Next Article in Journal
The Global Research Trend in Electrochemical Microfluidic Technology: A Bibliometric Review
Next Article in Special Issue
Hydrogen Sensors Based on In2O3 Thin Films with Bimetallic Pt/Pd Catalysts on the Surface and Tin and Dysprosium Impurities in the Bulk
Previous Article in Journal
Hierarchical Modelling of Raman Spectroscopic Data Demonstrates the Potential for Manufacturer and Caliber Differentiation of Smokeless Powders
Previous Article in Special Issue
Tetrafluorosubstituted Metal Phthalocyanines: Study of the Effect of the Position of Fluorine Substituents on the Chemiresistive Sensor Response to Ammonia
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Ti2CTx MXene Oxidation on Its Gas-Sensitive Properties

by
Artem S. Mokrushin
1,*,
Ilya A. Nagornov
1,
Philipp Yu. Gorobtsov
1,
Aleksey A. Averin
2,
Tatiana L. Simonenko
1,
Nikolay P. Simonenko
1,
Elizaveta P. Simonenko
1 and
Nikolay T. Kuznetsov
1
1
Kurnakov Institute of General and Inorganic Chemistry, Russian Academy of Sciences, 31 Leninsky pr., 119991 Moscow, Russia
2
Frumkin Institute of Physical Chemistry and Electrochemistry, Russian Academy of Sciences, 31 Leninsky pr., bldg. 4, 199071 Moscow, Russia
*
Author to whom correspondence should be addressed.
Chemosensors 2023, 11(1), 13; https://doi.org/10.3390/chemosensors11010013
Submission received: 30 November 2022 / Revised: 19 December 2022 / Accepted: 20 December 2022 / Published: 22 December 2022
(This article belongs to the Special Issue Gas Sensors for Monitoring Environmental Changes)

Abstract

:
The oxidation process was studied for the synthesized low-layer Ti2CTx MXene deposited on a special Al2O3/Pt sensor substrate using in situ Raman spectroscopy. It is noted that on the ceramic parts of the substrate (Al2O3), the beginning of oxidation (appearance of anatase mod phase) is observed already at 316 °C, in comparison with platinum, for which the appearance of anatase is noted only at 372 °C. At the temperature 447 °C, the initial MXene film is completely oxidized to TiO2. Using scanning electron microscopy and atomic force microscopy, the microstructure and dispersity of the obtained MXene film were studied. It was found that the obtained films exhibit chemoresistive responses to the detection of a wide group of gases, H2, CO, NH3, C6H6, C3H6O, CH4, C2H5OH and O2, at room temperature and RH = 50%. The highest sensitivity is observed for NH3. The partial oxidation of the Ti2CTx MXene was shown to favorably affect the gas-sensitive properties.

1. Introduction

Two-dimensional titanium-containing carbide MXenes with the general formula Tin+1CnTx have attracted increased attention from the scientific community in recent years due to their unique properties—a large surface-to-volume ratio, high electrical conductivity, which can be metallic or semiconducting, and surface functionalization by various groups (primarily −F, −OH, −O), whose composition can be adjusted in some limits by choice of synthesis method. [1,2,3]. This combination of unique MXene properties opens up the potential for their use in various fields: in lithium/sodium-ion batteries and supercapacitors [4,5], fuel cells [6,7], photocalysis [8,9], as well as chemical gas sensing [10,11,12,13,14] etc.
In contrast to metal oxide semiconductors (MOS), the classical receptor materials for chemoresistive gas sensors, Tin+1CnTx MXenes often have metallic conductivity, their outer surface is completely covered by functional groups, and their morphology refers to 2D-nanomaterials. This combination makes this class of compounds attractive for chemoresistive gas sensors with a high signal-to-noise ratio (SNR) [15], including room temperature operation.
The detection of the MXenes mechanism differs significantly from MOS materials. A universal model describing the interaction of the analyte gas with ion-adsorbed oxygen and the formation of an electron-depleted layer (EDL) on the surface layer of the sensitive material can be used for MOS-gas sensors [16]. At present, the question about the mechanism of gas detection by MXenes is debatable. Nevertheless, theoretical and experimental models allowing the description of the appearance of the chemoresistive signal in the atmosphere of different gases have already been observed [17,18,19].
MXenes, due to their predisposition to form hydrogen bonds with functional groups, are extremely sensitive to moisture. The harmonic structure characteristic of MXenes multilayer, the metallic type of conductivity, and the fact that water molecules do not form covalent bonds with the surface of MXenes make them extremely sensitive to humidity and allow a reproducible response. In Zhang C. et al. [20], using FT-IR spectroscopy and the study of the wetting angle on the MXenes surface, the authors found that water molecules from the gas phase can be chemically sorbed on the surface and in the interlayer space of Ti3C2Tx MXene. The authors proposed a mechanism whereby the hydroxyl groups thus formed on the surface provide an electrostatic field and prevent charge transfer, which is explained by an increase in electrical resistance (positive response: p-type) with increasing humidity. The nanoscale interlayer space (~1.01 nm) of Ti3C2Tx, as well as the surface super hydrophilicity found by the authors, prevents capillary condensation of water and limits the movement of hydroxyl groups, which are characteristic of bulk nanomaterials.
Structural water plays an important role in reducing the overall interaction between the MXenes layers and creating additional space for the adsorption and diffusion of gases with affinity to functional groups. This feature is an advantage over other 2D-layered materials, such as graphene, in which strong Van-der-Waals interactions between two-dimensional sheets and small interlayer distances prevent the intercalation of gas molecules. Koh H-J. et al., in their article [21], in situ using XRD demonstrated that the interlayer space of MXenes contains water fragments (hydroxyl groups or water molecules themselves), which can be removed when the material is exposed to dry nitrogen. The authors showed that ethanol molecules can chemically bind with Ti3C2Tx MXene in its interlayer space. As a result of these interactions, the MXene swells, and its interlayer distance increases due to the steric effect. The swollen two-dimensional MXene decreases the number of electrons involved in charge transfer between the layers during ethanol detection, which leads to an increase in electrical resistance and allows the positive (p-type) chemoresistance response to be recorded.
The MXenes show the best sensitivity when detecting gases able to form hydrogen bonds. Majhi S.M. et al., in article [18], showed a positive (p-type) response when acetone is detected by the Ti3C2Tx MXene. Kim S.J. et al. [15] also showed that when used as a receptor material, Ti3C2Tx MXene demonstrates the greatest response to gases capable of forming a hydrogen bond (acetone, ethanol, propanal and ammonia) than observed for acid gases (NO2, SO2 and CO2). The authors demonstrated the ability to detect 50–1000 ppb of acetone, ethanol and ammonia with a low SNR value. DFT calculations found that hydroxyl groups (−OH) made the largest contribution to acetone bonding to functional groups on the Ti3C2Tx MXene surface compared to oxygen (−O) and fluoride (−F) groups. It was found that for OH-groups, the bonding energy value is significantly more negative compared to −O and −F: −0.774 compared to −0.317 and −0.311 eV, respectively. The authors concluded that the terminal hydroxyl groups on the surface of MXenes play a key role in the detection of various gases to form hydrogen bonds [15].
In most of the papers presented in the literature, a positive chemoresistive response of the p-type is recorded when detecting various gases with carbide titanium-containing MXenes. Nevertheless, there are also studies in which the electrical resistance decreases when gases are injected (negative response: n-type). For example, Wu M. et al. [22] observed a decrease in electrical resistance when detecting ammonia by the Ti3C2Tx MXenes. The authors explain this by the strong functionalization of the surface by chloride groups (−Cl), which were formed on the surface of MXenes due to the peculiarities of the synthesis (etching MAX-phase in the Lewis acid ZnCl2 melt, which is different from the classical liquid-phase methods). This phenomenon is not fully explained in this paper. The authors suggest that chloride groups can affect the MXene band structure, which leads to the reverse chemoresistive response.
Nanocomposites based on them, primarily based on titanium dioxide Tin+1CnTx/TiO2, are no less interesting receptor materials for gas sensors as compared to individual Tin+1CnTx MXenes. Such nanocomposites can be obtained, in particular, by thermal oxidation in an oxygen-containing atmosphere of the initial Tin+1CnTx MXenes. Such Tin+1CnTx/TiO2 nanocomposites can have improved gas-sensitive characteristics. In [23], the authors oxidized a Ti3C2Tx film deposited on a special multi-sensor chip in air at 350°C. The resulting nanocomposite demonstrated an increased semiconductor response to ethanol at relatively high temperatures (300–350 °C), and a decrease in the response and recovery time of the sensor was noted. In [24], the authors used oxygen plasma to oxidize MXene. The resulting nanocomposite showed an increased response to ethanol by several orders of magnitude. The formation of Tin+1CnTx/TiO2 nanocomposites is also possible by holding MXene in an aqueous solution, specifically under the influence of light. In [25], the authors obtained Tin+1CnTx/TiO2 by holding the individual components in an ethanol-water mixture. For the synthesized nanocomposite, they were able to increase the sensitivity, as well as lower the threshold of sensitivity to ammonia to 100 ppb. In [26], the authors in situ-oxidized the initial Ti3C2Tx MXene using hydrothermal synthesis. The resulting oxidized nanocomposite demonstrated increased sensitivity and selectivity to NO2 detection. In [27], the authors used a “simple spray method” to spray an aqueous dispersion of TiO2 onto the surface of Ti3C2Tx MXene. The resulting nanocomposite showed improved gas-sensitive characteristics when detecting low NH3 concentrations at RT, as well as a more stable signal compared to the individual Ti3C2Tx MXene.
The present paper is devoted to the study of the oxidation process of the Ti3C2Tx MXene film using in situ Raman spectroscopy and to the study of the gas-sensitive chemoresistance properties of the obtained nanomaterials to a wide group of analyte gases.

2. Materials and Methods

2.1. Mxene Synthesis and Film Application

Reagents: powders of metallic titanium (99.9%, 0.5–100 µm, Ruskhim, Moscow, Russia), aluminum (99.2%, 30 µm, Ruskhim), graphite (MPG-8 grade, Ruskhim), potassium bromide KBr (99%, Ruskhim), sodium fluoride NaF (osc. Part 9-2, Reackhim, Moscow, Russia), hydrochloric acid HCl (>99%, Sigma Tech, Moscow, Russia), tetramethylammonium hydroxide solution (CH3)4N(OH) (TMAOH, 25%, aqueous solution, Technic, Saint-Denis, France).
The synthesis of Ti2CTx MXene was performed by selective etching of aluminum layers contained in the MAX-phase of Ti2AlC. For this purpose, Ti2AlC powder was added to a sodium fluoride solution in hydrochloric acid; the Ti2AlC and Ti2CTx synthesis techniques used in this study are described in detail in [28,29].
Briefly, to obtain Ti2AlC, powders of aluminum, titanium, graphite and potassium bromide were mixed in the ratio n(Ti):n(Al):n(C) = 2:1.2:0.8 and m(Ti+Al+C) = m(KBr), then co-milled, pressed into tablets and heat treated in a muffle furnace at 1000 °C in a protective melt KBr.
To obtain the multilayer Ti2CTx MXene, a 1 g sample of the MAX-phase Ti2AlC mass was slowly added to 20 mL of 6 M HCL solution with 1 g of NaF. The system was then incubated under stirring and at 40 ± 5 °C for 24 h. The resulting powder was separated by centrifugation and repeatedly washed with distilled water until pH ~5–6 was reached. Delamination was performed using tetramethylammonium hydroxide (CH3)4N(OH) solution under ultrasonic influence.
The obtained Ti2CTx MXene sample was held in an aqueous dispersion at 4–6 °C for 7 days and then used to apply the receptor layer.
The Ti2CTx MXene film was deposited by drop casting method on the surface of a specialized Al2O3 substrate with platinum interdigital electrodes (from the front side) and a platinum microheater (from the back side). For this purpose, the dispersion system (carbide nanosheets and deionized water in a volume of 50 µL) was applied to the substrate surface in the area of counter-pin electrodes using an automatic dispenser. Then, a step drying in the range of 25–150 °C under reduced pressure was performed to remove the solvent.

2.2. Instrumentation

The microstructure and chemical composition of the film surface were studied by scanning (NVision 40 scanning electron microscope, Carl Zeiss (Oberkochen, Germany, secondary electron detector, accelerating voltage 1–10 kV) and transmission electron microscope (JEOL JEM-1011, Akishima, Japan), as well as X-ray spectral elemental microanalysis (energy-dispersive X-ray (EDX) spectrometer INCA X-MAX 80, Oxford Instruments (Oxford, UK), accelerating voltage 20 kV). The powder phase composition and films were studied by means of X-ray diffraction (XRD) on D8 Advance device (Bruker, Billerica, MA, USA,), CuKα = 1.5418 Å, Ni filter, E = 40 keV, I = 40 mA; the range of 2θ: 5–45°; resolution: 0.02°; signal accumulation time in the point: 0.3 s).
The obtained Ti2CTx MXene films were examined by atomic force microscopy (AFM). Surface topography and Kelvin-probe force microscopy (KFM) were studied on a Solver Pro-M scanning probe microscope (NT-MDT production, Zelenograd, Moscow, Russia) using ETALON HA-HR probes with W2C-based conductive coating (tip curvature <35 nm). The microstructure of the obtained films was studied using SEM and AFM on the Al2O3 section between platinum microelectrodes.
Raman spectra were obtained with a Renishaw inVia Reflex Microscope system equipped with a Peltier-cooled CCD. The 532-nm lines of a Nd:YAG laser were used for excitation. The laser light was focused on the sample through a 50× objective to a spot size of ~2 μm. The power of the sample was <0.3 mW. The variable-temperature Raman scattering measurements were made using a THMS600 stage (Linkam Scientific Instruments Ltd., Redhill, UK). The in situ film heating rate was 5 °C/min. After conducting Raman in situ heating experiments on the samples, the temperature was recalibrated on a typical sample using a high-precision Testo 868 thermal imaging camera. Due to the fact that Raman spectroscopy can be destructive for some samples (especially those prone to oxidation, as in this case) due to local heating of the region from which the spectrum is taken after laser exposure, spectra were recorded in different areas of the film, which are nevertheless localized close to each other. This approach avoided the local overheating that could occur with repeated laser exposure, even at low power, which allowed more reliable data to be obtained. All spectra obtained were normalized with respect to the most intense spectrum.
The chemoresistive responses were obtained using a special laboratory setup, a detailed description of which can be found in our earlier papers [30,31]. The gas medium was created in a quartz cell using two Bronkhorst gas flow controllers with a maximum throughput of 100 and 200 mL/min. The quartz cell volume is 7 × 10−5 m3. The electrical resistance of the obtained oxide films was measured using a Fluke 8846A Digit Precision Multimeter, which has an upper detection limit of 1000 MΩ. The temperature of the sensor was monitored using a platinum microheater pre-calibrated with a high-precision Testo 868 thermal imaging camera. The measurements of gas-sensitive properties, among others, were carried out at room temperature. Before starting the gas-sensitive measurements, the film to be measured was incubated in a baseline gas atmosphere until a stable signal was established.
To measure the signal at different relative humidity (RH), we used a special unit with a bubbler flask; the RH of the gas mixture was controlled by a digital flow-through hygrometer “Excis”. The temperature value of the relative humidity was set and then measured at 20 °C.
All gas-sensing measurements were carried out at room temperature (RT) and 50% relative humidity. The response to H2, CO, NH3, benzene (C6H6), acetone (C3H6O), methane (CH4), ethanol (C2H5OH) and oxygen (O2) was calculated using the following ratio:
S 1 = R BL R g R BL × 100 %
where RBL is the baseline resistance (nitrogen (99.9999%) and was used as the baseline for oxygen detection and synthetic air for other gases, and Rg is the resistance at a given concentration of analyte gas.
The response to humidity was calculated using the following ratio:
S 2 = R BL R RH R BL × 100 %
where RBL is the resistance at 30% relative humidity, and RRH is a given relative humidity.

3. Results and Discussion

3.1. Investigation of the Phase Composition and Microstructure of the Obtained Ti2CTx Mxene

As shown in Figure 1a, the obtained MXene sample contains no significant admixtures of the original Ti2AlC, and the shift in the reflex position (002) toward lower angles to 2θ = 4.6° indicates the successful delamination of the multilayer aggregates formed by etching the MAX-phase with NaF-HCl. The shift of the reflex associated with the Ti2CTx phase for the coating on the specialized substrate toward higher angles to 2θ = 5.3° indicates that the reverse aggregation of the MXene sheets and the decrease in the interlayer distance during vacuum drying occurs during coating formation.
The TEM data indicate partial destruction of the MXene sheets (Figure 1b,c) while holding it in aqueous dispersion, as well as the appearance of loose nanoparticles on the sheet boundaries, probably related to oxidation products—titanium dioxide in different crystalline modification.

3.2. Raman Spectroscopy

In situ Raman spectra recording for Ti2CTx MXene films on an Al2O3 chip with Pt interdigital microelectrodes (Figure 2a) was performed both for platinum microelectrodes and for the Al2O3 area between them (Figure 2b). The results of which can be attributed to the inhomogeneity of the prepared film. It should be noted that modes on the Raman spectra characteristic of the MXene phase are more intensive on platinum.

3.2.1. MXene Raman Spectra for the Platinum Electrodes Area

A Ti2CTx MXene film is characterized by three normal modes, ω1, ω2 and ω3, in Raman spectra, located at 270, 380 and 697 cm−1, respectively (Figure 3a). Modes ω13 can be attributed both to vibrations of MAX-phase Ti2AlC, remaining after synthesis, and to Ti2CTx MXene itself [32,33,34,35]. Analysis of the spectrum obtained at RT also reveals other modes: ωR1 and ωR2 modes are located at 447 and 821 cm−1, which can be attributed to high convergence to Eg and B2g modes characteristic of rutile TiO2 [36,37]. For the rutile phase, the largest intensity should be observed for Eg and A1g modes with maxima at ~448 and 613 cm−1, respectively, while B2g should be the least intensive of all first-order modes characteristic of rutile [38]. In this case, however, the ωR2 (B2g) mode is the most intensive of all rutile phase modes, which is highly irregular for rutile TiO2 materials in the literature [38,39]. Such behavior allows us to suppose that the obtained MXene at RT does not include a separate rutile TiO2 phase, with observed ωR1 (Eg) and ωR2 (B2g) modes arising from Ti−O bonds, with energies close to analogous bonds in rutile, on the MXene surface. Moreover, the observed rutile species on the surface of Ti2CTx MXene likely formed as a result of keeping the synthesized sample in water suspension.
In the more long-wave region of the spectrum, ωD and ωG modes are located at 1349 and 1583 cm−1, which are characteristic carbon D- and G-bands [32], which are common for many carbon systems with sp2 hybridization of the C−C bond. Therefore, they can be attributed to carbon in MXene as well as to graphene layers resulting from the excessive etching of Ti2AlC by hydrofluoric acid and the partial removal of titanium atoms beside aluminum atoms [40,41,42].
In addition to the main MXene modes, low-intensity modes ωT1, ωT2 and ωT3 were found at 754, 935 and 1454 cm−1, respectively. Modes ωT1T3 are characteristic of tetramethylammonium hydroxide (TMAOH) [43], fragments of which had likely been captured in MXene interlayer space after delamination and washing. At 1042 cm−1, the ωN mode is also present, attributed to the nitro group (NO3) [44,45], which could have formed on the MXene surface after the gas-sensing properties studies in an atmosphere containing NO2. The possibility of gaseous NO2 transformation into an NO3− group on the surface of the Ti3C2Tx/TiO2 material has been considered in studies [46,47].
The described set of modes is retained by the sample under heating up to 109 °C. At higher temperatures, the spectra undergo significant changes. Starting with 147 °C, no ωT1T3 and ωN modes (characteristic of TMAOH and NO3 groups, respectively) appear in the spectra. This can be attributed to their desorption and/or decomposition with the following desorption of gaseous decomposition products from the MXene surface. Heating to ≥147 °C results in the overlap of ω2 mode with the rutile ωR1 band, yielding a new widened peak. Further heating leads to the transformation of this peak with maximum shifting to a lower wave number. At 410 °C, the maximum of this phase is located at 420 cm−1, which is close to the B1g mode of the titanium dioxide anatase phase [38]. This might indicate partial restructuring of the crystal lattice. Modes characteristic of Ti2CTx1 and ω3), as well as of TiO2 rutile (ωR2), still remain on the film spectra, with their maxima not undergoing any changes.
The most intensive Eg mode at 156 cm−1A1), characteristic of anatase, starts to appear on spectra at 372 °C. The intensity of this mode rises with the temperature, which indicates that MXene is being oxidized to titanium dioxide. The A1g mode at 608 cm−1R3), which is the most intensive mode of the rutile phase [38], also appears on the Raman spectra at temperatures ≥ 372 °C as a result of the oxidation process. The Raman spectrum at 447 °C significantly differs from those recorded for the MXene film at lower temperatures, having the following set of bands: ωA1 (Eg) and ωA2 (B1g) at 156 and 396 cm−1 from the anatase phase and ωR3 (A1g) and ωR2 (B2g) at 615 and 821 cm−1 from the rutile phase. The obtained values are in good agreement with the literature [37,38].
D- and G-modes are retained up to the temperature of 447 °C. An increase in temperature gives rise to the systematic shift of the G-mode to lower wave numbers. Thus, at 410 °C, the G-mode maximum is situated at 1570 cm−1. The shift of the maximum position can be related to temperature effects arising from in situ heating of the sample.
Thus, analysis of the recorded Raman spectra for MXene film on platinum allows us to infer that the oxidation of MXene begins at 372 °C. It is at this temperature that the Ti2CTx/TiO2 composite with a large content of rutile and anatase phases of TiO2 forms. At the temperature of 447 °C, the Ti2CTx/TiO2 film fully oxidizes TiO2, consisting of rutile and anatase.

3.2.2. Raman Spectra for the Al2O3 Area

Considering the features of sensing studies, it is the Raman spectra of MXene on a ceramic Al2O3 surface between platinum microelectrodes that are of the most interest since chemoresistive signals of the gas-sensitive film arise from this area of the sample. We have observed that Ti2CTx spectra recorded on local areas on the Al2O3 surface significantly differ from those recorded for MXene film areas situated on the surface of platinum electrodes. We were able to obtain intensive signals for Ti2CTx on the surface of the platinum electrodes related to vibrations of various phases. On the Al2O3 surface, however, most modes are of low intensity, and for the initial TI2CTx spectrum, large background luminescence was observed, which obstructs registering Raman scattering. For this reason, only the region up to 900 cm−1 is shown in the graph with characteristic MXene modes and oxidation products.
On all spectra of the MXene film (Figure 3b), the characteristic set of modes ωS1S5, typical for α-Al2O3 substrate, can be seen at 377, 417, 574, 644 and 749 cm−1 [48]. At RT, only weak ω1 mode of MXene at 276 cm−1 and intense mode ωR2 (B2g) at 828 cm−1 from rutile (these modes were the most intense for Raman spectra of MXene on the platinum surface—see Section 3.2.1) are observed. No other modes, characteristic of MXene, were found for the film on Al2O3‘s surface due to their low intensity and background luminescence. The described set of modes is retained up to 316 °C when the Eg (1) mode at 156 cm−1A1) from anatase, as well as Eg (3) at 634 cm−1A3), appear, with the latter mode not observed in spectra for the platinum surface. The obtained data are in good agreement with those in the literature [38]. It should be noted that the start of anatase formation owing to MXene oxidation already starts at 316 °C on the Al2O3 substrate, while for the platinum substrate, the same is only observed at 372 °C. Further increases in temperature to 410–447 °C result in full oxidation of MXene to TiO2. Modes characteristic of anatase can be observed in this case on the Raman spectra: strong Eg (1) at 156 cm−1A1) and weaker B1g(2) + A1g at 510 (ωA2) and Eg(3) at 634 cm−1A3), as well as B2g at 828 cm−1R2) from the rutile phase [38]. It should be noted that on Al2O3‘s surface, the oxidation of Ti2CTx MXene can be seen to proceed with the formation of mostly anatase phase. Only the weak B2g mode can be attributed to rutile, which was present from the start. In the case of the final Raman spectra for platinum substrate, ωR1 (Eg) and ωR3 (A1g) of the rutile phase can be observed. These seeming differences in oxidation behavior of the Ti2CTx MXene-receptive layer can be connected to the effect of the substrate material used for MXene film deposition on the recorded Raman spectra.
For further studies of the gas-sensing properties of the prepared Ti2CTx/TiO2 nanocomposite, we have chosen the film oxidized at the minimal temperature of 316 °C.

3.3. Microstructure

SEM micrographs for films on the special Al2O3/Pt substrate are given in Figure 4. The initial MXene film (Figure 4a,b) exhibits the folded wavelike microstructure: the surface is reminiscent of sea crests. The length of such crests can reach several microns. It can be seen from the micrograph that the film surface is rather developed, which is important for gas-sensing applications, which require a high specific surface area for better adsorption of gases from the atmosphere. Separate nanoparticles can also be seen on the micrographs, having a mostly spherical shape, with sizes of about 200 nm. Such particles might consist of impurities.
Heating up to 316 °C gives rise to significant microstructural changes in the MXene film (Figure 4c,d). It becomes more uniform, although the wavelike microstructure is retained. Spherical particles 1–3 μm in diameter start to appear, most likely being aggregates of TiO2 in rutile crystal modification. The enlargement of TiO2 particles can be attributed to the aggregation of smaller particles, as well as to overall phase transformation due to heat treatment of the films. Judging from Raman spectroscopy data given earlier, a nanocomposite is formed under such temperatures, consisting of the MXene phase and TiO2.
Further increases in heat treatment temperature to 447 °C result in even more pronounced changes in the microstructure (Figure 4e,f). The wavelike structure fully gives way to a smoother surface, consisting mostly of spherical aggregates 1–3 μm in diameter. Smaller spherical nanoparticles, 20–50 nm in size, can be seen from micrographs with higher magnification (Figure 4f). Raman spectroscopy analysis suggests that under such temperatures, the TiO2 film forms, consisting of a rutile and anatase phases mixture.
The microstructure of MXene films after their heat treatment at 316 and 447 °C was additionally studied using AFM. Topography scans are given in (Figure 5a,b), as well as the distribution of surface potential for the sample of MXene film oxidized at 316 °C. Topography results from AFM are in good agreement with SEM results (Figure 4c,d). Areas with spherical aggregates 2–4 μm in size, as well as narrow (width of ~100–250 nm) and elongated wavelike structures, roughly 100 nm in size, can be seen. Surface potential values for different film areas differ noticeably, with a work function changing varying from 4.57 to 5.00 eV, which is a consequence of film inhomogeneity. The mean work function value was 4.71 eV, which is rather close to the value for individual Ti2CTx MXene with a varyingly functionalized surface: 4.5–4.98 eV [49,50]. Thus, the obtained AFM results correlate well with Raman spectroscopy and SEM data and also confirm the formation of the Ti2CTx/TiO2 nanocomposite.
The topography scans are given in (Figure 5c,d), as well as the distribution of the surface potential for the sample of the MXene film oxidized at 447 °C. The obtained film surface is relatively smooth, consisting of low (50–80 nm) spherical aggregates 2–4 μm in size, with small spherical agglomerates 200–400 nm in size situated on their edges and in between them. Under higher magnification and in the phase contrast regime, nanoparticles with a shape close to spherical, size of ~45–95 nm, were found, which is in agreement with the SEM results. The mean square roughness on an area of 100 μm2 is just 20 nm. It can be seen from the map of the potential surface distribution obtained from KPFM (Figure 5d) that charge carriers are spread relatively uniformly on the film surface. The mean work function value for all scanned areas was 5.00 eV. This value is in good agreement with those reported in the literature for anatase [51]. For anatase, the work function lies in the range of 4.94–5.07 eV, while for rutile, it is ~4.80 eV. Thus, AFM also confirms the formation of TiO2 with mostly anatase crystal structure.

3.4. Gas-Sensing Chemoresistive Properties

The gas-sensitive chemoresistive properties at room temperature were studied for the initial Ti2CTx MXene film as well as the Ti2CTx/TiO2 film oxidized in situ in a Raman spectroscopy cell at 316 °C. A further increase in the operating temperature (including in an atmosphere of high humidity) led to a significant increase in electrical resistance (R > 1 GOhm), which made it impossible to measure the gas-sensitive properties.
In the first stage, the obtained Ti2CTx and Ti2CTx/TiO2 films had their baseline resistances measured in an atmosphere of dry air and dry nitrogen. In both cases, the resistance of the films exceeded 1 GOm, which did not allow their gas-sensitive properties to be measured in these atmospheres. The Ti2CTx and Ti2CTx/TiO2 films showed increased chemoresistance sensitivity to humidity: a significant decrease in the baseline was observed as the RH increased. Figure 6a shows the responses to humidity variations in the air atmosphere. As can be seen, when the RH was increased from 30% to 40%, 50%, 64% and 93%, the response value increased from 0% to 46%, 67%, 83% and 96% and to 50%, 76%, 87% and 91% for Ti2CTx and Ti2CTx/TiO2, respectively. The initial Ti2CTx film showed a higher sensitivity to humidity at elevated RH values (93%). The high MXenes sensitivity to humidity is typical and is well described in the literature. For further gas-sensitive measurements, a gas atmosphere with constant humidity RH = 50% was chosen.
In the following step, the chemoresistive responses were determined for the detection of 100 ppm CO, NH3, benzene (C6H6), acetone (C3H6O), ethanol (C2H5OH) and 1000 ppm methane (CH4), H2. The experimental responses (S1) to the above gases are shown in Figure 7a,b. The selectivity diagrams (Figure 7c,d) are plotted from the obtained response array. The bar chart (Figure 7c) shows the sign of the resulting response: a positive value corresponds to an increase in electrical resistance when the analyte gas is injected (p-response), and a negative value corresponds to a decrease (n-response). The radar chart (Figure 7d) shows the response values (S1 in %) for all analyzed gases.
For the initial Ti2CTx MXene film containing its oxidation products in water dispersion, the highest response (24%) was observed when NH3 was detected. Notable responses were recorded for CO (15%), C6H6 (13%), C2H5OH (11.5%) and C3H6O (8.3%); the response for H2 and CH4 detection was 4%. The sensitivity to 0.4–10% oxygen was also additionally measured (Figure 6b). A gradual increase in the response (S1) from 4% to 33% was observed when the oxygen concentration was increased from 1% to 10%. The response to oxygen by MXenes is practically not described in the literature but was nevertheless recorded by us in our previous study [29]. When all the above gases were injected, regardless of their chemical nature, the resistance decreased, i.e., an n-response was observed.
A significant change in gas-sensitive properties was observed for the Ti2CTx/TiO2 composite film obtained by oxidation in air at 316 °C. The highest response (61%) was recorded for NH3 detection, as for the initial Ti2CTx film. Further, notable responses were obtained for the detection of C2H5OH (32%) and NO2 (43%), and the response for all other gases did not exceed 7.6%. As can be clearly seen from the selectivity radar chart (Figure 7d), the oxidized Ti2CTx/TiO2 film showed an increased response to all analyzed gases (except oxygen) compared to the Ti2CTx film. Especially noticeable is the sensitivity to NO2 (the response increased by 10 times), as well as a completely absent response to oxygen. In contrast to the Ti2CTx film, for the air-oxidized Ti2CTx/TiO2 film, a different type of response was observed for the detection of almost all gases (except for NH3 and NO2, to which the maximum response was observed). When detecting CO, C6H6, C3H6O, C2H5OH, CH4, H2 and O2, a p-type response was obtained, and when detecting nitrogen-containing gases NH3 and NO2 an n-type response was obtained.
Figure 8 shows the response of 4–100 ppm NH3 and the signal reproducibility when detecting 10 ppm NH3 by the Ti2CTx/TiO2 composite film. As can be seen (Figure 8a), with an increase in NH3 concentration from 4 to 100 ppm, there is a gradual increase in the response (S1) from 16% to 61% with a noticeable drift of the baseline (which can be explained by the high interaction energy of the ammonia molecule and receptor material), and the response itself is well reproduced (Figure 8b)
During the consideration of the mechanisms of the detection of various gases by the obtained materials, it is necessary to take into account the fact that all measurements were carried out in an atmosphere of 50% RH. Therefore, it can be assumed that the MXene surface, in this case, contains a large amount of sorbed water, as well as hydroxyl groups (OH), i.e., hydroxyl groups can be directly involved in reactions with various gases that precede the chemoresistive response.
In the present work, only the n-type response was observed for the initial Ti2CTx sample. When detecting the reducing gases (all gases in this work except NO2 and O2), this behavior is typical for n-type metal oxide semiconductors [52] but is not typical for Tin+1CnTx MXenes, for which a p-type response is most often observed [15,18,53]. Raman spectroscopy and TEM data revealed that the surface of the initial MXene sheets contains oxygen groups, and an impurity of the rutile TiO2 phase was noted. Thus, in addition to the hydroxyl functional groups of MXene, as well as sorbed water molecules, TiO2 particles may also be involved in the detection mechanism. The presence of the n-type response detected in our study is not fully understood at this time and requires further investigation.
After oxidation of the initial MXene film in air to Ti2CTx/TiO2 composite, a change in the response character from n- to p-type is observed when detecting CO, H2 and VOC’s. According to Raman spectroscopy data, it is found that a single n-type TiO2 semiconductor phase is formed for this sample. The surface microstructure of the resulting composite also changes. As a rule, large-band-gap semiconductors, such as TiO2, have poor electrical conductivity at room temperature. At room temperature, it is the MXene phase, which is characterized by p-type conductivity, that makes the main contribution to the gas sensitivity. The increase in sensing signals can be related to the fact that at the Ti2CTx/TiO2 interface, a Schottky barrier appears, the height of which will change during the adsorption of different gases [23].
Table 1 provides information on the gas-sensitive properties of composites of MXenes Ti3C2Tx/TiO2 and Ti2CTx/TiO2 operating at room temperature for chemoresistive gas sensors reported in the literature. As can be seen, in the field of gas-sensing composites based on Ti3C2Tx, MXenes are mainly used, and the sensory properties of Ti2CTx MXenes and nanocomposite Ti2CTx/TiO2 are practically not studied. For many of the composites presented in Table 1, it is ammonia that shows the greatest response. This corresponds to the data of theoretical calculations [54] for the maximum oxidized compound, Ti2CO2, which is associated with high adsorption energy and the calculated value of charge transfer. A comparison of the experimental data on the gas-sensitive properties shows that the materials obtained in this work have a sufficiently high response (to a wide range of NH3 concentrations), significantly exceeding most of the literature analogs. The increased response can be attributed to the more flexible heat treatment conditions used in this work under the control of in situ Raman spectroscopy for the controlled formation of heterojunctions between Ti2CTx and TiO2. In addition, the responses obtained in this work were recorded at relative humidity RH = 50% (closest to the real conditions), which has not been noted in other works for TiO2-modified MXenes.

4. Conclusions

The in situ oxidation of the film of layered Ti2CTx MXene synthesized as a result of the exposure of the MAX-phase of Ti2AlC to sodium fluoride solution in hydrochloric acid, delaminated by tetramethylammonium hydroxide and held in water dispersion, has been studied, in detail, using Raman spectroscopy. The phenomenon of substrate (Al2O3 ceramics or platinum) influence on the intensity of Ti2CTx MXene phase modes, as well as the modes formed during oxidation in the air atmosphere of rutile and anatase phases, has been noted. It is shown that the formation of a new oxidation product phase (TiO2 with anatase structure) on the ceramic Al2O3 substrate can be fixed at a lower temperature (316 °C) than it is typical for the Ti2CTx layer on the platinum electrodes of the specialized substrate (372 °C). It was found that at 447 °C the full oxidation of the initial Ti2CTx MXene under the conditions of heating in an air atmosphere occurs.
The coating microstructure was studied using a complex of methods (SEM, TEM and AFM), and the work function was determined for both the Ti2CTx/TiO2 composite obtained at 316 °C and the fully oxidized composition using Kelvin-probe force microscopy.
It was found that the obtained Ti2CTx and Ti2CTx/TiO2 films exhibit chemoresistive responses when detecting a wide group of gases, H2, CO, NH3, C6H6, C3H6O, CH4, C2H5OH and O2, at room temperature and 50% relative humidity. It was shown that partial oxidation of MXene favors the gas-sensitive properties, and the most sensitivity was observed for NH3.

Author Contributions

Conceptualization, A.S.M. and E.P.S.; Methodology, I.A.N. and E.P.S.; Validation, E.P.S.; Formal analysis, P.Y.G., A.A.A., T.L.S. and N.P.S.; Investigation, A.S.M.; Writing—original draft, A.S.M.; Writing—review & editing, E.P.S. and N.T.K.; Visualization, A.S.M.; Supervision, N.T.K.; Funding acquisition, N.P.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Russian Science Foundation, project No. 21-73-10251, https://rscf.ru/en/project/21-73-10251/ (accessed on 29 November 2022).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare that they have no known competing financial interest or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Naguib, M.; Mashtalir, O.; Carle, J.; Presser, V.; Lu, J.; Hultman, L.; Gogotsi, Y.; Barsoum, M.W. Two-Dimensional Transition Metal Carbides. ACS Nano 2012, 6, 1322–1331. [Google Scholar] [CrossRef] [PubMed]
  2. Naguib, M.; Mochalin, V.N.; Barsoum, M.W.; Gogotsi, Y. 25th Anniversary Article: MXenes: A New Family of Two-Dimensional Materials. Adv. Mater. 2014, 26, 992–1005. [Google Scholar] [CrossRef] [PubMed]
  3. 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 Ti 3AlC 2. Adv. Mater. 2011, 23, 4248–4253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Kumar, K.S.; Choudhary, N.; Jung, Y.; Thomas, J. Recent Advances in Two-Dimensional Nanomaterials for Supercapacitor Electrode Applications. ACS Energy Lett. 2018, 3, 482–495. [Google Scholar] [CrossRef]
  5. Aslam, M.K.; Xu, M. A Mini-Review: MXene Composites for Sodium/Potassium-Ion Batteries. Nanoscale 2020, 12, 15993–16007. [Google Scholar] [CrossRef]
  6. Zhong, Y.; Xia, X.H.; Shi, F.; Zhan, J.Y.; Tu, J.P.; Fan, H.J. Transition Metal Carbides and Nitrides in Energy Storage and Conversion. Adv. Sci. 2016, 3, 1500286. [Google Scholar] [CrossRef]
  7. Khan, K.; Tareen, A.K.; Aslam, M.; Zhang, Y.; Wang, R.; Ouyang, Z.; Gou, Z.; Zhang, H. Recent Advances in Two-Dimensional Materials and Their Nanocomposites in Sustainable Energy Conversion Applications. Nanoscale 2019, 11, 21622–21678. [Google Scholar] [CrossRef]
  8. Li, K.; Zhang, S.; Li, Y.; Fan, J.; Lv, K. MXenes as Noble-Metal-Alternative Co-Catalysts in Photocatalysis. Chin. J. Catal. 2020, 42, 3–14. [Google Scholar] [CrossRef]
  9. Xie, X.; Zhang, N. Positioning MXenes in the Photocatalysis Landscape: Competitiveness, Challenges, and Future Perspectives. Adv. Funct. Mater. 2020, 30, 2002528. [Google Scholar] [CrossRef]
  10. Zhao, Q.; Zhou, W.; Zhang, M.; Wang, Y.; Duan, Z.; Tan, C.; Liu, B.; Ouyang, F.; Yuan, Z.; Tai, H.; et al. Edge-Enriched Mo2TiC2Tx/MoS2 Heterostructure with Coupling Interface for Selective NO2 Monitoring. Adv. Funct. Mater. 2022, 32, 2203528. [Google Scholar] [CrossRef]
  11. Wang, H.; Shi, X.; Liu, F.; Duan, T.; Sun, B. Non-Invasive Rapid Detection of Lung Cancer Biomarker Toluene with a Cataluminescence Sensor Based on the Two-Dimensional Nanocomposite Pt/Ti3C2Tx-CNT. Chemosensors 2022, 10, 333. [Google Scholar] [CrossRef]
  12. Choi, S.J.; Kim, I.D. Recent Developments in 2D Nanomaterials for Chemiresistive-Type Gas Sensors. Electron. Mater. Lett. 2018, 14, 221–260. [Google Scholar] [CrossRef]
  13. Lee, E.; Kim, D.-J. Review—Recent Exploration of Two-Dimensional MXenes for Gas Sensing: From a Theoretical to an Experimental View. J. Electrochem. Soc. 2020, 167, 037515. [Google Scholar] [CrossRef] [Green Version]
  14. Devaraj, M.; Rajendran, S.; Hoang, T.K.A.; Soto-Moscoso, M. A Review on MXene and Its Nanocomposites for the Detection of Toxic Inorganic Gases. Chemosphere 2022, 302, 134933. [Google Scholar] [CrossRef]
  15. 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] [Green Version]
  16. Ji, H.; Zeng, W.; Li, Y. Gas Sensing Mechanisms of Metal Oxide Semiconductors: A Focus Review. Nanoscale 2019, 11, 22664–22684. [Google Scholar] [CrossRef]
  17. Khakbaz, P.; Moshayedi, M.; Hajian, S.; Soleimani, M.; Narakathu, B.B.; Bazuin, B.J.; Pourfath, M.; Atashbar, M.Z. Titanium Carbide MXene as NH3 Sensor: Realistic First-Principles Study. J. Phys. Chem. C 2019, 123, 29794–29803. [Google Scholar] [CrossRef]
  18. Majhi, S.M.; Ali, A.; Greish, Y.E.; El-Maghraby, H.F.; Qamhieh, N.N.; Hajamohideen, A.R.; Mahmoud, S.T. Accordion-like-Ti3C2 MXene-Based Gas Sensors with Sub-Ppm Level Detection of Acetone at Room Temperature. ACS Appl. Electron. Mater. 2022, 4, 4094–4103. [Google Scholar] [CrossRef]
  19. 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]
  20. Zhang, C.; Zhang, Y.; Cao, K.; Guo, Z.; Han, Y.; Hu, W.; Wu, Y.; She, Y.; He, Y. Ultrasensitive and Reversible Room-Temperature Resistive Humidity Sensor Based on Layered Two-Dimensional Titanium Carbide. Ceram. Int. 2021, 47, 6463–6469. [Google Scholar] [CrossRef]
  21. Koh, H.J.; Kim, S.J.; Maleski, K.; Cho, S.Y.; Kim, Y.J.; Ahn, C.W.; Gogotsi, Y.; Jung, H.T. Enhanced Selectivity of MXene Gas Sensors through Metal Ion Intercalation: In Situ X-Ray Diffraction Study. ACS Sens. 2019, 4, 1365–1372. [Google Scholar] [CrossRef] [PubMed]
  22. Wu, M.; An, Y.; Yang, R.; Tao, Z.; Xia, Q.; Hu, Q.; Li, M.; Chen, K.; Zhang, Z.; Huang, Q.; et al. V2CTx and Ti3C2Tx MXenes Nanosheets for Gas Sensing. ACS Appl. Nano Mater. 2021, 4, 6257–6268. [Google Scholar] [CrossRef]
  23. Pazniak, H.; Plugin, I.A.; Loes, M.J.; Inerbaev, T.M.; Burmistrov, I.N.; Gorshenkov, M.; Polcak, J.; Varezhnikov, A.S.; Sommer, M.; Kuznetsov, D.V.; et al. Partially Oxidized Ti3C2Tx MXenes for Fast and Selective Detection of Organic Vapors at Part-per-Million Concentrations. ACS Appl. Nano Mater. 2020, 3, 3195–3204. [Google Scholar] [CrossRef]
  24. Wang, Z.; Wang, F.; Hermawan, A.; Zhu, J.; Yin, S. Surface Engineering of Ti3C2Tx MXene by Oxygen Plasma Irradiation as Room Temperature Ethanol Sensor. Funct. Mater. Lett. 2022, 15, 2251007. [Google Scholar] [CrossRef]
  25. Sun, Q.; Wang, J.; Wang, X.; Dai, J.; Wang, X.; Fan, H.; Wang, Z.; Li, H.; Huang, X.; Huang, W. Treatment-Dependent Surface Chemistry and Gas Sensing Behavior of the Thinnest Member of Titanium Carbide MXenes. Nanoscale 2020, 12, 16987–16994. [Google Scholar] [CrossRef]
  26. Liu, S.; Wang, M.; Liu, G.; Wan, N.; Ge, C.; Hussain, S.; Meng, H.; Wang, M.; Qiao, G. Enhanced NO2 Gas-Sensing Performance of 2D Ti3C2/TiO2 Nanocomposites by In-Situ Formation of Schottky Barrier. Appl. Surf. Sci. 2021, 567, 150747. [Google Scholar] [CrossRef]
  27. Tai, H.; Duan, Z.; He, Z.; Li, X.; Xu, J.; Liu, B.; Jiang, Y. Enhanced Ammonia Response of Ti3C2Tx Nanosheets Supported by TiO2 Nanoparticles at Room Temperature. Sens. Actuators B Chem. 2019, 298, 126874. [Google Scholar] [CrossRef]
  28. Simonenko, E.P.; Simonenko, N.P.; Nagornov, I.A.; Simonenko, T.L.; Mokrushin, A.S.; Sevastyanov, V.G.; Kuznetsov, N.T. Synthesis of MAX Phases in the Ti2AlC-V2AlC System as Precursors of Heterometallic MXenes Ti2–XVxC. Russ. J. Inorg. Chem. 2022, 67, 705–714. [Google Scholar] [CrossRef]
  29. Simonenko, E.P.; Simonenko, N.P.; Nagornov, I.A.; Simonenko, T.L.; Gorobtsov, P.Y.; Mokrushin, A.S.; Kuznetsov, N.T. Synthesis and Chemoresistive Properties of Single-Layer MXene Ti2CTx. Russ. J. Inorg. Chem. 2022, 67, 1838–1847. [Google Scholar] [CrossRef]
  30. Mokrushin, A.S.; Simonenko, T.L.; Simonenko, N.P.; Yu, P.; Bocharova, V.A.; Kozodaev, M.G.; Markeev, A.M.; Lizunova, A.A.; Volkov, I.A.; Simonenko, E.P.; et al. Microextrusion Printing of Gas-Sensitive Planar Anisotropic NiO Nanostructures and Their Surface Modification in an H2S Atmosphere. Appl. Surf. Sci. 2022, 578, 151984. [Google Scholar] [CrossRef]
  31. Mokrushin, A.S.; Nagornov, I.A.; Simonenko, T.L.; Simonenko, N.P.; Yu, P.; Khamova, T.V.; Kopitsa, G.P.; Evzrezov, A.N.; Simonenko, E.P.; Sevastyanov, V.G.; et al. Chemoresistive Gas-Sensitive ZnO/Pt Nanocomposites Films Applied by Microplotter Printing with Increased Sensitivity to Benzene and Hydrogen. Mater. Sci. Eng. B 2021, 271, 115233. [Google Scholar] [CrossRef]
  32. Melchior, S.A.; Raju, K.; Ike, I.S.; Erasmus, R.M.; Kabongo, G.; Sigalas, I.; Iyuke, S.E.; Ozoemena, K.I. High-Voltage Symmetric Supercapacitor Based on 2D Titanium Carbide (MXene, Ti2CTx)/Carbon Nanosphere Composites in a Neutral Aqueous Electrolyte. J. Electrochem. Soc. 2018, 165, A501–A511. [Google Scholar] [CrossRef] [Green Version]
  33. Habib, I.; Ferrer, P.; Ray, S.C.; Ozoemena, K.I. Interrogating the Impact of Onion-like Carbons on the Supercapacitive Properties of MXene (Ti2CTX). J. Appl. Phys. 2019, 126, 134301. [Google Scholar] [CrossRef]
  34. Lioi, D.B.; Neher, G.; Heckler, J.E.; Back, T.; Mehmood, F.; Nepal, D.; Pachter, R.; Vaia, R.; Kennedy, W.J. Electron-Withdrawing Effect of Native Terminal Groups on the Lattice Structure of Ti3C2Tx MXenes Studied by Resonance Raman Scattering: Implications for Embedding MXenes in Electronic Composites. ACS Appl. Nano Mater. 2019, 2, 6087–6091. [Google Scholar] [CrossRef]
  35. Spanier, J.E.; Gupta, S.; Amer, M.; Barsoum, M.W. Vibrational Behavior of the Mn+1AXn Phases from First-Order Raman Scattering (M = Ti, V, Cr, A = Si, X = C, N). Phys. Rev. B 2005, 71, 012103. [Google Scholar] [CrossRef] [Green Version]
  36. Ma, R.; Fukuda, K.; Sasaki, T.; Osada, M.; Bando, Y. Structural Features of Titanate Nanotubes/Nanobelts Revealed by Raman, X-Ray Absorption Fine Structure and Electron Diffraction Characterizations. J. Phys. Chem. B 2005, 109, 6210–6214. [Google Scholar] [CrossRef]
  37. Ma, H.L.; Yang, J.Y.; Dai, Y.; Zhang, Y.B.; Lu, B.; Ma, G.H. Raman Study of Phase Transformation of TiO2 Rutile Single Crystal Irradiated by Infrared Femtosecond Laser. Appl. Surf. Sci. 2007, 253, 7497–7500. [Google Scholar] [CrossRef]
  38. Frank, O.; Zukalova, M.; Laskova, B.; Kürti, J.; Koltai, J.; Kavan, L. Raman Spectra of Titanium Dioxide (Anatase, Rutile) with Identified Oxygen Isotopes (16, 17, 18). Phys. Chem. Chem. Phys. 2012, 14, 14567–14572. [Google Scholar] [CrossRef]
  39. Zhang, J.Z.; Shen, Y.D.; Li, Y.W.; Hu, Z.G.; Chu, J.H. Composition Dependence of Microstructure, Phonon Modes, and Optical Properties in Rutile TiO2:Fe Nanocrystalline Films Prepared by a Nonhydrolytic Sol-Gel Route. J. Phys. Chem. C 2010, 114, 15157–15164. [Google Scholar] [CrossRef]
  40. Gan, Y.; Zhu, F.; Shi, Y.; Wen, Q. Single Frequency Fiber Laser Base on MXene with KHz Linewidth. J. Mater. Chem. C 2021, 9, 2276–2281. [Google Scholar] [CrossRef]
  41. Murugan, N.; Jerome, R.; Preethika, M.; Sundaramurthy, A.; Sundramoorthy, A.K. 2D-Titanium Carbide (MXene) Based Selective Electrochemical Sensor for Simultaneous Detection of Ascorbic Acid, Dopamine and Uric Acid. J. Mater. Sci. Technol. 2021, 72, 122–131. [Google Scholar] [CrossRef]
  42. Liu, X.; Ji, L.; Zhu, F.; Gan, Y.; Wen, Q. Linear-Cavity-Based Single Frequency Fiber Laser with a Loop Mirror and Ti2CTx Quantum Dots. Opt. Mater. 2021, 122, 111686. [Google Scholar] [CrossRef]
  43. Tan, Z.; Sato, K.; Ohara, S. Synthesis of Layered Nanostructured TiO2 by Hydrothermal Method. Adv. Powder Technol. 2015, 26, 296–302. [Google Scholar] [CrossRef]
  44. Waterland, M.R.; Stockwell, D.; Kelley, A.M. Symmetry Breaking Effects in NO3-: Raman Spectra of Nitrate Salts and Ab Initio Resonance Raman Spectra of Nitrate-Water Complexes. J. Chem. Phys. 2001, 114, 6249–6258. [Google Scholar] [CrossRef]
  45. Waterland, M.R.; Kelley, A.M. Far-Ultraviolet Resonance Raman Spectroscopy of Nitrate Ion in Solution. J. Chem. Phys. 2000, 113, 6760–6773. [Google Scholar] [CrossRef]
  46. Choi, J.; Kim, Y.J.; Cho, S.Y.; Park, K.; Kang, H.; Kim, S.J.; Jung, H.T. In Situ Formation of Multiple Schottky Barriers in a Ti3C2 MXene Film and Its Application in Highly Sensitive Gas Sensors. Adv. Funct. Mater. 2020, 30, 2003998. [Google Scholar] [CrossRef]
  47. Wang, Z.; Wang, F.; Hermawan, A.; Asakura, Y.; Hasegawa, T.; Kumagai, H.; Kato, H.; Kakihana, M.; Zhu, J.; Yin, S. SnO-SnO2 Modified Two-Dimensional MXene Ti3C2Tx for Acetone Gas Sensor Working at Room Temperature. J. Mater. Sci. Technol. 2021, 73, 128–138. [Google Scholar] [CrossRef]
  48. Roy, A.; Sood, A.K. Phonons and Fractons in Sol-Gel Alumina: Raman Study. Pramana 1995, 44, 201–209. [Google Scholar] [CrossRef] [Green Version]
  49. Xu, J.; Shim, J.; Park, J.H.; Lee, S. MXene Electrode for the Integration of WSe2 and MoS2 Field Effect Transistors. Adv. Funct. Mater. 2016, 26, 5328–5334. [Google Scholar] [CrossRef]
  50. Khazaei, M.; Arai, M.; Sasaki, T.; Ranjbar, A.; Liang, Y.; Yunoki, S. OH-Terminated Two-Dimensional Transition Metal Carbides and Nitrides as Ultralow Work Function Materials. Phys. Rev. B 2015, 92, 075411. [Google Scholar] [CrossRef]
  51. Mansfeldova, V.; Zlamalova, M.; Tarabkova, H.; Janda, P.; Vorokhta, M.; Piliai, L.; Kavan, L. Work Function of TiO2 (Anatase, Rutile, and Brookite) Single Crystals: Effects of the Environment. J. Phys. Chem. C 2021, 125, 1902–1912. [Google Scholar] [CrossRef]
  52. Deng, Y. Sensing mechanism and evaluation criteria of semiconducting metal oxides gas sensors. In Semiconducting Metal Oxides for Gas Sensing; Springer: Singapore, 2019; pp. 23–51. ISBN 9789811358524. [Google Scholar]
  53. 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]
  54. Yu, X.F.; Li, Y.C.; Cheng, J.B.; Liu, Z.B.; Li, Q.Z.; Li, W.Z.; Yang, X.; Xiao, B. Monolayer Ti2CO2: A Promising Candidate for NH3 Sensor or Capturer with High Sensitivity and Selectivity. ACS Appl. Mater. Interfaces 2015, 7, 13707–13713. [Google Scholar] [CrossRef]
  55. Hou, M.; Guo, S.; Yang, L.; Gao, J.; Hu, T.; Wang, X.; Li, Y. Improvement of Gas Sensing Property for Two-Dimensional Ti3C2Tx Treated with Oxygen Plasma by Microwave Energy Excitation. Ceram. Int. 2021, 47, 7728–7737. [Google Scholar] [CrossRef]
  56. 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]
  57. Zhou, Y.; Wang, Y.; Wang, Y.; Yu, H.; Zhang, R.; Li, J.; Zang, Z.; Li, X. MXene Ti3C2Tx-Derived Nitrogen-Functionalized Heterophase TiO2 Homojunctions for Room-Temperature Trace Ammonia Gas Sensing. ACS Appl. Mater. Interfaces 2021, 13, 56485–56497. [Google Scholar] [CrossRef]
Figure 1. The X-ray patterns of the initial Ti2AlC, low-layer Ti2CTx MXene and its coating on a specialized sensor substrate (a), and TEM microphotographs of Ti2CTx MXene (b,c).
Figure 1. The X-ray patterns of the initial Ti2AlC, low-layer Ti2CTx MXene and its coating on a specialized sensor substrate (a), and TEM microphotographs of Ti2CTx MXene (b,c).
Chemosensors 11 00013 g001
Figure 2. The scheme of an Al2O3 substrate with Pt microelectrodes and Ti2CTx MXene film (a) and a microphotograph of the Ti2CTx film on the substrate surface with an indication of the areas where in situ Raman spectra were measured (b).
Figure 2. The scheme of an Al2O3 substrate with Pt microelectrodes and Ti2CTx MXene film (a) and a microphotograph of the Ti2CTx film on the substrate surface with an indication of the areas where in situ Raman spectra were measured (b).
Chemosensors 11 00013 g002
Figure 3. In situ Raman spectra at different temperatures (20–447 °C) of the thin film at different substrate areas: on Pt (a) and Al2O3 (b).
Figure 3. In situ Raman spectra at different temperatures (20–447 °C) of the thin film at different substrate areas: on Pt (a) and Al2O3 (b).
Chemosensors 11 00013 g003
Figure 4. The SEM micrographs of films on a special Al2O3/Pt substrate: initial MXene (a,b), oxidized at 316 °C (c,d) and 447 °C (e,f).
Figure 4. The SEM micrographs of films on a special Al2O3/Pt substrate: initial MXene (a,b), oxidized at 316 °C (c,d) and 447 °C (e,f).
Chemosensors 11 00013 g004
Figure 5. The AFM results: topography (a) and surface potential distribution map (b) of Ti2CTx/TiO2 MXene film oxidized at 316 °C; topography (c) and surface potential distribution map (d) of Ti2CTx/TiO2 film oxidized at 447 °C.
Figure 5. The AFM results: topography (a) and surface potential distribution map (b) of Ti2CTx/TiO2 MXene film oxidized at 316 °C; topography (c) and surface potential distribution map (d) of Ti2CTx/TiO2 film oxidized at 447 °C.
Chemosensors 11 00013 g005
Figure 6. The responses of Ti2CTx and Ti2CTx/TiO2 films to humidity (a) and Ti2CTx films to 0.4–10% O2 at 50%RH (b).
Figure 6. The responses of Ti2CTx and Ti2CTx/TiO2 films to humidity (a) and Ti2CTx films to 0.4–10% O2 at 50%RH (b).
Chemosensors 11 00013 g006
Figure 7. The gas-sensing properties of the samples at RT and 50%RH: responses to 100 and 1000 ppm of different gases of the Ti2CTx film (a) and Ti2CTx/TiO2 (b); selectivity diagrams plotted from the responses to the different gases (c,d).
Figure 7. The gas-sensing properties of the samples at RT and 50%RH: responses to 100 and 1000 ppm of different gases of the Ti2CTx film (a) and Ti2CTx/TiO2 (b); selectivity diagrams plotted from the responses to the different gases (c,d).
Chemosensors 11 00013 g007
Figure 8. The gas-sensitive properties of the samples at RT and 50%RH. Responses of Ti2CTx/TiO2 films to 4–100 ppm NH3 (a) and 10 ppm NH3 (b).
Figure 8. The gas-sensitive properties of the samples at RT and 50%RH. Responses of Ti2CTx/TiO2 films to 4–100 ppm NH3 (a) and 10 ppm NH3 (b).
Chemosensors 11 00013 g008
Table 1. Comparison of gas-sensitive characteristics of chemoresistive gas sensors based on Ti3C2Tx/TiO2 and Ti2CTx/TiO2 MXenes operating at room temperature that is presented in the literature.
Table 1. Comparison of gas-sensitive characteristics of chemoresistive gas sensors based on Ti3C2Tx/TiO2 and Ti2CTx/TiO2 MXenes operating at room temperature that is presented in the literature.
No.YearCompositionTarget GasConc., ppmResponseRH, %Ref.
12019Ti3C2Tx/TiO2NH310 ppm3%60[27]
22020Ti2CTx/TiO2NH310 ppm1.9%0[25]
32020Ti3C2Tx/TiO2NO25 ppm16.05%0[46]
42021Ti3C2Tx/TiO2C2H5OH100 ppm22.47%0[55]
52021Ti3C2Tx/TiO2hexanal100 ppm8.8%0[56]
62021Ti3C2Tx/TiO2NO2100 ppm4%0[26]
72021Ti3C2Tx (N-doped)/TiO2NH3200 ppb7.3%0[57]
82022Ti3C2Tx/TiO2C2H5OH90 ppm91 a.u.0[24]
92022Ti2CTx/TiO2NH34–100 ppm 16–61%50This work
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

Mokrushin, A.S.; Nagornov, I.A.; Gorobtsov, P.Y.; Averin, A.A.; Simonenko, T.L.; Simonenko, N.P.; Simonenko, E.P.; Kuznetsov, N.T. Effect of Ti2CTx MXene Oxidation on Its Gas-Sensitive Properties. Chemosensors 2023, 11, 13. https://doi.org/10.3390/chemosensors11010013

AMA Style

Mokrushin AS, Nagornov IA, Gorobtsov PY, Averin AA, Simonenko TL, Simonenko NP, Simonenko EP, Kuznetsov NT. Effect of Ti2CTx MXene Oxidation on Its Gas-Sensitive Properties. Chemosensors. 2023; 11(1):13. https://doi.org/10.3390/chemosensors11010013

Chicago/Turabian Style

Mokrushin, Artem S., Ilya A. Nagornov, Philipp Yu. Gorobtsov, Aleksey A. Averin, Tatiana L. Simonenko, Nikolay P. Simonenko, Elizaveta P. Simonenko, and Nikolay T. Kuznetsov. 2023. "Effect of Ti2CTx MXene Oxidation on Its Gas-Sensitive Properties" Chemosensors 11, no. 1: 13. https://doi.org/10.3390/chemosensors11010013

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