Next Article in Journal
Electrochemical Redox Cycling Behavior of Gold Nanoring Electrodes Microfabricated on a Silicon Micropillar
Previous Article in Journal
Identification of the Position of a Tethered Delivery Catheter to Retrieve an Untethered Magnetic Robot in a Vascular Environment
Previous Article in Special Issue
Tapered Optical Fiber Sensor Coated with Single-Walled Carbon Nanotubes for Dye Sensing Application
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Gas-Sensitive Properties of ZnO/Ti2CTx Nanocomposites

by
Elizaveta P. Simonenko
1,*,
Ilya A. Nagornov
1,
Artem S. Mokrushin
1,
Aleksey A. Averin
2,
Yulia M. Gorban
1,3,
Tatiana L. Simonenko
1,
Nikolay P. Simonenko
1 and
Nikolay T. Kuznetsov
1
1
Kurnakov Institute of General and Inorganic Chemistry, Russian Academy of Sciences, Moscow 119991, Russia
2
Frumkin Institute of Physical Chemistry and Electrochemistry, Russian Academy of Sciences, Moscow 199071, Russia
3
Mendeleev University of Chemical Technology of Russia, Moscow 125047, Russia
*
Author to whom correspondence should be addressed.
Micromachines 2023, 14(4), 725; https://doi.org/10.3390/mi14040725
Submission received: 27 February 2023 / Revised: 17 March 2023 / Accepted: 22 March 2023 / Published: 24 March 2023
(This article belongs to the Special Issue Advanced Functional Nanostructures for Chemical Sensing)

Abstract

:
At present, a new class of 2D nanomaterials, MXenes, is of great scientific and applied interest, and their application prospects are very broad, including as effective doping components for receptor materials of MOS sensors. In this work we have studied the influence on the gas-sensitive properties of nanocrystalline zinc oxide synthesized by atmospheric pressure solvothermal synthesis, with the addition of 1–5% of multilayer two-dimensional titanium carbide Ti2CTx, obtained by etching Ti2AlC with NaF solution in hydrochloric acid. It was found that all the obtained materials have high sensitivity and selectivity with respect to 4–20 ppm NO2 at a detection temperature of 200 °C. It is shown that the selectivity towards this compound is best for the sample containing the highest amount of Ti2CTx dopant. It has been found that as the MXene content increases, there is an increase in nitrogen dioxide (4 ppm) from 1.6 (ZnO) to 20.5 (ZnO-5 mol% Ti2CTx). reactions which the responses to nitrogen dioxide increase. This may be due to the increase in the specific surface area of the receptor layers, the presence of MXene surface functional groups, as well as the formation of the Schottky barrier at the interface between the phases of the components.

1. Introduction

The development of synthesis methods and research into receptor materials for chemical gas sensors remains one of the most important tasks in the field of functional materials creation [1], because with the development of industry, environmental problems and their impact on human health are becoming more acute. The detection of both toxic gaseous contaminants in the air and marker gases for various serious diseases requires further increases in the sensitivity of sensor materials as the concentrations to be detected move from the ppm level to the ppb and ppt range [2]. For example, ammonia has been found to be the most important respiratory metabolite with a concentration of 833 ppb, followed by acetone (477 ppb), methanol (461 ppb), ethanol (112 ppb), isoprene (106 ppb), acetaldehyde (22 ppb), propanol (18 ppb), etc.
Perhaps an even more urgent task is to create materials with maximum selectivity with respect to a particular gaseous component, since real measurements for environmental or human health monitoring are inevitably associated with operation in complex gaseous media containing many interfering impurities in addition to the target component. Certainly, the use of multi-sensor devices with mathematical processing of a signal by methods PCA (principal component analysis), LDA (linear discriminant analysis) and others, as shown by the studies carried out [3,4,5,6,7], allows researchers to distinguish and define the concentration even when close to structural substances (for example, from a set of aliphatic alcohols) [8,9]. However, the use of receptor materials with the most varied selectivity in the set will significantly increase the accuracy of determination of individual components of the gaseous medium.
Zinc oxide is one of the most popular metal oxide semiconductor sensor (MOS sensor) materials, probably second only to the traditional tin oxide SnO2 [10]. It is known to be highly sensitive to toxic analytes such as NO2 [11,12,13], H2S [14,15], NH3 [16,17], as well as to volatile organic compounds [18,19,20,21]. The negative characteristics of this promising substance include rather low selectivity as well as a relatively high detection temperature, mainly in the temperature range of 200 to 400 °C. Studies by many scientific groups show that these disadvantages of ZnO can be compensated to some extent by doping it with various components: n- and p-type semiconductors [22,23,24,25]; noble metal nanoparticles [26,27,28,29,30]; carbon nanotubes or graphene [31,32,33,34,35]; etc.
In recent years, a new class of substances with interesting gas-sensitive properties—the MXenes—has attracted much attention [36,37,38,39,40,41,42,43,44,45,46,47]. These compounds have the general formula Mn+1CnTx (where M is a transition metal from the series Ti, V, Nb, Mo, etc., and T; surface functional groups, most often -F, -OH, =O, -Cl), which can exhibit both metallic and semiconducting properties depending on the method of synthesis [48]. They have been found to exhibit gas sensitivity to some gaseous components (NH3, NO2, volatile organic compounds or VOC) even at room temperature [49,50,51,52,53], which is extremely useful for the fabrication of sensors on flexible polymer substrates for their incorporation into portable electronic devices [54]. However, individual MXenes are characterized by very low response values (fractions and units of per cent) and sensor response and recovery times (tens of minutes) that are too long for practically important tasks [55,56,57].
Nevertheless, the use of MXenes as a doping component in MOS sensors often allows for the improvement of properties, e.g., in increasing the sensitivity to certain gases or in reducing the operating temperature of the sensor. For example, in [58] it was shown that growing ZnO nanorods on zinc oxide particles deposited on the surface of Ti3C2Tx (content ~36.5%) produces a composite material with high sensitivity to NO2 (up to the ppt level) when activated by UV irradiation. The work by of Liu et al. [59] investigated the effect of doping ZnO with low concentrations of Ti3C2Tx (0.5–3%) on its sensitivity to NO2. It was found that this leads to the possibility of lowering the detection temperature from 220 to 160 °C—which the authors attribute to an increase in the specific surface area and the number of adsorptions centers—an increase in the electrical conductivity of Ti3C2Tx (which allows rapid transport of charge carriers) and the formation of the Schottky barrier at the interface. In a study by Zhu [60] for ZnO/Ti3C2Tx nanocomposites (the amount of MXene varied from 1 to 3 wt.%), a higher sensitivity to acetone at the operating temperature of 320 °C was observed for samples containing delaminated MXene compared to the accordion-like ones. In the article by Bu et al. [61] it was shown that doping zinc oxide with ~5.7 wt.% Ti3C2Tx does not allow lowering the detection temperature with maximum accuracy below 300 °C. However, a sharp increase of the response value at 100 ppm ethanol compared to single ZnO and Ti3C2Tx, low response/recovery times and good long-term stability of the detector were observed.
Improvements in the sensor properties of MXenes have also been observed when they are doped with ZnO nanoparticles [62,63], in particular increased sensitivity to the analyte and reduced response and recovery time of the sensor (especially under UV exposure).
Thus, an analysis of the literature shows that by doping ZnO with MXenes it is possible to improve its gas sensitive properties. However, the few available articles in this field focus on the effect of the most common Ti3C2Tx MXene, which was the first to be synthesized [64], while information on the doping of semiconductor nanodispersed oxides with another two-dimensional titanium carbide Ti2CTx is lacking.
The aim of this work is to develop methods for the synthesis of ZnO/Ti2CTx nanocomposites containing 1–5 mol% of multilayer MXene and to investigate their gas sensitive properties.

2. Materials and Methods

2.1. Synthesis and Application

Reagents: powders of metallic titanium (99.9%, 0.5–100 µm, Moscow, Russia, Ruskhim), aluminum (99.2%, 30 µm, Moscow, Russia, Ruskhim), graphite (MPG-8 grade, MPG-8, Technocarb, Chelyabinsk, Russia), potassium bromide KBr (99%, Moscow, Russia, Ruskhim), sodium fluoride NaF (99.9%, Moscow, Russia, Reackhim), hydrochloric acid HCl (>99%, Moscow, Russia, Sigma Tech).
The synthesis of the initial Ti2AlC MAX-phase is described in detail in our previous work [65]. In the present work: the powder ratio of n(Ti):n(Al):n(C) = 2:1.2:0.8; the weight of KBr powder added was equal to the total weight of titanium, aluminum and graphite powders; and the synthesis temperature in the molten salt was 1000 °C, the duration was 5 h.
The Ti2CTx MXene multilayer was obtained by the method described in [37,40] (without the delamination step). For this purpose, the selective etching of the Ti2AlC MAX-phase aluminum layers was carried out by exposure to NaF solution in hydrochloric acid (6M) at 40 °C 24 h. After extraction of the MXenes by centrifugation and washing with distilled water to pH~6–7, their dispersion in ethanol was subjected to an ultrasonic bath for 30 min. Then the powder was dried under a vacuum at 150 °C.
The synthesis of ZnO/Ti2CTx composites containing 1, 3 and 5 mol% of MXene (hereafter referred to as samples Z, Z1T, Z3T and Z5T, respectively) was carried out according to the adapted method [66,67]. For this purpose, a sample of MXene sample was dispersed under the influence of ultrasound in a solution of the precursor, monohydrate zinc acetylacetonate monohydrate in butanol, under ultrasonic exposure for 15 min. The reaction system was then heated in an oil bath at 145 ± 5 °C in a round-bottomed flask with a reflux condenser. The duration of solvothermal treatment was 6 h. The resulting precipitate was separated from the mother liquor by centrifugation (3500 rpm, 30 min), washed repeatedly with distilled water and ethyl alcohol solution (95 vol.%) to remove impurity organic fragments, and dried to constant weight under a vacuum at 150 °C.
The resulting ZnO/Ti2CTx composite powders were used to prepare the pastes required to form the gas sensitive layers of the sensor. For this purpose, the powder sample was ground in an agate mortar in 1-butanol for 1 h.
Receptor layers were deposited on specialized sensor substrates (Al2O3 wafers with platinum interdigital electrodes and a platinum heater on the backside) by microplotter printing [36,68,69,70]. The sensors were dried under reduced pressure at a temperature of 150 °C (1 h).

2.2. Instrumentation

The following instruments and methods were used to study the physico-chemical properties of the samples obtained. The following equipment was used to study the chemical composition of the surface of the powders and films, and to take microphotographs of the surface of the samples: NVision 40 scanning electron microscope, Carl Zeiss, (Oberkochen, Germany, secondary electron detector, accelerating voltage 1–10 kV); transmission electron microscope (JEOL, JEM-1011, Akishima, Japan); and INCA X-MAX 80 energy dispersive X-ray (EDX) spectrometer, Oxford Instruments (Oxford, UK), accelerating voltage 20 kV. Phase composition was investigated using a D8 Advance (Bruker, Billerica, MA, USA, CuKα = 1.5418 Å, Ni filter, E = 40 keV, I = 40 mA; 2θ range: 5–45°; Resolution: 0.02°; point accumulation time: 0.3 s). The thermal behavior of the synthesized Ti2CTx-MXene powder was studied using a combined DSC/DTA/TG analyzer SDT-Q600 (TA Instruments) in Al2O3 crucibles in air and Ar streams (250 mL/min); heating rate 10°/min, temperature range 25–1000 °C.
Raman spectra were obtained using an inVia Reflex «Renishaw» spectrometer (New Mills, Wotton-under-Edge, Gloucestershire, United Kingdom, GL12 8JR) with a 405 nm diode laser as the excitation source. All spectra were recorded in the 100–1750 cm−1 range with a spectral resolution of ~3 cm−1 through a 50x (NA 0.5, FN 26.5) magnification lens with an irradiated spot diameter of ~2 μm. The incident power was less than 0.2 mW. Lattice: 2400, signal accumulation time was 300 s.
The gas sensitive properties were measured using a specialised precision apparatus. The gas environment in the quartz cell was created using four Bronkhorst gas flow controllers with maximum flow rates of 1, 50, 100 and 200 mL/min. The temperature of the sensor element was controlled by a built-in platinum microheater, which was pre-calibrated using a Testo 868 thermal imager. The resulting film was tested for sensitivity to the following analyte gases: CO, H2, CH4, NH3, benzene, acetone, ethanol and NO2. The corresponding test gas mixtures in air were used as the source of the analyzed gases and synthetic air was used to establish a baseline. To measure the signal at different relative humidity (RH), we used a special unit with a barbometer, while the RH of the gas mixture was controlled by a digital flow hygrometer “Excis” (EXIS, Russia, Moscow). The electrical resistance of the oxide films was measured using a Fluke 8846A (6.5-digit precision multimeter) with an upper limit of 1 GΩ. The sensitivity to CO, H2, CH4, NH3, benzene, acetone and ethanol was calculated using the following formula.
S = RAir/R,
where RAir is the resistance of the oxide film in the synthetic air medium; R—in the medium with a given concentration of the analyte gas. The response to NO2 was calculated using the inverse relationship (1).
The selectivity coefficient was calculated according to the formula:
Sel = SNO2/Sg,
where SNO2 is the response at 4 ppm NO2; Sg is the response at a given gas concentration where the highest response value was observed after NO2.

3. Results and Discussion

3.1. Synthesis and Investigation of Ti2CTx MXene Powder and ZnO/Ti2CTx Composite Powders

As can be seen in Figure 1a, the Ti2CTx MXene powder obtained by selective MAX-phase etching contains a small admixture of the parent compound Ti2AlC, but there are no common crystalline impurities of Al2O3, TiC, TiO2 and low-soluble aluminum fluorides including Na5Al3F14. The shift of reflex position (002) from 12.9 to 7.5° indicates an increase in the interlayer distance with removal of aluminum layers and formation of surface functional groups.
The investigation of the particle microstructure by TEM (Figure 1b–d) showed that in addition to the multi-layered, coarse, accordion-like Ti2CTx particles in the powder, a certain number of low-layered particles with increased defects are present. In addition to MXene layer stacks with relatively long planes (up to 1–1.5 μm), aggregates with much smaller diameters (up to 100–200 nm) are found in approximately equal amounts. A similar conclusion can be drawn from the analysis of the SEM micrographs (Figure 1e–g): the sample contains MXene stacks with different diameters. The small Ti2CTx particles could not be fixed by this method. The EDX analysis of the Ti2CTx powder shows that the aluminum content does not exceed 1.3 at. % and the ratio n(F):n(Cl) = 3.7.
The thermal behavior of Ti2CTx multilayer powder has been studied in a flow of inert gas and air, shown in Figure 2. As can be seen, in both cases a loss of mass begins at low temperatures, which is probably related primarily to the desorption of water molecules. In this case, up to the temperature range of 150 to 170 °C, the TGA curves for air and argon coincided, while at higher temperatures the mass of the sample in the air flow was significantly higher than in the argon flow. This may be related to the onset of oxidation of the most chemically active particles of the MXene low-layer MXene present in the sample. At temperatures > 385 °C during heating in the air flow (Figure 2a), the stabilization of the mass of the sample in the air flow is followed by a decrease in mass, then an increase (>445 °C) and again a decrease (>505 °C). These mass changes are associated with the parallel processes of detachment of MXene surface functional groups (mass decrease) and oxidation of two-dimensional titanium carbide, accompanied by exothermic effects with a maximum at temperatures of 428 and 532 °C. The total cumulative mass loss at 1000 °C was 8.2%.
However, when heated in an argon stream (Figure 2b), a stepwise decrease in mass is observed, which at low temperatures (<200–220 °C) is probably mainly due to the removal of adsorbed water molecules. Heating to a higher temperature resulted in an increase in the rate of mass decrease due to the removal of surface hydroxyl groups, which is completed at 380 °C. A further step in the mass loss is observed at temperatures > 700 °C, probably due to the decomposition of the MXene by detachment removal of the fluoride and chloride functional groups already present.
In order to study the behavior of Ti2CTx powder when heated to relatively low temperatures, at which water desorption and detachment of surface OH-groups are assumed, we studied the change in MXene reflex position (002) after its incubation in an argon stream at 250 °C under DSC/TGA experiment (heating rate was 5°/min), shown in Figure 3. It is shown that the mass loss in this regime was 9.1% (Figure 3a), which is practically the same as that observed for thermal analysis in an argon stream above a temperature of 380 °C (Figure 2b). At the same time, the mass of the sample was already stabilized after 15 min of heating to the temperature 250 °C. X-ray diffraction of the obtained sample in the interval 2θ = 6–10° confirmed the shift of the reflex position (002) of the Ti2CTx MXene from 7.5° to 8.1°, which indicates the decrease of the interlayer distance from 11.8 to 11.0 Å as a result of the desorption of water molecules. At the same time, keeping the sample in a humid atmosphere for 2 days leads to the initiation of the opposite process, shifting the reflex (002) position towards lower angles from 8.1 to 7.8°. The data obtained correspond to the situation described in [71] for another two-dimensional titanium carbide, Ti3C2Tx, the shift resulting from the purging of the chamber with nitrogen for 200 min to remove adsorbed and intercalated molecules of reflex position (002) and from the corresponding reduction of the interlayer distance by 0.84 Å.
The phase composition of the obtained ZnO/Ti2CTx composite powders was investigated by XRD (Figure 4a) and Raman spectroscopy (Figure 4b). The X-ray diffraction patterns of all samples show intense but broadened reflexes of well-crystallized zinc oxide with wurtzite structure (Figure 4a), whereas for the Z5T sample (containing maximum amount of Ti2CTx–5 mol.%) at 2θ = 7.5° an additional low intensity reflex of accordion-like Ti2CTx is observed. All the Raman spectra of the powders obtained were recorded at low power to avoid oxidation of the oxides by the laser, so the spectra were obtained with some noise. As can be seen from Figure 4b, the spectra of the ZnO and Z5T samples show a characteristic set of ZnO peaks of the wurtzite phase, at 333, 438, 578 and 1154 cm−1 belonging to modes (E2high -E2low), E2high, A1 (LO) and [2A1 (LO), E1 (LO), 2LO], respectively [72,73]. The Raman spectrum of the Ti2CTx MXene powder ω1-3 with broad MXene own bands with maxima at 247, 416 and 668 cm−1 [74,75], and a weakly intense D- and intense G-band at 1361 and 1583 cm−1 [76], related to carbon modes, are present. The D- and G-bands are characteristic for MXenes and other carbon systems with sp2-hybridization of the C-C bond. The spectrum of the Z5T sample lacks the intense ω1-3 and D-bands of MXenes due to the low content of MXene, but the G-band is clearly visible, which is a sign of the presence of carbon structures in the nanocomposites (in this case, MXenes).
TEM of powder sample Z5T with maximum MXene content (Figure 5) shows that in addition to the rod aggregates (diameter 10–30 nm and length 40–90 nm) common to ZnO nanoparticles synthesized by the above method, the sample also contains buds [77] in the sample in which these ZnO aggregates are attached to accordion-like (Figure 5b) and small layered MXene particles (Figure 5c). Despite the low MXene dopant content in the compositions, such aggregates in which there are phase contacts between ZnO and Ti2CTx occur quite frequently in the total particle mass, which may be due to the fact that Ti2CTx acted as a seed for the formation of the ZnO phase.
Scanning electron microscopy was performed on additively deposited receptor layers of ZnO and ZnO/Ti2CTx compositions on the surface of specialized substrates (Figure 6). As can be seen from the micrographs, when doped with MXene, uniform ZnO aggregates in the form of a porous and uniform layer (Figure 6a–c) are replaced by a less regular aggregate size distribution (Figure 6d–l). At the same time, the size and shape of the aggregates forming the receptor layer is determined by the MXene stacks morphology (length 1–2.5 μm and thickness 0.8–1.8 μm, which corresponds to the size of the initial Ti2CTx aggregates including the ZnO ‘coat’, Figure 1), whose packing leads to an increase in the average pore size. Another peculiarity of Z1T-Z5T samples is that due to the less dense packing of the ZnO rods in the cross-links, the proportion of unaggregated particles increases. As the ZnO content increases, the number of accordion-like ZnO particles appearing on the surface of the composite layer (partially uncoated by ZnO rods) increases. The size of the zinc oxide rods themselves does not vary significantly with changes in the composition of the n(ZnO):n(Ti2CTx) ratio: the length of a single rod varies from 50 to 100 nm and the thickness varies from 10 to 30 nm. The mutual distribution of the elements Zn, Ti and O is shown in Figure 7. As can be seen, for the sample Z5T with the highest Ti2CTx content the following is observed: the absence of ZnO is clearly visible in the place of titanium-containing inclusions, i.e., zinc oxide nanoparticles are localized around rather large accordion-like MXene aggregates.

3.2. Chemoresistance Properties of ZnO/Ti2CTx Composites

The obtained films of ZnO/Ti2CTx composition showed electrical conductivity, which allowed the study of their chemoresistive gas-sensitive properties only at elevated temperatures. The main phase of the studied nanocomposites is ZnO, a broad band n-type semiconductor (Eg = 3.37 eV), therefore the electrophysical properties of ZnO/Ti2CTx nanocomposites are mainly determined by ZnO, which is characterized by medium and high detection operating temperatures. Modification of ZnO with Ti2CTx MXene, which has metallic conductivity, did not significantly reduce the electrical resistance. All other chemosensory measurements were carried out at the optimum operating temperature of 200 °C, at which sufficient nanomaterial resistance was observed for the measurements and the structure of the Ti2CTx MXene was preserved (oxidation to TiO2 starts at temperatures around 300 °C to TiO2 [40]).
In a first step, detection responses were obtained for all samples for a broad group of analyte gases at specified concentrations (1000 ppm: H2 and CH4; 100 ppm: CO, C6H6, NH3, ethanol, acetone and 4 ppm NO2). The highest response, which was significantly higher than for all other gases, was obtained for NO2. Figure 8a shows a histogram of the selectivity, which indicates that the individual zinc oxide also has a high sensitivity with respect to other gases besides NO2 (the highest response was obtained for 100 ppm ethanol and acetone S = 3.1, while the response for 4 ppm NO2 was 5.0). When ZnO is modified with MXene, the response to 4 ppm NO2 increases to 7.6, 16.4 and 26.7 for Z1T, Z3T and Z5T respectively, and the response to other gases decreases to 1.3–1.9. For a convenient numerical estimation of the selectivity, a special coefficient (Sel) was calculated according to Formula (2); the higher the coefficient, the greater the difference between the responses to different gases (in this case, the responses to other gases were compared with the response to 4 ppm NO2) and the better the selectivity. Figure 8b shows the dependence of Sel on the MXene content of ZnO/Ti2CTx nanocomposites. It was found that the selectivity to NO2 detection improves with increasing MXene content in the nanocomposites.
Figure 9a shows the experimental data on the sensitivity of the obtained receptor materials to different concentrations of NO2. As can be seen, the zinc oxide-based coatings exhibit a high response over a wide range of very low NO2 concentrations (4–20 ppm), exceeding the threshold limit value (TLV) recommended by the National Institute for Occupational Safety and Health [78]. It was found that with increasing concentrations from 4 to 15 ppm NO2, the response value increased from 5 to 14.2, from 7.6 to 32.7, from 16.4 to 61.2 and from 26.7 to 73.3 for samples Z, Z1T, Z3T and Z5T respectively. Thus, the response to NO2 increased significantly over the whole concentration range with increasing as the MXene content was increased. For sample Z5T, which showed the highest sensitivity to NO2, the response cut off at 20 ppm NO2 because the electrical resistance of the material at 20 ppm NO2 exceeded 1 GΩ, which did not allow the gas sensitivity of the material to be correctly estimated on the equipment used in this study.
Figure 9b shows the dependence of the response on the concentration of NO2 in the gas mixture. The observed dependence of the sensor response on the NO2 concentration is well described by the Freundlich isotherm equation: S = kCa, where k and a are proportional and exponential constants representing the adsorption capacity and the adsorption enhancement, respectively. The character of the obtained dependences agrees well with the data previously obtained for ZnO-based nanocomposites [11,18].
Analyzing the shape of the received signals, it can be seen that at 4–8 ppm NO2 the shape of the signals from all samples is bell-shaped, whereas at 10–20 ppm it is almost rectangular. The shape of the signals has a direct effect on the response time (t90). For example, when detecting 4–8 ppm NO2, the ZnO sample shows a t90 in the range of 40–45 s and at 10–20 ppm it is about 30–35 s. When modified with Ti2CTx MXene at low NO2 concentrations (4–8 ppm), the response times increase to 93–172, 55–124 and 45–68 s for Z1T, Z3T and Z5T respectively. When 10–20 ppm NO2 is detected, the response time decreases significantly to ~22, 25 and 35 s, respectively. Thus, when ZnO is modified with Ti2CTx MXene, there is an increase in response time (compared to single ZnO) when detecting low concentrations and a decrease when detecting higher concentrations.
Since zinc oxide is the main gas sensitive material in ZnO/Ti2CTx nanocomposites, the detection mechanism will be determined to a greater extent by the behaviour of ZnO. To describe the NO2 detection mechanism, the generally accepted model related to the presence of ion-adsorbed oxygen (O) and electron depletion layer (EDL) on the ZnO surface (n-type semiconductor) should be used [79]. Sensitivity of single ZnO high response in the detection to NO2 by single ZnO is related to the reactions that take place on the surface of the material. Of all the metal oxide materials used as receptors in MOS sensors, ZnO probably has the strongest basic properties [80]. In the work of Mei Chen et al. [81] and Liu et al. [82], as in our previous study [18], it was shown that the highest response to NO2 is observed in the operating temperature range of 125–200 °C. Due to the strong basic properties of ZnO, the NO2 influx leads to the formation of nitrite (NO2), and then nitrate (NO3) in the first few minutes, involving various defects in the ZnO crystal lattice and ion-adsorbed oxygen. The formation of nitro-groups occurs with the participation of electrons from the conduction band of ZnO, which explains the high chemoresistance response. The modification of zinc oxide by the accordion-like Ti2CTx MXene leads to a consistent increase in sensitivity to NO2, which can be associated primarily with a marked increase in the specific surface area of the receptor material, leading to an increase in the number of adsorption centers and the diffusion rate of gases in the volume of the receptor layer. In addition, the presence of polar functional groups in the MXene structure can further contribute to the adsorption of gases capable of forming hydrogen bonds. The formation of heterojunctions at the interface between the components should lead to charge separation and increase the sensitivity to electron donor or acceptor gases [59,62,63,82,83,84].
Table 1 compares the gas sensing characteristics of ZnO/Ti2CTx and ZnO/Ti3C2Tx nanocomposites presented in the literature. As can be seen, ZnO/Ti3C2Tx nanocomposites are the main ones described in the literature. It was found that the vast majority of materials exhibit an increased response to NO2, which is also typical of individual zinc oxide. Nevertheless, the response to NO2 of ZnO/Ti2CTx nanocomposites obtained in this study is significantly higher than that of their analogues. The Sel parameter, which characterizes selectivity towards this gaseous analyte, was also estimated to be higher for ZnO/Ti2CTx nanocomposites obtained using the solvothermal method as compared to the analogues described in the literature.

4. Conclusions

Using atmospheric pressure solvothermal synthesis, composite powders in which nanodispersed zinc oxide was doped with 1–5 mol% of the accordion-like Ti2CTx MXene were synthesized, and their phase composition and microstructure were studied. It was found that the MXene powder addition in the synthesis process resulted in the formation of ZnO aggregates with a more porous structure compared to individual ZnO. This is probably due to the fact that the growth of ZnO particle nucleation on the porous MXene surface limited the consolidation.
The study of the chemoresistive properties of the obtained samples at the detection temperature of 200 °C for the most practically required analyte gases (explosive, toxic inorganic compounds and VOCs) revealed an extremely high sensitivity of all materials to NO2. At the same time, the selectivity increased with the increase of the MXene component: the ratio of the response value at 4 ppm NO2 to the second highest response value (at 100 ppm acetone/ethanol) increased from 1.6 (ZnO) to 20.5 (sample Z5T). The response time for the detection of 4–20 ppm NO2 for single ZnO was 30–45 s. When ZnO is modified with Ti2CTx MXene, there is an increase in the response time for the detection of low concentrations (up to 45–124 s, 4–8 ppm NO2) and a decrease for the detection of higher concentrations (up to 22–35 s). In addition to the high selectivity with increasing MXene content in the composite powder, the sensitivity to NO2 increased over the whole range of concentrations studied (4–20 ppm). For example, the response value increased by 4 ppm from 5 for ZnO to 26.7 for the sample containing 5 mol% Ti2CTx. This can be attributed to the increase in the specific surface area of the material, which causes an increase in the number of adsorption centers and the rate of analyte diffusion, as well as to the presence of polar functional groups on the MXene surface and the formation of a Schottky barrier at the interface.

Author Contributions

Conceptualization, E.P.S., N.P.S., A.S.M. and N.T.K.; methodology, E.P.S. and N.T.K.; validation, E.P.S., A.S.M. and N.P.S.; investigation, I.A.N., A.S.M., A.A.A., Y.M.G., T.L.S. and N.P.S.; resources, N.T.K.; writing—original draft preparation, E.P.S., A.S.M. and I.A.N.; writing—review and editing, I.A.N., A.A.A., Y.M.G. and T.L.S.; supervision, N.T.K.; project administration, N.P.S.; 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/.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Dhall, S.; Mehta, B.R.; Tyagi, A.K.; Sood, K. A review on environmental gas sensors: Materials and technologies. Sens. Int. 2021, 2, 100116. [Google Scholar] [CrossRef]
  2. Wilson, A. Advances in Electronic-Nose Technologies for the Detection of Volatile Biomarker Metabolites in the Human Breath. Metabolites 2015, 5, 140–163. [Google Scholar] [CrossRef] [PubMed]
  3. Tai, H.; Wang, S.; Duan, Z.; Jiang, Y. Evolution of breath analysis based on humidity and gas sensors: Potential and challenges. Sens. Actuators B Chem. 2020, 318, 128104. [Google Scholar] [CrossRef]
  4. Solovieva, S.; Karnaukh, M.; Panchuk, V.; Andreev, E.; Kartsova, L.; Bessonova, E.; Legin, A.; Wang, P.; Wan, H.; Jahatspanian, I.; et al. Potentiometric multisensor system as a possible simple tool for non-invasive prostate cancer diagnostics through urine analysis. Sens. Actuators B Chem. 2019, 289, 42–47. [Google Scholar] [CrossRef]
  5. Deisingh, A.K.; Stone, D.C.; Thompson, M. Applications of electronic noses and tongues in food analysis. Int. J. Food Sci. Technol. 2004, 39, 587–604. [Google Scholar] [CrossRef]
  6. Devabharathi, N.; Umarji, A.M.; Dasgupta, S. Fully Inkjet-Printed Mesoporous SnO2-Based Ultrasensitive Gas Sensors for Trace Amount NO2 Detection. ACS Appl. Mater. Interfaces 2020, 12, 57207–57217. [Google Scholar] [CrossRef]
  7. Thai, N.X.; Tonezzer, M.; Masera, L.; Nguyen, H.; Van Duy, N.; Hoa, N.D. Multi gas sensors using one nanomaterial, temperature gradient, and machine learning algorithms for discrimination of gases and their concentration. Anal. Chim. Acta 2020, 1124, 85–93. [Google Scholar] [CrossRef]
  8. Fedorov, F.S.; Simonenko, N.P.; Trouillet, V.; Volkov, I.A.; Plugin, I.A.; Rupasov, D.P.; Mokrushin, A.S.; Nagornov, I.A.; Simonenko, T.L.; Vlasov, I.S.; et al. Microplotter-Printed On-Chip Combinatorial Library of Ink-Derived Multiple Metal Oxides as an “Electronic Olfaction” Unit. ACS Appl. Mater. Interfaces 2020, 12, 56135–56150. [Google Scholar] [CrossRef]
  9. Goikhman, B.V.; Fedorov, F.S.; Simonenko, N.P.; Simonenko, T.L.; Fisenko, N.A.; Dubinina, T.S.; Ovchinnikov, G.; Lantsberg, A.V.; Lipatov, A.; Simonenko, E.P.; et al. Quantum of selectivity testing: Detection of isomers and close homologs using an AZO based e-nose without a prior training. J. Mater. Chem. A 2022, 10, 8413–8423. [Google Scholar] [CrossRef]
  10. Jeong, S.Y.; Kim, J.S.; Lee, J.H. Rational Design of Semiconductor-Based Chemiresistors and their Libraries for Next-Generation Artificial Olfaction. Adv. Mater. 2020, 32, 2002075. [Google Scholar] [CrossRef]
  11. Mokrushin, A.S.; Nagornov, I.A.; Simonenko, T.L.; Simonenko, N.P.; Gorobtsov, P.Y.; Arkhipushkin, I.A.; Simonenko, E.P.; Sevastyanov, V.G.; Kuznetsov, N.T. Gas-sensitive nanostructured ZnO films praseodymium and europium doped: Electrical conductivity, selectivity, influence of UV irradiation and humidity. Appl. Surf. Sci. 2022, 589, 152974. [Google Scholar] [CrossRef]
  12. Vorobyeva, N.; Rumyantseva, M.; Filatova, D.; Spiridonov, F.; Zaytsev, V.; Zaytseva, A.; Gaskov, A. Highly Sensitive ZnO(Ga, In) for Sub-ppm Level NO2 Detection: Effect of Indium Content. Chemosensors 2017, 5, 18. [Google Scholar] [CrossRef] [Green Version]
  13. Nagornov, I.A.; Mokrushin, A.S.; Simonenko, E.P.; Simonenko, N.P.; Gorobtsov, P.Y.; Sevastyanov, V.G.; Kuznetsov, N.T. Zinc oxide obtained by the solvothermal method with high sensitivity and selectivity to nitrogen dioxide. Ceram. Int. 2020, 46, 7756–7766. [Google Scholar] [CrossRef]
  14. Rumyantseva, M.N.; Vladimirova, S.A.; Platonov, V.B.; Chizhov, A.S.; Batuk, M.; Hadermann, J.; Khmelevsky, N.O.; Gaskov, A.M. Sub-ppm H2S sensing by tubular ZnO-Co3O4 nanofibers. Sens. Actuators B Chem. 2020, 307, 127624. [Google Scholar] [CrossRef]
  15. Zhou, X.; Lin, X.; Yang, S.; Zhu, S.; Chen, X.; Dong, B.; Bai, X.; Wen, X.; Geyu, L.; Song, H. Highly dispersed Metal–Organic-Framework-Derived Pt nanoparticles on three-dimensional macroporous ZnO for trace-level H2S sensing. Sens. Actuators B Chem. 2020, 309, 127802. [Google Scholar] [CrossRef]
  16. Li, Z.; Yi, J. Drastically Enhanced Ammonia Sensing of Pt/ZnO Ordered Porous Ultra-Thin Films. Sens. Actuators B Chem. 2020, 317, 128217. [Google Scholar] [CrossRef]
  17. Shingange, K.; Tshabalala, Z.P.; Ntwaeaborwa, O.M.; Motaung, D.E.; Mhlongo, G.H. Highly selective NH3 gas sensor based on Au loaded ZnO nanostructures prepared using microwave-assisted method. J. Colloid Interface Sci. 2016, 479, 127–138. [Google Scholar] [CrossRef]
  18. Mokrushin, A.S.; Nagornov, I.A.; Simonenko, T.L.; Simonenko, N.P.; Gorobtsov, P.Y.; 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]
  19. Mokrushin, A.S.; Nagornov, I.A.; Gorban, Y.M.; Simonenko, T.L.; Simonenko, N.P.; Arkhipushkin, I.A.; Simonenko, E.P.; Kuznetsov, N.T. Effect of platinum nanoparticles on the chemoresistive gas sensitive properties of the ZnO/Pt composite. Ceram. Int. 2023. [Google Scholar] [CrossRef]
  20. Du, Y.; Gao, S.; Mao, Z.; Zhang, C.; Zhao, Q.; Zhang, S. Aerobic and anaerobic H2 sensing sensors fabricated by diffusion membranes depositing on Pt-ZnO film. Sens. Actuators B Chem. 2017, 252, 239–250. [Google Scholar] [CrossRef]
  21. Hassan, K.; Uddin, A.S.M.I.; Ullah, F.; Kim, Y.S.; Chung, G.-S. Platinum/palladium bimetallic ultra-thin film decorated on a one-dimensional ZnO nanorods array for use as fast response flexible hydrogen sensor. Mater. Lett. 2016, 176, 232–236. [Google Scholar] [CrossRef]
  22. Li, W.; Ma, S.; Li, Y.; Yang, G.; Mao, Y.; Luo, J.; Gengzang, D.; Xu, X.; Yan, S. Enhanced ethanol sensing performance of hollow ZnO–SnO2 core–shell nanofibers. Sens. Actuators B Chem. 2015, 211, 392–402. [Google Scholar] [CrossRef]
  23. Han, C.; Li, X.; Shao, C.; Li, X.; Ma, J.; Zhang, X.; Liu, Y. Composition-controllable p-CuO/n-ZnO hollow nanofibers for high-performance H2S detection. Sens. Actuators B Chem. 2019, 285, 495–503. [Google Scholar] [CrossRef]
  24. Li, H.; Chu, S.; Ma, Q.; Fang, Y.; Wang, J.; Che, Q.; Wang, G.; Yang, P. Novel Construction of Morphology-Tunable C–N/SnO2/ZnO/Au Microspheres with Ultrasensitivity and High Selectivity for Triethylamine under Various Temperature Detections. ACS Appl. Mater. Interfaces 2019, 11, 8601–8611. [Google Scholar] [CrossRef]
  25. Lee, J.H.; Kim, J.Y.; Kim, J.H.; Mirzaei, A.; Kim, H.W.; Kim, S.S. Co3O4-loaded ZnO nanofibers for excellent hydrogen sensing. Int. J. Hydrog. Energy 2019, 44, 27499–27510. [Google Scholar] [CrossRef]
  26. Zhou, Q.; Hong, C.; Li, Z.; Peng, S.; Wu, G.; Wang, Q.; Zhang, Q.; Xu, L. Facile Hydrothermal Synthesis and Enhanced Methane Sensing Properties of Pt-Decorated ZnO Nanosheets. J. Nanosci. Nanotechnol. 2018, 18, 3335–3340. [Google Scholar] [CrossRef]
  27. Gong, Y.; Wu, X.; Chen, J.; Li, W.; Han, N.; Zhang, D.; Chen, Y. Enhanced gas-sensing performance of metal@ZnO core–shell nanoparticles towards ppb–ppm level benzene: The role of metal–ZnO hetero-interfaces. New J. Chem. 2019, 43, 2220–2230. [Google Scholar] [CrossRef]
  28. Wang, H.T.; Kang, B.S.; Ren, F.; Tien, L.C.; Sadik, P.W.; Norton, D.P.; Pearton, S.J.; Lin, J. Hydrogen-selective sensing at room temperature with ZnO nanorods. Appl. Phys. Lett. 2005, 86, 243503. [Google Scholar] [CrossRef]
  29. Van Tran, T.; Ahemad, M.J.; Kim, D.-S.; Le, T.D.; Dao, V.; Yu, Y.-T. Ultra high response for hydrogen gas on Pd–Augr-alloy@ZnO core-shell nanoparticles with Pd–Au gradient composition alloy core. Mater. Today Nano 2023, 21, 100292. [Google Scholar] [CrossRef]
  30. Anjitha, R.G.; Basu, P.K. Dual Activation and Photocatalyst-Induced Selective Detection in Defect Tuned Pd–ZnO Thin Films for Low ppm Detection of CH and CO. IEEE Sens. J. 2023, 23, 143–153. [Google Scholar] [CrossRef]
  31. Schütt, F.; Postica, V.; Adelung, R.; Lupan, O. Single and Networked ZnO–CNT Hybrid Tetrapods for Selective Room-Temperature High-Performance Ammonia Sensors. ACS Appl. Mater. Interfaces 2017, 9, 23107–23118. [Google Scholar] [CrossRef]
  32. Sinha, M.; Neogi, S.; Mahapatra, R.; Krishnamurthy, S.; Ghosh, R. Material dependent and temperature driven adsorption switching (p- to n- type) using CNT/ZnO composite-based chemiresistive methanol gas sensor. Sens. Actuators B Chem. 2021, 336, 129729. [Google Scholar] [CrossRef]
  33. Singh, G.; Choudhary, A.; Haranath, D.; Joshi, A.G.; Singh, N.; Singh, S.; Pasricha, R. ZnO decorated luminescent graphene as a potential gas sensor at room temperature. Carbon N. Y. 2012, 50, 385–394. [Google Scholar] [CrossRef]
  34. Anand, K.; Singh, O.; Singh, M.P.; Kaur, J.; Singh, R.C. Hydrogen sensor based on graphene/ZnO nanocomposite. Sens. Actuators B Chem. 2014, 195, 409–415. [Google Scholar] [CrossRef]
  35. Hussein, H.T.; Kareem, M.H.; Abdul Hussein, A.M. Synthesis and characterization of carbon nanotube doped with zinc oxide nanoparticles CNTs-ZnO/PS as ethanol gas sensor. Optik 2021, 248, 168107. [Google Scholar] [CrossRef]
  36. Simonenko, N.P.; Glukhova, O.E.; Plugin, I.A.; Kolosov, D.A.; Nagornov, I.A.; Simonenko, T.L.; Varezhnikov, A.S.; Simonenko, E.P.; Sysoev, V.V.; Kuznetsov, N.T. The Ti0.2V1.8C MXene Ink-Prepared Chemiresistor: From Theory to Tests with Humidity versus VOCs. Chemosensors 2022, 11, 7. [Google Scholar] [CrossRef]
  37. 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, 1850–1859. [Google Scholar] [CrossRef]
  38. Mokrushin, A.S.; Nagornov, I.A.; Averin, A.A.; Simonenko, T.L.; Simonenko, N.P.; Simonenko, E.P.; Kuznetsov, N.T. Chemoresistive Properties of V2CTx MXene and the V2CTx/V3O7 Nanocomposite Based on It. Chemosensors 2023, 11, 142. [Google Scholar] [CrossRef]
  39. Kedambaimoole, V.; Harsh, K.; Rajanna, K.; Sen, P.; Nayak, M.M.; Kumar, S. MXene wearables: Properties, fabrication strategies, sensing mechanism and applications. Mater. Adv. 2022, 3, 3784–3808. [Google Scholar] [CrossRef]
  40. 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 2022, 11, 13. [Google Scholar] [CrossRef]
  41. Peng, B.; Huang, X. Research status of gas sensing performance of Ti3C2Tx-based gas sensors: A mini review. Front. Chem. 2022, 10, 1037732. [Google Scholar] [CrossRef] [PubMed]
  42. Pazniak, H.; Varezhnikov, A.S.; Kolosov, D.A.; Plugin, I.A.; Di Vito, A.; Glukhova, O.E.; Sheverdyaeva, P.M.; Spasova, M.; Kaikov, I.; Kolesnikov, E.A.; et al. 2D Molybdenum Carbide MXenes for Enhanced Selective Detection of Humidity in Air. Adv. Mater. 2021, 33, 2104878. [Google Scholar] [CrossRef] [PubMed]
  43. 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] [PubMed]
  44. Rasheed, T. MXenes as emerging two-dimensional analytical modalities for potential recognition of hazardous environmental contaminants. Mater. Today Chem. 2022, 24, 100859. [Google Scholar] [CrossRef]
  45. Alwarappan, S.; Nesakumar, N.; Sun, D.; Hu, T.Y.; Li, C.-Z. 2D metal carbides and nitrides (MXenes) for sensors and biosensors. Biosens. Bioelectron. 2022, 205, 113943. [Google Scholar] [CrossRef]
  46. Nahirniak, S.; Saruhan, B. MXene Heterostructures as Perspective Materials for Gas Sensing Applications. Sensors 2022, 22, 972. [Google Scholar] [CrossRef]
  47. Simonenko, E.P.; Simonenko, N.P.; Mokrushin, A.S.; Simonenko, T.L.; Gorobtsov, P.Y.; Nagornov, I.A.; Korotcenkov, G.; Sysoev, V.V.; Kuznetsov, N.T. Application of Titanium Carbide MXenes in Chemiresistive Gas Sensors. Nanomaterials 2023, 13, 850. [Google Scholar] [CrossRef]
  48. 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]
  49. 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]
  50. Li, X.; An, Z.; Lu, Y.; Shan, J.; Xing, H.; Liu, G.; Shi, Z.; He, Y.; Chen, Q.; Han, R.P.S.; et al. Room Temperature VOCs Sensing with Termination-Modified Ti3C2Tx MXene for Wearable Exhaled Breath Monitoring. Adv. Mater. Technol. 2022, 7, 2100872. [Google Scholar] [CrossRef]
  51. 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]
  52. 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]
  53. Yao, L.; Tian, X.; Cui, X.; Zhao, R.; Xiao, M.; Wang, B.; Xiao, X.; Wang, Y. Two-dimensional Ti3C2Tx MXene/SnO nanocomposites: Towards enhanced response and selective ammonia vapor sensor at room temperature. Sens. Actuators B Chem. 2022, 358, 131501. [Google Scholar] [CrossRef]
  54. Bag, A.; Lee, N. Recent Advancements in Development of Wearable Gas Sensors. Adv. Mater. Technol. 2021, 6, 2000883. [Google Scholar] [CrossRef]
  55. 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]
  56. 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]
  57. Lee, J.; Kang, Y.C.; Koo, C.M.; Kim, S.J. Ti3C2Tx MXene Nanolaminates with Ionic Additives for Enhanced Gas-Sensing Performance. ACS Appl. Nano Mater. 2022, 5, 11997–12005. [Google Scholar] [CrossRef]
  58. Wang, J.; Yang, Y.; Xia, Y. Mesoporous MXene/ZnO nanorod hybrids of high surface area for UV-activated NO2 gas sensing in ppb-level. Sens. Actuators B Chem. 2022, 353, 131087. [Google Scholar] [CrossRef]
  59. Liu, X.; Zhang, H.; Song, Y.; Shen, T.; Sun, J. Facile solvothermal synthesis of ZnO/Ti3C2Tx MXene nanocomposites for NO2 detection at low working temperature. Sens. Actuators B Chem. 2022, 367, 132025. [Google Scholar] [CrossRef]
  60. Zhu, Y.; Ma, Y.; Wu, D.; Jiang, G. Preparation and gas sensing properties of ZnO/MXene composite nanomaterials. Sens. Actuators A Phys. 2022, 344, 113740. [Google Scholar] [CrossRef]
  61. Bu, X.; Wu, Q.; Yuan, Y.; Wu, H.; Liu, W.; Li, X.; Han, C. The highly sensitive ethanol sensor based on nanocomposites of ZnO in situ grown on 2D Ti3C2Tx nanosheet. Nanotechnology 2023, 34, 105707. [Google Scholar] [CrossRef] [PubMed]
  62. Yang, Z.; Jiang, L.; Wang, J.; Liu, F.; He, J.; Liu, A.; Lv, S.; You, R.; Yan, X.; Sun, P.; et al. Flexible resistive NO2 gas sensor of three-dimensional crumpled MXene Ti3C2Tx/ZnO spheres for room temperature application. Sens. Actuators B Chem. 2021, 326, 128828. [Google Scholar] [CrossRef]
  63. Fan, C.; Shi, J.; Zhang, Y.; Quan, W.; Chen, X.; Yang, J.; Zeng, M.; Zhou, Z.; Su, Y.; Wei, H.; et al. Fast and recoverable NO2 detection achieved by assembling ZnO on Ti3C2Tx MXene nanosheets under UV illumination at room temperature. Nanoscale 2022, 14, 3441–3451. [Google Scholar] [CrossRef] [PubMed]
  64. Naguib, M.; Kurtoglu, M.; Presser, V.; Lu, J.; Niu, J.; Heon, M.; Hultman, L.; Gogotsi, Y.; Barsoum, M.W. Two-Dimensional Nanocrystals Produced by Exfoliation of Ti3AlC2. Adv. Mater. 2011, 23, 4248–4253. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  65. 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]
  66. Nagornov, I.A.; Mokrushin, A.S.; Simonenko, E.P.; Simonenko, N.P.; Sevastyanov, V.G.; Kuznetsov, N.T. Liquid-Phase Growth of Nanocrystalline ZnO Thin Films and Their Gas-Sensitive Properties. Russ. J. Inorg. Chem. 2022, 67, 539–546. [Google Scholar] [CrossRef]
  67. Mokrushin, A.S.; Nagornov, I.A.; Averin, A.A.; Simonenko, N.P.; Simonenko, T.L.; Simonenko, E.P.; Sevastyanov, V.G.; Kuznetsov, N.T. Effect of the Addition of Cerium Acetylacetonate on the Synthesis of ZnO Nanopowder. Russ. J. Inorg. Chem. 2021, 66, 638–644. [Google Scholar] [CrossRef]
  68. Simonenko, N.P.; Fisenko, N.A.; Fedorov, F.S.; Simonenko, T.L.; Mokrushin, A.S.; Simonenko, E.P.; Korotcenkov, G.; Sysoev, V.V.; Sevastyanov, V.G.; Kuznetsov, N.T. Printing Technologies as an Emerging Approach in Gas Sensors: Survey of Literature. Sensors 2022, 22, 3473. [Google Scholar] [CrossRef]
  69. Simonenko, T.L.; Simonenko, N.P.; Simonenko, E.P.; Vlasov, I.S.; Volkov, I.A.; Kuznetsov, N.T. Microplotter Printing of Hierarchically Organized Planar NiCo2O4 Nanostructures. Russ. J. Inorg. Chem. 2022, 67, 1848–1854. [Google Scholar] [CrossRef]
  70. Simonenko, T.L.; Simonenko, N.P.; Gorobtsov, P.Y.; Simonenko, E.P.; Kuznetsov, N.T. Microextrusion Printing of Multilayer Hierarchically Organized Planar Nanostructures Based on NiO, (CeO2)0.8(Sm2O3)0.2 and La0.6Sr0.4Co0.2Fe0.8O3−δ. Micromachines 2022, 14, 3. [Google Scholar] [CrossRef]
  71. 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]
  72. Güell, F.; Martínez-Alanis, P.R.; Khachadorian, S.; Rubio-García, J.; Franke, A.; Hoffmann, A.; Santana, G. Raman and photoluminescence properties of ZnO nanowires grown by a catalyst-free vapor-transport process using ZnO nanoparticle seeds. Phys. Status Solidi 2016, 253, 883–888. [Google Scholar] [CrossRef]
  73. Šćepanović, M.; Grujić-Brojčin, M.; Vojisavljević, K.; Bernik, S.; Srećković, T. Raman study of structural disorder in ZnO nanopowders. J. Raman Spectrosc. 2010, 41, 914–921. [Google Scholar] [CrossRef]
  74. 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]
  75. 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]
  76. 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]
  77. Nagornov, I.A.; Maeder, T.; Kuznetsov, N.T.; Mokrushin, A.S.; Simonenko, N.P.; Vasiliev, A.A.; Volkov, I.A.; Vlasov, I.S.; Gorobtsov, P.Y.; Simonenko, E.P.; et al. Nanocrystalline ZnO Obtained by the Thermal Decomposition of [Zn(H2O)(O2C5H7)2] in 1-Butanol: Synthesis and Testing as a Sensing Material. Russ. J. Inorg. Chem. 2018, 63, 1519–1528. [Google Scholar] [CrossRef]
  78. NOISH. Documentation for Immediately Dangerous to Life or Health Concentrations (IDLHs). 1996. Available online: https://www.cdc.gov/niosh/idlh/pdfs/1994-IDLH-ValuesBackgroundDocs.pdf (accessed on 15 March 2023).
  79. Kim, H.-J.; Lee, J.-H. Highly sensitive and selective gas sensors using p-type oxide semiconductors: Overview. Sens. Actuators B Chem. 2014, 192, 607–627. [Google Scholar] [CrossRef]
  80. Marikutsa, A.; Rumyantseva, M.; Konstantinova, E.A.; Gaskov, A. The Key Role of Active Sites in the Development of Selective Metal Oxide Sensor Materials. Sensors 2021, 21, 2554. [Google Scholar] [CrossRef]
  81. Chen, M.; Wang, Z.; Han, D.; Gu, F.; Guo, G. Porous ZnO Polygonal Nanoflakes: Synthesis, Use in High-Sensitivity NO2 Gas Sensor, and Proposed Mechanism of Gas Sensing. J. Phys. Chem. C 2011, 115, 12763–12773. [Google Scholar] [CrossRef]
  82. 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]
  83. Choi, J.; Kim, Y.; Cho, S.; Park, K.; Kang, H.; Kim, S.J.; Jung, H. 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]
  84. 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 Ti3C2T for acetone gas sensor working at room temperature. J. Mater. Sci. Technol. 2021, 73, 128–138. [Google Scholar] [CrossRef]
Figure 1. The X-ray powder patterns of the original Ti2AlC MAX-phase and the obtained Ti2CTx MXene (a) and the Ti2CTx MXene microstructure particles according to TEM (bd) and SEM (eg).
Figure 1. The X-ray powder patterns of the original Ti2AlC MAX-phase and the obtained Ti2CTx MXene (a) and the Ti2CTx MXene microstructure particles according to TEM (bd) and SEM (eg).
Micromachines 14 00725 g001
Figure 2. DSC (blue) and TGA (green) curves of Ti2CTx MXene obtained by thermal analysis in air flow (a) and argon (b).
Figure 2. DSC (blue) and TGA (green) curves of Ti2CTx MXene obtained by thermal analysis in air flow (a) and argon (b).
Micromachines 14 00725 g002
Figure 3. TGA curve after heating the resulting accordion-like Ti2CTx MXene powder in an argon flow to the temperature 250 °C and holding for 30 min (a), and also sections of the X-ray diagram in the interval 2θ = 6–10° showing the reflex position (002) (b) in the initial powder (1), after heating in an argon flow to 250 °C for 30 min (2) and after holding the sample in a humid atmosphere for 2 days (3).
Figure 3. TGA curve after heating the resulting accordion-like Ti2CTx MXene powder in an argon flow to the temperature 250 °C and holding for 30 min (a), and also sections of the X-ray diagram in the interval 2θ = 6–10° showing the reflex position (002) (b) in the initial powder (1), after heating in an argon flow to 250 °C for 30 min (2) and after holding the sample in a humid atmosphere for 2 days (3).
Micromachines 14 00725 g003
Figure 4. X-ray diffraction patterns of the obtained powders Z, Z1T, Z3T and Z5T (a) and Raman spectra of Ti2CTx MXene powders and Z and Z5T samples (b).
Figure 4. X-ray diffraction patterns of the obtained powders Z, Z1T, Z3T and Z5T (a) and Raman spectra of Ti2CTx MXene powders and Z and Z5T samples (b).
Micromachines 14 00725 g004
Figure 5. Particle morphology of ZnO/Ti2CTx (Z5T) composite powder by TEM.
Figure 5. Particle morphology of ZnO/Ti2CTx (Z5T) composite powder by TEM.
Micromachines 14 00725 g005
Figure 6. Microstructure of the receptor layers of samples Z (ac), Z1T (df), Z3T (gi) and Z5T (jl) deposited on a sensor substrate by microplotter printing, as determined by SEM.
Figure 6. Microstructure of the receptor layers of samples Z (ac), Z1T (df), Z3T (gi) and Z5T (jl) deposited on a sensor substrate by microplotter printing, as determined by SEM.
Micromachines 14 00725 g006
Figure 7. Distribution of elements O, Zn and Ti obtained by EDX mapping for coating samples Z, Z1T, Z3T and Z5T.
Figure 7. Distribution of elements O, Zn and Ti obtained by EDX mapping for coating samples Z, Z1T, Z3T and Z5T.
Micromachines 14 00725 g007
Figure 8. Selectivity histogram constructed from the responses to different gases (H2, CH4, CO, C6H6, NH3, ethanol, acetone and NO2) (a) and the dependence of the selectivity coefficient (Sel) of the receptor layers of samples Z, Z1T, Z3T and Z5T (b); detection temperature 200 °C.
Figure 8. Selectivity histogram constructed from the responses to different gases (H2, CH4, CO, C6H6, NH3, ethanol, acetone and NO2) (a) and the dependence of the selectivity coefficient (Sel) of the receptor layers of samples Z, Z1T, Z3T and Z5T (b); detection temperature 200 °C.
Micromachines 14 00725 g008
Figure 9. Sensitivity to 4–20 ppm NO2 (a) and response as a function of concentration (b) of ZnO/Ti2CTx coatings at 200 °C detection temperature.
Figure 9. Sensitivity to 4–20 ppm NO2 (a) and response as a function of concentration (b) of ZnO/Ti2CTx coatings at 200 °C detection temperature.
Micromachines 14 00725 g009
Table 1. Comparison of gas sensing properties of ZnO/Ti2CTx and ZnO/Ti3C2Tx nanocomposites reported in the literature.
Table 1. Comparison of gas sensing properties of ZnO/Ti2CTx and ZnO/Ti3C2Tx nanocomposites reported in the literature.
YearCompositionGasConc., ppmTemp, °CResponseResponse Time (s)Sel 1Ref.
2021ZnO—Ti3C2TxNO2100RT41.39 234~8[62]
2022ZnO—Ti3C2TxNO220RT367.63 222~6[63]
2022ZnO—36.5% Ti3C2TxNO20.05RT, UV81 217~4[58]
2022ZnO—2% Ti3C2TxAcetone10032014.4 38~1.03[60]
2022ZnO—Ti3C2TxNO281603.6 3191~3.6[59]
2023ZnO—3.7% Ti3C2TxEthanol5030059 322~5[61]
2023ZnO—(1–5%)Ti2CTxNO220200>73.3 32220.5This work
1 Formula used for Sel: SRanalyte/SR2, where SR2 is the highest response among the gases tested in this article, excluding response to gas analyte. 2 Formula used for SR = (ΔR/Ra) × 100. 3 Formula used for SR = Ra/Rg.
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

Simonenko, E.P.; Nagornov, I.A.; Mokrushin, A.S.; Averin, A.A.; Gorban, Y.M.; Simonenko, T.L.; Simonenko, N.P.; Kuznetsov, N.T. Gas-Sensitive Properties of ZnO/Ti2CTx Nanocomposites. Micromachines 2023, 14, 725. https://doi.org/10.3390/mi14040725

AMA Style

Simonenko EP, Nagornov IA, Mokrushin AS, Averin AA, Gorban YM, Simonenko TL, Simonenko NP, Kuznetsov NT. Gas-Sensitive Properties of ZnO/Ti2CTx Nanocomposites. Micromachines. 2023; 14(4):725. https://doi.org/10.3390/mi14040725

Chicago/Turabian Style

Simonenko, Elizaveta P., Ilya A. Nagornov, Artem S. Mokrushin, Aleksey A. Averin, Yulia M. Gorban, Tatiana L. Simonenko, Nikolay P. Simonenko, and Nikolay T. Kuznetsov. 2023. "Gas-Sensitive Properties of ZnO/Ti2CTx Nanocomposites" Micromachines 14, no. 4: 725. https://doi.org/10.3390/mi14040725

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