.; Huang, J.; Wang, L.; Liu, Y.-N. Ru Nanoclusters Supported on Ti3C2Tx Nanosheets for Catalytic Hydrogenation of Quinolines. ACS Appl. Nano Mater. 2022, 5, 6213–6220. [Google Scholar] [CrossRef]
  171. Lee, J.T.; Wyatt, B.C.; Davis, G.A., Jr.; Masterson, A.N.; Pagan, A.L.; Shah, A.; Anasori, B.; Sardar, R. Covalent Surface Modification of Ti3C2Tx MXene with Chemically Active Polymeric Ligands Producing Highly Conductive and Ordered Microstructure Films. ACS Nano 2021, 15, 19600–19612. [Google Scholar] [CrossRef] [PubMed]
  172. Hermawan, A.; Zhang, B.; Taufik, A.; Asakura, Y.; Hasegawa, T.; Zhu, J.; Shi, P.; Yin, S. CuO Nanoparticles/Ti3C2Tx MXene Hybrid Nanocomposites for Detection of Toluene Gas. ACS Appl. Nano Mater. 2020, 3, 4755–4766. [Google Scholar] [CrossRef]
  173. Guo, W.; Surya, S.G.; Babar, V.; Ming, F.; Sharma, S.; Alshareef, H.N.; Schwingenschlögl, U.; Salama, K.N. Selective Toluene Detection with Mo2CTx MXene at Room Temperature. ACS Appl. Mater. Interfaces 2020, 12, 57218–57227. [Google Scholar] [CrossRef]
  174. Park, H.J.; Hong, S.Y.; Chun, D.H.; Kang, S.W.; Park, J.C.; Lee, D.-S. A Highly Susceptive Mesoporous Hematite Microcube Architecture for Sustainable P-Type Formaldehyde Gas Sensors. Sens. Actuators B Chem. 2019, 287, 437–444. [Google Scholar] [CrossRef]
  175. Fu, X.; Yang, P.; Xiao, X.; Zhou, D.; Huang, R.; Zhang, X.; Cao, F.; Xiong, J.; Hu, Y.; Tu, Y.; et al. Ultra-Fast and Highly Selective Room-Temperature Formaldehyde Gas Sensing of Pt-Decorated MoO3 Nanobelts. J. Alloys Compd. 2019, 797, 666–675. [Google Scholar] [CrossRef]
  176. Zhao, R.; Zhang, X.; Peng, S.; Hong, P.; Zou, T.; Wang, Z.; Xing, X.; Yang, Y.; Wang, Y. Shaddock Peels as Bio-Templates Synthesis of Cd-Doped SnO2 Nanofibers: A High Performance Formaldehyde Sensing Material. J. Alloys Compd. 2020, 813, 152170. [Google Scholar] [CrossRef]
  177. Shuaishuai, J.; Wan, X.; Cuimin, Z.; Feihu, L.; Bo, M.; Zili, Z. Improving the Formaldehyde Gas Sensing Performance of the ZnO/SnO2 Nanoparticles by PdO Decoration. J. Mater. Sci. Mater. Electron. 2020, 31, 684–692. [Google Scholar] [CrossRef]
  178. Chen, X.; Shen, Y.; Zhou, P.; Zhao, S.; Zhong, X.; Li, T.; Han, C.; Wei, D.; Meng, D. NO2 Sensing Properties of One-Pot-Synthesized ZnO Nanowires with Pd Functionalization. Sens. Actuators B Chem. 2019, 280, 151–161. [Google Scholar] [CrossRef]
  179. Zhang, D.; Mi, Q.; Wang, D.; Li, T. MXene/Co3O4 Composite Based Formaldehyde Sensor Driven by ZnO/MXene Nanowire Arrays Piezoelectric Nanogenerator. Sens. Actuators B Chem. 2021, 339, 129923. [Google Scholar] [CrossRef]
  180. Niu, G.; Zhang, M.; Wu, B.; Zhuang, Y.; Ramachandran, R.; Zhao, C.; Wang, F. Nanocomposites of Pre-Oxidized Ti3C2Tx MXene and SnO2 Nanosheets for Highly Sensitive and Stable Formaldehyde Gas Sensor. Ceram. Int. 2023, 49, 2583–2590. [Google Scholar] [CrossRef]
  181. Mirzaei, A.; Leonardi, S.G.; Neri, G. Detection of Hazardous Volatile Organic Compounds (VOCs) by Metal Oxide Nanostructures-Based Gas Sensors: A Review. Ceram. Int. 2016, 42, 15119–15141. [Google Scholar] [CrossRef]
  182. Wang, X.; Sun, K.; Li, K.; Li, X.; Gogotsi, Y. Ti3C2Tx/PEDOT:PSS Hybrid Materials for Room-Temperature Methanol Sensor. Chin. Chem. Lett. 2020, 31, 1018–1021. [Google Scholar] [CrossRef]
  183. Liu, M.; Wang, Z.; Song, P.; Yang, Z.; Wang, Q. In2O3 Nanocubes/Ti3C2Tx MXene Composites for Enhanced Methanol Gas Sensing Properties at Room Temperature. Ceram. Int. 2021, 47, 23028–23037. [Google Scholar] [CrossRef]
  184. Jiang, B.; Lu, J.; Han, W.; Sun, Y.; Wang, Y.; Cheng, P.; Zhang, H.; Wang, C.; Lu, G. Hierarchical Mesoporous Zinc Oxide Microspheres for Ethanol Gas Sensor. Sens. Actuators B Chem. 2022, 357, 131333. [Google Scholar] [CrossRef]
  185. Zhang, J.; Ma, S.; Wang, B.; Pei, S. Hydrothermal Synthesis of SnO2-CuO Composite Nanoparticles as a Fast-Response Ethanol Gas Sensor. J. Alloys Compd. 2021, 886, 161299. [Google Scholar] [CrossRef]
  186. Sharma, B.; Karuppasamy, K.; Vikraman, D.; Jo, E.-B.; Sivakumar, P.; Kim, H.-S. Porous, 3D-Hierarchical α-NiMoO4 Rectangular Nanosheets for Selective Conductometric Ethanol Gas Sensors. Sens. Actuators B Chem. 2021, 347, 130615. [Google Scholar] [CrossRef]
  187. Tseng, S.-F.; Chen, C.-W. Synthesis of Nanograiny SnO2 Films on Laser-Patterned Graphene/Ceramic Substrates for Low-Temperature Ethanol Gas Sensors. Ceram. Int. 2021, 47, 33498–33508. [Google Scholar] [CrossRef]
  188. Mao, J.N.; Hong, B.; Chen, H.D.; Gao, M.H.; Xu, J.C.; Han, Y.B.; Yang, Y.T.; Jin, H.X.; Jin, D.F.; Peng, X.L.; et al. Highly Improved Ethanol Gas Response of N-Type α-Fe2O3 Bunched Nanowires Sensor with High-Valence Donor-Doping. J. Alloys Compd. 2020, 827, 154248. [Google Scholar] [CrossRef]
  189. Zhang, S.; Song, P.; Sun, J.; Ding, Y.; Wang, Q. MoO3/Ti3C2Tx MXene Nanocomposites with Rapid Response for Enhanced Ethanol-Sensing at a Low Temperature. Sens. Actuators B Chem. 2023, 378, 133216. [Google Scholar] [CrossRef]
  190. Wang, C.; Li, R.; Feng, L.; Xu, J. The SnO2/MXene Composite Ethanol Sensor Based on MEMS Platform. Chemosensors 2022, 10, 109. [Google Scholar] [CrossRef]
  191. Ai, T.; Li, J.; Nie, S.; Yin, Y.; Lu, J.; Bao, S.; Yan, L. High-Performance H2 Gas Sensor Based on Mn-Doped α-Fe2O3 Polyhedrons from N, N-Dimethylformamide Assisted Hydrothermal Synthesis. Int. J. Hydrogen Energy 2022, 47, 20561–20571. [Google Scholar] [CrossRef]
  192. Long, Y.; Xu, H.; He, J.; Li, C.; Zhu, M. Piezoelectric Polarization of BiOCl via Capturing Mechanical Energy for Catalytic H2 Evolution. Surf. Interfaces 2022, 31, 102056. [Google Scholar] [CrossRef]
  193. Meng, X.; Bi, M.; Xiao, Q.; Gao, W. Ultrasensitive Gas Sensor Based on Pd/SnS2/SnO2 Nanocomposites for Rapid Detection of H2. Sens. Actuators B Chem. 2022, 359, 131612. [Google Scholar] [CrossRef]
  194. Zhu, S.; Tian, Q.; Wu, G.; Bian, W.; Sun, N.; Wang, X.; Li, C.; Zhang, Y.; Dou, H.; Gong, C.; et al. Highly Sensitive and Stable H2 Gas Sensor Based on P-PdO-n-WO3-Heterostructure-Homogeneously-Dispersing Thin Film. Int. J. Hydrogen Energy 2022, 47, 17821–17834. [Google Scholar] [CrossRef]
  195. Wang, L.; Xiao, Z.; Yao, X.; Yu, X.; Tu, S.-T.; Chen, S. Pt=Pd Separation Modified Ti3C2TX MXene for Hydrogen Detection at Room Temperature. Int. J. Hydrogen Energy 2023, 48, 30205–30217. [Google Scholar] [CrossRef]
  196. Xia, Y.; Zhou, L.; Yang, J.; Du, P.; Xu, L.; Wang, J. Highly Sensitive and Fast Optoelectronic Room-Temperature NO2 Gas Sensor Based on ZnO Nanorod-Assembled Macro-/Mesoporous Film. ACS Appl. Electron. Mater. 2020, 2, 580–589. [Google Scholar] [CrossRef]
  197. Guo, F.; Feng, C.; Zhang, Z.; Zhang, L.; Xu, C.; Zhang, C.; Lin, S.; Wu, H.; Zhang, B.; Tabusi, A.; et al. A Room-Temperature NO2 Sensor Based on Ti3C2TX MXene Modified with Sphere-like CuO. Sens. Actuators B Chem. 2023, 375, 132885. [Google Scholar] [CrossRef]
  198. Gasso, S.; Kaur Sohal, M.; Chand Singh, R.; Mahajan, A. NOx Sensor Based on Semiconductor Metal Oxide and MXene Nanostructures. Mater. Today Proc. 2023, 92, 632–635. [Google Scholar] [CrossRef]
  199. Sun, B.; Qin, F.; Jiang, L.; Gao, J.; Liu, Z.; Wang, J.; Zhang, Y.; Fan, J.; Kan, K.; Shi, K. Room-Temperature Gas Sensors Based on Three-Dimensional Co3O4/Al2O3@Ti3C2Tx MXene Nanocomposite for Highly Sensitive NOx Detection. Sens. Actuators B Chem. 2022, 368, 132206. [Google Scholar] [CrossRef]
  200. Kuang, D.; Wang, L.; Guo, X.; She, Y.; Du, B.; Liang, C.; Qu, W.; Sun, X.; Wu, Z.; Hu, W.; et al. Facile Hydrothermal Synthesis of Ti3C2Tx-TiO2 Nanocomposites for Gaseous Volatile Organic Compounds Detection at Room Temperature. J. Hazard. Mater. 2021, 416, 126171. [Google Scholar] [CrossRef]
  201. Zhou, Y.; Wang, Y.; Wang, Y.; Li, X. Humidity-Enabled Ionic Conductive Trace Carbon Dioxide Sensing of Nitrogen-Doped Ti3C2Tx MXene/Polyethyleneimine Composite Films Decorated with Reduced Graphene Oxide Nanosheets. Anal. Chem. 2020, 92, 16033–16042. [Google Scholar] [CrossRef] [PubMed]
  202. Wang, J.; Xu, R.; Xia, Y.; Komarneni, S. Ti2CTx MXene: A Novel p-Type Sensing Material for Visible Light-Enhanced Room Temperature Methane Detection. Ceram. Int. 2021, 47, 34437–34442. [Google Scholar] [CrossRef]
  203. Persson, P.; Rosen, J. Current State of the Art on Tailoring the MXene Composition, Structure, and Surface Chemistry. Curr. Opin. Solid. State Mater. Sci. 2019, 23, 100774. [Google Scholar] [CrossRef]
  204. Rhouati, A.; Berkani, M.; Vasseghian, Y.; Golzadeh, N. MXene-Based Electrochemical Sensors for Detection of Environmental Pollutants: A Comprehensive Review. Chemosphere 2022, 291, 132921. [Google Scholar] [CrossRef]
  205. Park, Y.J.; Peñas-Defrutos, M.N.; Drummond, M.J.; Gordon, Z.; Kelly, O.R.; Garvey, I.J.; Gullett, K.L.; García-Melchor, M.; Fout, A.R. Secondary Coordination Sphere Influences the Formation of Fe(III)-O or Fe(III)-OH in Nitrite Reduction: A Synthetic and Computational Study. Inorg. Chem. 2022, 61, 8182–8192. [Google Scholar] [CrossRef]
  206. Hao, D.; Liu, Y.; Gao, S.; Arandiyan, H.; Bai, X.; Kong, Q.; Wei, W.; Shen, P.K.; Ni, B.-J. Emerging Artificial Nitrogen Cycle Processes through Novel Electrochemical and Photochemical Synthesis. Mater. Today 2021, 46, 212–233. [Google Scholar] [CrossRef]
  207. Hou, C.-Y.; Fu, L.-M.; Ju, W.-J.; Wu, P.-Y. Microfluidic Colorimetric System for Nitrite Detection in Foods. Chem. Eng. J. 2020, 398, 125573. [Google Scholar] [CrossRef]
  208. Guadagnini, L.; Tonelli, D. Carbon Electrodes Unmodified and Decorated with Silver Nanoparticles for the Determination of Nitrite, Nitrate and Iodate. Sens. Actuators B Chem. 2013, 188, 806–814. [Google Scholar] [CrossRef]
  209. Wang, T.; Wang, C.; Xu, X.; Li, Z.; Li, D. One-Step Electrodeposition Synthesized Aunps/Mxene/ERGO for Selectivity Nitrite Sensing. Nanomaterials 2021, 11, 1892. [Google Scholar] [CrossRef]
  210. Wang, Y.; Zeng, Z.; Qiao, J.; Dong, S.; Liang, Q.; Shao, S. Ultrasensitive Determination of Nitrite Based on Electrochemical Platform of AuNPs Deposited on PDDA-Modified MXene Nanosheets. Talanta 2021, 221, 121605. [Google Scholar] [CrossRef]
  211. Feng, X.; Han, G.; Cai, J.; Wang, X. Au@Carbon Quantum Dots-MXene Nanocomposite as an Electrochemical Sensor for Sensitive Detection of Nitrite. J. Colloid. Interface Sci. 2022, 607, 1313–1322. [Google Scholar] [CrossRef]
  212. Zhuang, J.; Pan, H.; Feng, W. 3D Urchin–like CoVO/MXene Nanosheet Composites for Enhanced Detection Signal of Nitrite. Sens. Actuators B Chem. 2023, 378, 133207. [Google Scholar] [CrossRef]
  213. Schmidt, E.W. Hydrazine and Its Derivatives: Preparation, Properties, Applications, 2 Volume Set; John Wiley & Sons: Hoboken, NJ, USA, 2001; ISBN 0471415537. [Google Scholar]
  214. Mackinson, F.W. Occupational Health Guidelines for Chemical Hazards. NIOSH OSHA 1981, 4, 81–123. [Google Scholar]
  215. Soundiraraju, B.; Raghavan, R.; George, B.K. Chromium Carbide Nanosheets Prepared by Selective Etching of Aluminum from Cr2AlC for Hydrazine Detection. ACS Appl. Nano Mater. 2020, 3, 11007–11016. [Google Scholar] [CrossRef]
  216. Soundiraraju, B.; Anthony, A.M.; Pandurangan, P.; George, B.K. Electrochemical Behavior of Chromium Carbide MXene Modified Electrodes: Hydrazine Sensing. Mater. Today Commun. 2022, 32, 103982. [Google Scholar] [CrossRef]
  217. Yao, Y.; Han, X.; Yang, X.; Zhao, J.; Chai, C. Detection of Hydrazine at MXene/ZIF-8 Nanocomposite Modified Electrode†. Chin. J. Chem. 2021, 39, 330–336. [Google Scholar] [CrossRef]
  218. Gul, S.; Serna, M.I.; Zahra, S.A.; Arif, N.; Iqbal, M.; Akinwande, D.; Rizwan, S. Un-Doped and Er-Adsorbed Layered Nb2C MXene for Efficient Hydrazine Sensing Application. Surf. Interfaces 2021, 24, 101074. [Google Scholar] [CrossRef]
  219. Yang, H.; Li, S.; Yu, H.; Zheng, F.; Lin, L.; Chen, J.; Li, Y.; Lin, Y. In Situ Construction of Hollow Carbon Spheres with N, Co, and Fe Co-Doping as Electrochemical Sensors for Simultaneous Determination of Dihydroxybenzene Isomers. Nanoscale 2019, 11, 8950–8958. [Google Scholar] [CrossRef]
  220. Shen, Y.; Rao, D.; Sheng, Q.; Zheng, J. Simultaneous Voltammetric Determination of Hydroquinone and Catechol by Using a Glassy Carbon Electrode Modified with Carboxy-Functionalized Carbon Nanotubes in a Chitosan Matrix and Decorated with Gold Nanoparticles. Microchim. Acta 2017, 184, 3591–3601. [Google Scholar] [CrossRef]
  221. Wu, L.; Lu, X.; Dhanjai; Wu, Z.-S.; Dong, Y.; Wang, X.; Zheng, S.; Chen, J. 2D Transition Metal Carbide MXene as a Robust Biosensing Platform for Enzyme Immobilization and Ultrasensitive Detection of Phenol. Biosens. Bioelectron. 2018, 107, 69–75. [Google Scholar] [CrossRef]
  222. Lei, L.; Li, C.; Huang, W.; Wu, K. Simultaneous Detection of 4-Chlorophenol and 4-Nitrophenol Using a Ti3C2Tx MXene Based Electrochemical Sensor. Analyst 2021, 146, 7593–7600. [Google Scholar] [CrossRef]
  223. Krishnamoorthy, R.; Muthumalai, K.; Nagaraja, T.; Rajendrakumar, R.T.; Das, S.R. Chemically Exfoliated Titanium Carbide MXene for Highly Sensitive Electrochemical Sensors for Detection of 4-Nitrophenols in Drinking Water. ACS Omega 2022, 7, 42644–42654. [Google Scholar] [CrossRef] [PubMed]
  224. Huang, R.; Liao, D.; Liu, Z.; Yu, J.; Jiang, X. Electrostatically Assembling 2D Hierarchical Nb2CTx and Zifs-Derivatives into Zn-Co-NC Nanocage for the Electrochemical Detection of 4-Nitrophenol. Sens. Actuators B Chem. 2021, 338, 129828. [Google Scholar] [CrossRef]
  225. Rasheed, P.A.; Pandey, R.P.; Jabbar, K.A.; Mahmoud, K.A. Platinum Nanoparticles/Ti3C2Tx (MXene) Composite for the Effectual Electrochemical Sensing of Bisphenol A in Aqueous Media. J. Electroanal. Chem. 2021, 880, 114934. [Google Scholar] [CrossRef]
  226. Huang, R.; Liao, D.; Chen, S.; Yu, J.; Jiang, X. A Strategy for Effective Electrochemical Detection of Hydroquinone and Catechol: Decoration of Alkalization-Intercalated Ti3C2 with MOF-Derived N-Doped Porous Carbon. Sens. Actuators B Chem. 2020, 320, 128386. [Google Scholar] [CrossRef]
  227. Tu, X.; Gao, F.; Ma, X.; Zou, J.; Yu, Y.; Li, M.; Qu, F.; Huang, X.; Lu, L. Mxene/Carbon Nanohorn/β-Cyclodextrin-Metal-Organic Frameworks as High-Performance Electrochemical Sensing Platform for Sensitive Detection of Carbendazim Pesticide. J. Hazard. Mater. 2020, 396, 122776. [Google Scholar] [CrossRef] [PubMed]
  228. Xie, Y.; Gao, F.; Tu, X.; Ma, X.; Xu, Q.; Dai, R.; Huang, X.; Yu, Y.; Lu, L. Facile Synthesis of MXene/Electrochemically Reduced Graphene Oxide Composites and Their Application for Electrochemical Sensing of Carbendazim. J. Electrochem. Soc. 2019, 166, B1673. [Google Scholar] [CrossRef]
  229. He, Y.; Zhou, X.; Zhou, L.; Zhang, X.; Ma, L.; Jiang, Y.; Gao, J. Self-Reducing Prussian Blue on Ti3C2Tx MXene Nanosheets as a Dual-Functional Nanohybrid for Hydrogen Peroxide and Pesticide Sensing. Ind. Eng. Chem. Res. 2020, 59, 15556–15564. [Google Scholar] [CrossRef]
  230. Zhao, F.; Yao, Y.; Jiang, C.; Shao, Y.; Barceló, D.; Ying, Y.; Ping, J. Self-Reduction Bimetallic Nanoparticles on Ultrathin MXene Nanosheets as Functional Platform for Pesticide Sensing. J. Hazard. Mater. 2020, 384, 121358. [Google Scholar] [CrossRef]
  231. Zhou, L.; Zhang, X.; Ma, L.; Gao, J.; Jiang, Y. Acetylcholinesterase/Chitosan-Transition Metal Carbides Nanocomposites-Based Biosensor for the Organophosphate Pesticides Detection. Biochem. Eng. J. 2017, 128, 243–249. [Google Scholar] [CrossRef]
  232. Lin, C.; Song, X.; Ye, W.; Liu, T.; Rong, M.; Niu, L. Recent Progress in Optical Sensors Based on MXenes Quantum Dots and MXenes Nanosheets. J. Anal. Test. 2023, 8, 95–113. [Google Scholar] [CrossRef]
  233. Wu, L.; Yuan, X.; Tang, Y.; Wageh, S.; Al-Hartomy, O.A.; Al-Sehemi, A.G.; Yang, J.; Xiang, Y.; Zhang, H.; Qin, Y. MXene Sensors Based on Optical and Electrical Sensing Signals: From Biological, Chemical, and Physical Sensing to Emerging Intelligent and Bionic Devices. PhotoniX 2023, 4, 15. [Google Scholar] [CrossRef]
  234. Bhardwaj, S.K.; Singh, H.; Khatri, M.; Kim, K.-H.; Bhardwaj, N. Advances in MXenes-Based Optical Biosensors: A Review. Biosens. Bioelectron. 2022, 202, 113995. [Google Scholar] [CrossRef]
  235. Chaudhuri, K.; Alhabeb, M.; Wang, Z.; Shalaev, V.M.; Gogotsi, Y.; Boltasseva, A. Highly Broadband Absorber Using Plasmonic Titanium Carbide (MXene). ACS Photonics 2018, 5, 1115–1122. [Google Scholar] [CrossRef]
  236. Li, R.; Zhang, L.; Shi, L.; Wang, P. MXene Ti3C2: An Effective 2D Light-to-Heat Conversion Material. ACS Nano 2017, 11, 3752–3759. [Google Scholar] [CrossRef]
  237. Cui, Y.; Yang, K.; Zhang, F.; Lyu, Y.; Zhang, Q.; Zhang, B. Ultra-Light MXene/CNTs/PI Aerogel with Neat Arrangement for Electromagnetic Wave Absorption and Photothermal Conversion. Compos. Part A Appl. Sci. Manuf. 2022, 158, 106986. [Google Scholar] [CrossRef]
  238. Li, W.; Miao, Y.; Zheng, Y.; Zhang, K.; Yao, J. Nb2CTx MXene Integrated Tapered Microfiber Based on Light-Controlled Light for Ultra-Sensitive and Wide-Range Hemoglobin Detection. IEEE Sens. J. 2022, 22, 11456–11462. [Google Scholar] [CrossRef]
  239. Kumar, A.; Verma, P.; Jindal, P. Surface Plasmon Resonance Biosensor Based on a D-Shaped Photonic Crystal Fiber Using Ti3C2Tx MXene Material. Opt. Mater. 2022, 128, 112397. [Google Scholar] [CrossRef]
  240. Ma, C.; Cao, Y.; Gou, X.; Zhu, J.-J. Recent Progress in Electrochemiluminescence Sensing and Imaging. Anal. Chem. 2020, 92, 431–454. [Google Scholar] [CrossRef]
  241. Sojic, N. Analytical Electrogenerated Chemiluminescence: From Fundamentals to Bioassays; Royal Society of Chemistry: London, UK, 2019; Volume 15, ISBN 1788014146. [Google Scholar]
  242. Ramanaviciene, A.; German, N.; Kausaite-Minkstimiene, A.; Voronovic, J.; Kirlyte, J.; Ramanavicius, A. Comparative Study of Surface Plasmon Resonance, Electrochemical and Electroassisted Chemiluminescence Methods Based Immunosensor for the Determination of Antibodies against Human Growth Hormone. Biosens. Bioelectron. 2012, 36, 48–55. [Google Scholar] [CrossRef]
  243. Ma, X.; Gao, W.; Du, F.; Yuan, F.; Yu, J.; Guan, Y.; Sojic, N.; Xu, G. Rational Design of Electrochemiluminescent Devices. ACC Chem. Res. 2021, 54, 2936–2945. [Google Scholar] [CrossRef]
  244. Kerr, E.; Farr, R.; Doeven, E.H.; Nai, Y.H.; Alexander, R.; Guijt, R.M.; Prieto-Simon, B.; Francis, P.S.; Dearnley, M.; Hayne, D.J.; et al. Amplification-Free Electrochemiluminescence Molecular Beacon-Based MicroRNA Sensing Using a Mobile Phone for Detection. Sens. Actuators B Chem. 2021, 330, 129261. [Google Scholar] [CrossRef]
  245. Daley, G.Q.; Van Etten, R.A.; Baltimore, D. Induction of Chronic Myelogenous Leukemia in Mice by the P210bcr/Abl Gene of the Philadelphia Chromosome. Science 1990, 247, 824–830. [Google Scholar] [CrossRef]
  246. Cheng, W.; Lin, Z.; Zhao, L.; Fan, N.; Bai, H.; Cheng, W.; Zhao, M.; Ding, S. CeO2/MXene Heterojunction-Based Ultrasensitive Electrochemiluminescence Biosensing for BCR-ABL Fusion Gene Detection Combined with Dual-Toehold Strand Displacement Reaction for Signal Amplification. Biosens. Bioelectron. 2022, 210, 114287. [Google Scholar] [CrossRef]
  247. Zhou, H.; Liu, J.; Xu, J.-J.; Zhang, S.; Chen, H.-Y. Chapter Two—Advances in DNA/RNA Detection Using Nanotechnology. In Advances in Clinical Chemistry; Makowski, G.S., Ed.; Elsevier: Amsterdam, The Netherlands, 2019; Volume 91, pp. 31–98. ISBN 0065-2423. [Google Scholar]
  248. Viter, R.; Kunene, K.; Genys, P.; Jevdokimovs, D.; Erts, D.; Sutka, A.; Bisetty, K.; Viksna, A.; Ramanaviciene, A.; Ramanavicius, A. Photoelectrochemical Bisphenol S Sensor Based on ZnO-Nanoroads Modified by Molecularly Imprinted Polypyrrole. Macromol. Chem. Phys. 2020, 221, 1900232. [Google Scholar] [CrossRef]
  249. Zhao, W.-W.; Xu, J.-J.; Chen, H.-Y. Photoelectrochemical Bioanalysis: The State of the Art. Chem. Soc. Rev. 2015, 44, 729–741. [Google Scholar] [CrossRef]
  250. Zhou, Q.; Tang, D. Recent Advances in Photoelectrochemical Biosensors for Analysis of Mycotoxins in Food. TrAC Trends Anal. Chem. 2020, 124, 115814. [Google Scholar] [CrossRef]
  251. Zhao, W.-W.; Xu, J.-J.; Chen, H.-Y. Photoelectrochemical DNA Biosensors. Chem. Rev. 2014, 114, 7421–7441. [Google Scholar] [CrossRef]
  252. Liu, F.; Zhao, J.; Liu, X.; Zhen, X.; Feng, Q.; Gu, Y.; Yang, G.; Qu, L.; Zhu, J.-J. PEC-SERS Dual-Mode Detection of Foodborne Pathogens Based on Binding-Induced DNA Walker and C3N4/MXene-Au NPs Accelerator. Anal. Chem. 2023, 95, 14297–14307. [Google Scholar] [CrossRef]
  253. Wang, L.; Zhu, F.; Chen, M.; Zhu, Y.; Xiao, J.; Yang, H.; Chen, X. Rapid and Visual Detection of Aflatoxin B1 in Foodstuffs Using Aptamer/G-Quadruplex DNAzyme Probe with Low Background Noise. Food Chem. 2019, 271, 581–587. [Google Scholar] [CrossRef] [PubMed]
  254. Viter, R.; Savchuk, M.; Iatsunskyi, I.; Pietralik, Z.; Starodub, N.; Shpyrka, N.; Ramanaviciene, A.; Ramanavicius, A. Analytical, Thermodynamical and Kinetic Characteristics of Photoluminescence Immunosensor for the Determination of Ochratoxin A. Biosens. Bioelectron. 2018, 99, 237–243. [Google Scholar] [CrossRef]
  255. Lu, X.; Wang, C.; Qian, J.; Ren, C.; An, K.; Wang, K. Target-Driven Switch-on Fluorescence Aptasensor for Trace Aflatoxin B1 Determination Based on Highly Fluorescent Ternary CdZnTe Quantum Dots. Anal. Chim. Acta 2019, 1047, 163–171. [Google Scholar] [CrossRef] [PubMed]
  256. Wu, Z.; Sun, D.-W.; Pu, H.; Wei, Q. A Novel Fluorescence Biosensor Based on CRISPR/Cas12a Integrated MXenes for Detecting Aflatoxin B1. Talanta 2023, 252, 123773. [Google Scholar] [CrossRef]
  257. Bandodkar, A.J.; Gutruf, P.; Choi, J.; Lee, K.; Sekine, Y.; Reeder, J.T.; Jeang, W.J.; Aranyosi, A.J.; Lee, S.P.; Model, J.B.; et al. Battery-Free, Skin-Interfaced Microfluidic/Electronic Systems for Simultaneous Electrochemical, Colorimetric, and Volumetric Analysis of Sweat. Sci. Adv. 2023, 5, eaav3294. [Google Scholar] [CrossRef] [PubMed]
  258. Li, Y.; Kang, Z.; Kong, L.; Shi, H.; Zhang, Y.; Cui, M.; Yang, D.-P. MXene-Ti3C2/CuS Nanocomposites: Enhanced Peroxidase-like Activity and Sensitive Colorimetric Cholesterol Detection. Mater. Sci. Eng. C 2019, 104, 110000. [Google Scholar] [CrossRef]
  259. Zeng, Y.; Hu, R.; Wang, L.; Gu, D.; He, J.; Wu, S.-Y.; Ho, H.-P.; Li, X.; Qu, J.; Gao, B.Z.; et al. Recent Advances in Surface Plasmon Resonance Imaging: Detection Speed, Sensitivity, and Portability. Nanophotonics 2017, 6, 1017–1030. [Google Scholar] [CrossRef]
  260. Balciunas, D.; Plausinaitis, D.; Ratautaite, V.; Ramanaviciene, A.; Ramanavicius, A. Towards Electrochemical Surface Plasmon Resonance Sensor Based on the Molecularly Imprinted Polypyrrole for Glyphosate Sensing. Talanta 2022, 241, 123252. [Google Scholar] [CrossRef]
  261. Kausaite-Minkstimiene, A.; Ramanavicius, A.; Ruksnaite, J.; Ramanaviciene, A. A Surface Plasmon Resonance Immunosensor for Human Growth Hormone Based on Fragmented Antibodies. Anal. Methods 2013, 5, 4757–4763. [Google Scholar] [CrossRef]
  262. Zhu, X.; Gao, T. Chapter 10—Spectrometry. In Nano-Inspired Biosensors for Protein Assay with Clinical Applications; Li, G., Ed.; Elsevier: Amsterdam, The Netherlands, 2019; pp. 237–264. ISBN 978-0-12-815053-5. [Google Scholar]
  263. Chen, R.; Kan, L.; Duan, F.; He, L.; Wang, M.; Cui, J.; Zhang, Z.; Zhang, Z. Surface Plasmon Resonance Aptasensor Based on Niobium Carbide MXene Quantum Dots for Nucleocapsid of SARS-CoV-2 Detection. Microchim. Acta 2021, 188, 316. [Google Scholar] [CrossRef]
  264. Baniukevic, J.; Hakki Boyaci, I.; Goktug Bozkurt, A.; Tamer, U.; Ramanavicius, A.; Ramanaviciene, A. Magnetic Gold Nanoparticles in SERS-Based Sandwich Immunoassay for Antigen Detection by Well Oriented Antibodies. Biosens. Bioelectron. 2013, 43, 281–288. [Google Scholar] [CrossRef]
  265. Liu, R.; Jiang, L.; Yu, Z.; Jing, X.; Liang, X.; Wang, D.; Yang, B.; Lu, C.; Zhou, W.; Jin, S. MXene (Ti3C2Tx)-Ag Nanocomplex as Efficient and Quantitative SERS Biosensor Platform by in-Situ PDDA Electrostatic Self-Assembly Synthesis Strategy. Sens. Actuators B Chem. 2021, 333, 129581. [Google Scholar] [CrossRef]
Figure 1. Elements of the known MXenes (A) and MAX phases (B).
Figure 1. Elements of the known MXenes (A) and MAX phases (B).
Nanomaterials 14 00447 g001
Figure 2. Schematic depiction of Nb2CTz MXene synthesis. Reprinted from [39].
Figure 2. Schematic depiction of Nb2CTz MXene synthesis. Reprinted from [39].
Nanomaterials 14 00447 g002
Figure 3. Schematic illustration of the crack propagation mechanism behind MXene-based strain sensors. Reprinted from [54].
Figure 3. Schematic illustration of the crack propagation mechanism behind MXene-based strain sensors. Reprinted from [54].
Nanomaterials 14 00447 g003
Figure 4. Schematic illustration of the mechanism behind MXene-based hydrogel strain sensors.
Figure 4. Schematic illustration of the mechanism behind MXene-based hydrogel strain sensors.
Nanomaterials 14 00447 g004
Figure 5. Schematic illustration of the mechanism behind MXene-based pressure sensors.
Figure 5. Schematic illustration of the mechanism behind MXene-based pressure sensors.
Nanomaterials 14 00447 g005
Figure 6. Crack mechanism of temperature sensors with MXenes.
Figure 6. Crack mechanism of temperature sensors with MXenes.
Nanomaterials 14 00447 g006
Figure 7. Experimental setup for temperature sensing. Reprinted from [106].
Figure 7. Experimental setup for temperature sensing. Reprinted from [106].
Nanomaterials 14 00447 g007
Figure 8. Gas sensing mechanism of the pristine 2D MXenes and the corresponding energy band diagram. Reprinted from [126].
Figure 8. Gas sensing mechanism of the pristine 2D MXenes and the corresponding energy band diagram. Reprinted from [126].
Nanomaterials 14 00447 g008
Figure 9. Nitrite detection scheme of AuNPs/MXene/ERGO-based sensor. Reprinted from [209].
Figure 9. Nitrite detection scheme of AuNPs/MXene/ERGO-based sensor. Reprinted from [209].
Nanomaterials 14 00447 g009
Table 1. MXene application in strain sensors.
Table 1. MXene application in strain sensors.
CompositionSensitivity (GF)Test Range, Strain %LOD, Strain %Ref
Ti3C2Tx/MWCNTs/TPU13–3630–100-[58]
Ti3C2Tx MXene/polyurethane22.9–22858–1500.1[59]
CNT/Ti3C2Tx MXene/PDMS13.30–60.3-[53]
MXene/PANIF97.6–2369.10–800.1538[60]
Ti3C2Tx MXene/paper17.40–0.60.1[61]
Ti3C2Tx MXene/P(VDF-TrFE)6.35–108.845–66-[62]
PVA/CMC/TA/MXene hydrogel2.90–7001[63]
PVA-CA-MXene hydrogel2.30–200-[64]
Polyvinyl alcohol/polyacrylamide/CaCl2/MXene3.00–300-[65]
Liquid metal/MXene7.85–15.470–400-[66]
PAA/PEDOT:PSS/MXene hydrogel9–20.860–1000-[67]
MXene/AgNWs/TPU33.10–120-[68]
Table 2. MXene application in pressure sensors.
Table 2. MXene application in pressure sensors.
CompositionSensitivity, kPa−1 Test Range, kPaLOD, PaRef
MXene/CTAB/CMF381.91 0−30.1[95]
MXene/PPNs/MXene/TPUEM34.9–177.30–30.4-[96]
MXene/Polyetherimide20 (150 °C)–80 (−5 °C)0–409[97]
MXene/rGO/PS115–224 0–20.65-[98]
MXene/bacterial cellulose (BC)0.94–95.20–100.4[99]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Navitski, I.; Ramanaviciute, A.; Ramanavicius, S.; Pogorielov, M.; Ramanavicius, A. MXene-Based Chemo-Sensors and Other Sensing Devices. Nanomaterials 2024, 14, 447. https://doi.org/10.3390/nano14050447

AMA Style

Navitski I, Ramanaviciute A, Ramanavicius S, Pogorielov M, Ramanavicius A. MXene-Based Chemo-Sensors and Other Sensing Devices. Nanomaterials. 2024; 14(5):447. https://doi.org/10.3390/nano14050447

Chicago/Turabian Style

Navitski, Ilya, Agne Ramanaviciute, Simonas Ramanavicius, Maksym Pogorielov, and Arunas Ramanavicius. 2024. "MXene-Based Chemo-Sensors and Other Sensing Devices" Nanomaterials 14, no. 5: 447. https://doi.org/10.3390/nano14050447

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

Article Metrics

Back to TopTop