Next Article in Journal
Monitoring of Chlorophylls during the Maturation Stage of Plums by Multivariate Calibration of RGB Data from Digital Images
Previous Article in Journal
Li6BaLa2Ta2O12 Solid-State Probe for Studying Li Activity in Molten Sn-Li Alloys
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

The Ti0.2V1.8C MXene Ink-Prepared Chemiresistor: From Theory to Tests with Humidity versus VOCs

by
Nikolay P. Simonenko
1,*,
Olga E. Glukhova
2,
Ilya A. Plugin
3,
Dmitry A. Kolosov
2,
Ilya A. Nagornov
1,
Tatiana L. Simonenko
1,
Alexey S. Varezhnikov
3,
Elizaveta P. Simonenko
1,
Victor V. Sysoev
3,* and
Nikolay T. Kuznetsov
1
1
Kurnakov Institute of General and Inorganic Chemistry of the Russian Academy of Sciences, 31 Leninsky pr., 119991 Moscow, Russia
2
Institute of Physics, Saratov State University, 410012 Saratov, Russia
3
Department of Physics, Yuri Gagarin State Technical University of Saratov, 77 Polytechnicheskaya str., 410054 Saratov, Russia
*
Authors to whom correspondence should be addressed.
Chemosensors 2023, 11(1), 7; https://doi.org/10.3390/chemosensors11010007
Submission received: 23 November 2022 / Revised: 14 December 2022 / Accepted: 20 December 2022 / Published: 22 December 2022
(This article belongs to the Section Materials for Chemical Sensing)

Abstract

:
The 2D structure of MXenes attracts wide research attention toward an application of these materials in gas sensors. These structures are extremely sensitive to minor variations in their composition, which are employed for tuning their functional properties. Here, we consider the partially substituted MXenes of the composition of TixV2-xC, where x = 0.2, via quantum chemical calculations, and test their chemiresistive characteristics as a receptor component of the planar-type sensor and on-chip multisensor array. We thoroughly discuss the synthesis process of Ti0.2V1.8AlC MAX-phase and the corresponding MXenes, to prepare functional inks and, furthermore, deposit the films by microextrusion printing over an array of planar multi-electrode structures at the surface of a pen-sized chip. The crystal structure of the obtained materials is evaluated via X-ray diffraction analysis. The developed chip has been exposed upon few gaseous analytes, of alcohol VOCs, NH3, and H2O, of a 500–16,000 ppm concentration, at room temperature to ensure that we could observe the positive chemiresistive effect matured from resistance enhancing, with up to 10% vs. water vapors. The calculations carried in the framework of the density-functional theory for V2C, Ti2C, and Ti0.2V1.8C crystals ensured that the variations in their electronic structure were almost consistent with the experiment fundings: the most prominent effect is observed in relation to the H2O vapors. Therefore, these Ti0.2V1.8C structures could be considered for applying them in room temperature-operated hygrometers.

Graphical Abstract

1. Introduction

Nowadays, metal oxides in the form of thick or thin films, which operate at elevated temperatures, ordinarily up to 400 °C, are widely used to design gas sensors of a chemiresistive type [1,2,3]. In order to reduce the working temperature and to advance the sensitivity, other materials are considered to meet the challenges of various current applications, including the Internet of Things [4]. Since the beginning of the XXI century, two-dimensional materials, primarily matured from a one atom-thick layer of carbon, called graphene [5], started to develop rapidly [1,6]. The lack of bulk in these structures allows for the electrical characteristics to entirely depend on gas adsorption and other environmental changes and, so, to develop sensors operating even at room temperature [7]. However, carbon structures have still much lower values of the chemiresistive response compared to oxide ones. Therefore, other two-dimensional materials are being intensively searched for an application in the sensors, like sulfides, selenides, nitrides, and carbides of transition metals [8,9,10,11,12,13,14]. The last two ones are called as MXenes [15]. The MXene structures are already shown to exhibit a suitable gas-sensing performance in forward to numerous gaseous analytes, including volatile organic compounds (VOCs) at room temperature [16,17,18].
The control of the humidity content in the atmosphere is essential for many processes and applications [19] that leads to intensive research of MXenes towards exposing the H2O vapors [20,21]. In order to advance the performance of such sensors, the researchers attempt to employ modified MXenes primarily via designing heterostructures. For instance, He et al. demonstrated [22] that adding SnO2 to Ti3C2Tx improves both the chemiresistive response of this heterostructure when compared to pure MXenes, and its stability. Alternatively, a transition metal substitution is another way to modify the chemiresistive characteristics of MXenes [23]. It was shown that the presence of vanadium in the MXenes structure is supposed to increase the sensor response to humid air [24,25,26], which was confirmed [24] by an example of the V2CTx MXenes-based sensor.
Still, the important aspect of the gas sensors is a selectivity to analytes. To date, all the available sensors of the chemiresistive type have a rather low selectivity. To advance this issue, sensors are assembled into sets or multisensor arrays whose vector signal depends on the analyte’s type in a higher degree following an adequate choice of sensor components [27,28]. For the ability to produce such arrays in a mass scale, they are frequently designed in an on-chip platform via forming multiple electrodes on the same substrate to be covered by a gas-sensitive layer (see, for instance, [29]). The array’s vector signals generated in the presence of analytes are processed via pattern recognition technologies to enable their identification, while the gas concentration could be extracted from a normal calibration of any of the sensor’s signal in the array [30].
It is worth mentioning that the surface-like 2D structure of MXenes allows one to perform the theoretical studying of the adsorption of the analytes from a gaseous phase. The major approach goes via applying quantum-chemical calculations in the framework of the density functional theory (DFT) method with a generalized gradient approximation (GGA) and Perdew–Berke–Ernzerhof (PBE) parameterization [31,32,33,34,35,36,37,38]. In the case of MXenes with a semiconductor conductivity type, researchers turn to employ Hubbard corrections [39,40,41] and hybrid functionals due to the presence of strongly correlated d-electrons in the transition metal elements [40,42,43]. This powerful method makes it possible to accurately interpret the experimental results and explain the mechanism of chemisorption in MXenes of a different chemical composition.
According to our survey of the current literature on the application of MXenes as the receptor components of chemiresistive gas sensors, the partially substituted V2C-based MXenes have not previously been considered for gas sensors and multisensor arrays. Moreover, the quantum-chemical calculations to detail the adsorption features of various analytes over such a material have not been performed. Thus, the aim of this work was studying (i) the synthesis process of TixV2-xC MXene structures, where the degree of vanadium substitution with titan, x, is equal to 0.2, (ii) the DFT calculation of the adsorption of various VOCs and H2O molecules over the Ti0.2V1.8C structure in comparison to Ti2C and V2C ones, (iii) the pilot measurements of the chemiresistive effect in Ti0.2V1.8C structures upon exposing to the VOCs and H2O at room temperature, and (iv) applying a pattern recognition technique to the vector signal generated by an on-chip Ti0.2V1.8C-based sensor array for a selective discrimination of the test analytes.

2. Materials and Methods

The scheme of the technology route to synthesize the Ti0.2V1.8C MXene and to study its chemosensory properties is drawn in Figure 1. The stages are described in detail in the following subsections.

2.1. Materials

Graphite (99%, MPG-8, Technocarb, Chelyabinsk, Russia), titanium (99%, Lanhit, Moscow, Russia), vanadium (99%, Lanhit, Moscow, Russia) and aluminum powders (99%, Lanhit, Moscow, Russia), KBr (99.99%, Rushim, Moscow, Russia), hydrofluoric acid (50%, Honeywell International Inc., Charlotte, USA), hydrochloric acid (36%, Sigma Tek, Moscow, Russia), ethylene glycol (99%, TK Spektr-Chim, Moscow, Russia), and tetramethylammonium hydroxide N(CH3)4(OH) (water solution, Technic, Saint-Denis, France) of an analytical grade were used in this work without a further purification.

2.2. Synthesis of the Ti0.2V1.8AlC MAX-Phase

At the first stage, the Ti0.2V1.8AlC MAX-phase synthesis was carried out according to the earlier reported protocols [44,45]. The choice of an aluminum-containing MAX-phase, where the Al atoms are located between the (Ti,V)2C layers, is based on their higher reactivity upon an interaction with HF for a subsequent MXene synthesis, when compared, for example, with Si-containing MAX-phases. For this purpose, the metallic powders of Ti, V, Al, and graphite, of a 99% purity, were mixed in the molar ratios of 0.2:1.8:1.2:0.8 following a number of preliminary experiments with the purpose of minimizing the forming of the by-products (Figure 1, pos.1). Then, KBr powder in the mass ratio of 1:1 was added to the obtained reaction mixture to be a subject of combined grinding in the planetary ball mill at ethanol under 100–500 rpm for 12 h to improve the homogenization of the material (Figure 1, pos.2). Next, the powder was dried under an advanced temperature of 70–80 °C to be formed into dense molds by pressing at 100 bar (Figure 1, pos.3). These samples were placed in ceramic crucibles covered with a protective layer of KBr powder and heat-treated in the muffle furnace at 1100 °C for 3 h (Figure 1, pos.4) following the inertial cooling down to room temperature (RT). After cooling, the samples were washed from the KBr residuals with a hot distilled H2O and dried at 110 °C until the completion of the weight loss (Figure 1, pos. 5).

2.3. Preparation of Ti0.2V1.8C MXene and Its Delamination

The MAX-phase powder (Figure 1, pos.6) was further used as a precursor to obtain the Ti0.2V1.8C MXene through a selective chemical Al etching. For this purpose, the Ti0.2V1.8AlC powder was introduced into a mixture of concentrated HF and HCl acids mixed in the volume ratio of 3:2, stirred at RT for 30 min, and then incubated at 40 °C for 90 h (Figure 1, pos.7). This resulted in a selective removal of the Al layers located between the M2C, where M = Ti, V, the layers and the formation of the Ti0.2V1.8C MXene multilayer agglomerates, whose surface is functionalized mainly by F- and O- ions, and OH-groups (Figure 1, pos.7). After the chemical etching, the MXene-contained suspension was repeatedly washed from by-products with the distilled H2O by a centrifugation at 3500 rpm until the hydrogen potential of the solution (pH) was equal to 5–6 at least.
The obtained multilayer agglomerates of Ti0.2V1.8C MXene (Figure 1, pos.8) were further delaminated into separate two-dimensional structures. The 12% aqueous solution of tetramethylammonium hydroxide, N(CH3)4(OH), in the ratio of 10 g/l was added to the corresponding powder, followed by ultrasonic treating for 20 min (Figure 1, pos.9). In order to isolate the single-layer nanosheets of Ti0.2V1.8C MXene and to eliminate the delamination agent, the resulting suspension was centrifuged at 3500 rpm for 60 min, the supernatant was removed, and the solid phase was washed twice with the distilled H2O (Figure 1, pos.10).

2.4. Fabrication of Ti0.2V1.8C MXene-Based Multielectrode Chip

In order to employ the MXenes for the sensor, the functional inks were prepared as a suspension by mixing the synthesized MXene structures with ethylene glycol. For this purpose, the aqueous MXene suspension was centrifuged at 15,000 rpm for 1 h, the solid phase was washed with an absolute ethanol, and then ethylene glycol was added. The resulting suspension was dispersed in an ultrasonic bath for 15 min. The output inks were applied to the dielectric substrate made of oxidized Si via a microextrusion printing at 10 µm spot resolution (Figure 1, pos.11). For the printing, we employed a home-made three-coordinate positioning system equipped with the pneumatic dozer and the capillary dispenser in the form of a hollow needle with an inner diameter of 150 µm. The chip substrate has been prior equipped with multiple co-planar Pt strip electrodes of a 50 µm width with an inter-electrode distance of 50 µm, meander-shaped heaters, and thermoresistors [46], all at the frontal side. The latter two ones were not employed in the course of these studies because of the operation at RT. The substrate containing the MXene-based layer was then dried at 150 °C under a reduced pressure, below 15 kPa, to avoid a spontaneous oxidation of the material. The MXene layer confined between each couple of strip electrodes has been served as a sensor element of the chemiresistor type. All the sensor elements located over the electrode set have been considered as a multisensor array.
Next, the chip with the deposited layer of the Ti0.2V1.8C MXene structures was wired (Figure 1, pos.12) into the ceramic holder [47] containing electric tracks to connect the measuring electrodes with the output multi-pins of an interface socket (SMC 1.27 mm, Erni Electronics, Adelberg, Germany).

2.5. Quantum-Chemical Calculations

An experimental identification of the chemisorption processes in MXenes is a challenging task. However, the theoretical study using a density functional theory (DFT) could provide an in-depth understanding of the mechanism of interaction of MXenes with gaseous analytes [48,49,50,51]. In our work, the first principle study of the MXene structures was performed using the Siesta 4.1.5 DFT code [52,53,54,55,56] under Grimme corrections, DFT-D2, in order to consider the Van der Waals interaction. We have employed the split valence basis set, including the polarization functions, of double-zeta plus polarization, with a generalized gradient approximation (GGA) and Perdew–Berke–Ernzerhoff (PBE) parameterization, since these calculation approaches have been proved to be appropriate in terms of the accuracy of the calculation and the calculation time. The 5 × 5 × 1 Monkhorst–Pack grid [54] is applied for integration into an inverse space. To minimize the total energy, the geometry relaxation of the atomic coordinates and translation vectors of the MXene structures were approached via considering the Hellman–Feynman forces under Bullet-type’s corrections in the framework of the modified Broyden’s method [55]. When optimizing the MXenes cells, the force action between atoms was chosen to be 0.04 eV/Å, while the 50 Å vacuum layer over the MXene surface was taken into consideration. The geometry relaxation resulted in atomistic models with the lattice vectors equal to 15.38 Å for Ti2C, 14.74 Å for V2C, and 14.68 Å for Ti0.2V1.8C. The real-space grid cutoff was 350 Ry. A calculation of the chemiresistive response of the MXene structure has been carried out within the Landauer–Büttiker formalism using non-equilibrium Green–Keldysh functions [56] according to:
G = 2 e h T ( E ) F T ( E E F ) d E
where e is the charge of the electron, h is Planck’s constant, E F is the Fermi energy of the contacts, and T(E) is the probability of the electron transmittance through the channel to be defined as:
T ( E ) = T r   [ Γ L ( E ) G ( E ) Γ R ( E ) G ( E ) ] .
Here, G ( E ) ,   G ( E )   are the lead and lag Green’s functions which describe a contact with the electrodes, and ΓL(E), ΓD(E) are the level broadening matrices defined as:
Γ L / R ( E ) = i ( Σ L / R ( E ) Σ L / R ( E ) )
where Σ L / R   are the eigen energy matrices of the left and right electrodes. The Green’s matrices are calculated according to:
G ( E ) = 1 ( E S C H C L ( E ) R ( E ) )
where S C is the matrix of the overlaps of the conducting channel atomic orbitals, E is the electron energy, and H C is the Hamiltonian of the conducting channel.
To calculate the chemiresistive response, the number of k-points, perpendicular to the current transfer, was chosen to be 100.

2.6. Instrumentation

X-ray diffraction (XRD) analysis of the materials under study, the MAX-phase powder, and the corresponding MXene film deposited on the glass substrate surface was performed using a D8-Advance diffractometer (Bruker, Billerica, MA, USA) equipped with a CuKα source, 1.5418 Å, and Ni filter. The measurements were taken under 40 keV of energy at a current of 40 mA. The integration time was equal to 0.3 s/point with a step of 0.02°.
The microstructure and the chemical composition of the obtained Ti0.2V1.8C MXene was analyzed with the help of a scanning electron microscope NVision 40 three-beam workstation (Carl Zeiss, Inc., Oberkochen, Germany) combined with an energy-dispersive X-ray (EDX) spectrometer of INCA X-MAX 80 (Oxford Instruments, Abingone, UK). The scanning electron microscopy (SEM) images were acquired at an accelerating voltage U = 10 kV (secondary electron detector) and EDX spectrum at U = 20 kV.
The gas-sensing measurements were carried out in a home-made gas-mixing setup composing of a few bubblers containing the liquid analytes of an analytical grade, namely methanol, ethanol, isopropanol, ammonia solution, and distilled water (Figure 2). The bubblers were purged with dry air sourced from the corresponding generator (Peak Scientific, Glasgow, UK) with the constant flowrate of 400 sccm. The multi-line gas flows were managed via high-precision mass-flow controllers (Bronkhorst, Achterhoek, the Netherlands) to be mixed in the specified ratio. The gas mixtures were delivered to the chip under study installed into a sealed steel-stainless chamber equipped with entry and output tubes. The concentration of the analytes was calculated via [46] as:
C = F g a s P s a t F g a s P s a t + ( P a t m P s a t ) F g a s + P a t m F a i r
where Fgas is the flow rate of gas through the bubbler (cm3/min), Fair is the flow rate of the background air (cm3/min), Psat is the saturated vapor pressure of the analyte solution (mm Hg), and Patm is the atmosphere pressure (mm Hg).
The resistance of the MXene layer located over the chip was measured via the corresponding multimeter (Keithley Instruments, Cleveland, OH, USA) assembled with National Instruments’ (Austin, TX, USA) multiple-entry NI-DAQ unit. All of the setup was driven by PC via Labview@ software (National Instruments, Austin, TX, USA).

3. Results and Discussion

3.1. Characterization of the MAX-Phase and MXene under Study

In order to study the MXene crystal’s structure, the functional ink was used to deposit the carbide film on the surface of a glass substrate. After drying the ink layer at 150 °C under a reduced pressure, the material has been found to have a crystalline structure, characteristic for MXene.
The XRD patterns characterizing the target Ti0.2V1.8AlC MAX-phase and the corresponding MXene of Ti0.2V1.8C are given in Figure 3. It could be seen that the obtained Ti0.2V1.8C MXene structure does not contain any of the crystalline impurities left after the initial components or byproducts.
The results of the EDX analysis showed the content of Ti and V atoms equal to 1.4 at.% and 13.2% at.%, correspondingly, which further confirms the given [Ti]/[V] ratio to be equal to approx. 0.11, and the appearance of MXene in the desired composition.
The microstructure of the Ti0.2V1.8C MXene layer placed over the surface of the glass substrate from the ink was studied by SEM (Figure 4). As can be seen from the electron microscopy images, the material appears as a homogeneous layer in a characteristic smooth morphology. Further, a detailed analysis of the carbide film’s structural features revealed that the material consists of an array made up of individual large-area thin nanosheets. Thus, these data indicate a high degree of delamination of the resulting Ti0.2V1.8C MXene at the stage of the preparation of the functional ink, as well as the retention of the two-dimensional structure of the corresponding disperse phase elements at the stage of the film’s formation on the substrate’s surface.

3.2. Quantum-Chemical Calculations

Three atomistic models of MXenes were considered for quantum-chemical calculations as Ti2C, V2C, and Ti0.2V1.8C MXene of the mixed composition at the 1T phase. Initially, we have employed the atomic configurations corresponding to the lowest possible supercells with lattice constants of 3.07 Å for Ti2C and 2.95 Å for V2C. Following the optimization of the atomic coordinates and lattice constants, the elementary structures were extended by translating the supercells. Thus, the atomistic models were obtained with lattice constants of 15.38 Å characterizing Ti2C and 14.74 Å characterizing V2C. By replacing the vanadium atoms in the V2C structure with titanium ones, we have derived the atomistic model of Ti0.2V1.8C with lattice vectors equal to 14.68 Å. Such a rather large area of the supercells was required to consider a possibility for the analyte molecules to land over the surface of the MXenes. Originally, the analyte molecule was located above the surface of MXene at the distance of about 3.2 Å, where the total energy minimum of the MXene-analyte system was identified accounting for the geometry relaxation using Grimme dispersion corrections. The atomistic models under study are shown in Figure 5.
Further, in order to reveal the resistance of the supercell and its change due to the MXene interaction with analyte molecules, we have estimated the electron charge density distribution over the supercell atoms according to the Mulliken procedure [57]. The obtained values of the Mulliken charge are given in Table 1. Here, the negative value of the charge indicates a charge transfer from the MXene film to the analyte. As can be seen from the table, the transferred charge does not exceed the magnitude of 0.11e, which characterizes the interaction of Ti0.2V1.8C with the ethanol molecule.
The observed low values of the charge appear because all the analyte molecules are bound by Van der Waals forces to the MXene surface, as indicated by the binding energies yielded in Table 2. As one can see, the binding energies do not exceed −0.128 eV, characteristic of an ethanol interaction with Ti2C. In the case of Ti0.2V1.8C MXene, the lowest values of the binding energy are observed in range of 0.044–0.062 eV depending on the analyte. It indicates the easier adsorption/desorption of molecules over this surface when compared to Ti2C and V2C where the energies are higher. It is worth noting that these energies match well the conditions of an RT operation of the Ti0.2V1.8C MXene-based sensors.
In order to evaluate the contribution of the analytes to the change in the electronic structure of MXene, we have estimated the electronic density of the states (DOS) accounting for setting the smearing parameter equal to zero. In this case, even a minor DOS change due to some charge overflow is clearly noticeable than one taking the conventional DOS pattern with the smearing of the states. Figure 6 yields the DOS patterns calculated for Ti0.2V1.8C MXene without smearing at a pristine state and following the adsorption of the analytes under interest (alcohols, ammonia, and water).
It is well known that the resistance of a material depends on the number of electronic states near the Fermi energy. The contribution of electronic states into the conductivity decreases with a displacement along the energy axis to be relative to the Fermi energy, E-Ef, with the major one to lie in the approximate energy band of E-Ef = [−0.5 eV, +0.5 eV].
It can be seen from Figure 6 that the analyte adsorption results in a shift of the DOS peak positions, and the modulation of their magnitude near the Fermi energy, especially in the conduction band. In case of pristine Ti0.2V1.8C MXene, the three DOS peaks are concentrated in the range of ~50 meV off the Fermi energy, and the amplitude of one of the maxima is equal to 1780 (eV)−1. The nearest peak is located in the conduction band, as a positive region of the energy axis, with the energy of ~25 meV and an amplitude of 1146 (eV)−1. In this regard, we observe the lowest resistance value of 4.118 kΩ. When the H2O molecule approaches the MXene surface, the peak at the energy of ~25 meV disappears and the amplitudes of the peaks closest to the Fermi energy are reduced down to values of about 750 (eV)−1. A similar picture is also observed for an ethanol interaction with the MXene. The nearest peak arises at 51 meV in energy with an amplitude of higher than 500 (eV)−1. An interesting picture is observed with the isopropanol molecule, whose interaction yields the nearest DOS to be at 17 meV with an amplitude of 1489 (eV)−1. This process should lead to a significant reduction in the electrical resistance of the crystal under study, but this peak is the only one in the energy range up to 0.24 eV which levels out its contribution to the electrical conductivity. All these features in the DOS’ behavior lie behind the change in the MXene resistance upon the analyte adsorption.
We have estimated the resistance of the Ti0.2V1.8C MXene in a pristine state and following the interaction with the analytes via the method of nonequilibrium Green–Keldysh functions. Primarily, the introduction of Ti atoms into the V2C crystal structure enhanced approx. two-fold the resistivity of the Ti0.2V1.8C structure. Figure 6b, the bottom plot, summarizes the resistance values derived for the Ti0.2V1.8C supercell in a pristine state and upon an analyte adsorption. Obviously, we cannot compare the exact values of the supercell resistance with the experimental data which relate to much larger structures. However, we could estimate the change in the MXene resistance upon the analyte’s appearance according to the definition of the conventional chemiresistive response as:
S = ( R ( a n a l y t e ) R ( p r i s t i n e ) 1 ) · 100 %
where R(analyte) and R(pristine) are the resistance of the MXene supercell with and without analytes on the surface. Figure 6b, the upper plot, shows the calculated S values which well correlate with the observed shifts in the DOS. For the pristine layer, the peak of the DOS shifts from the Fermi energy by approx. 25 meV, while the adsorption of the analytes leads to its larger shifting by approx. 50 meV. Such a shift yields an increase in the supercell’s resistance according to the above discussion regarding the regularities of the location of the DOS’ peaks. A similar pattern is observed for the chemiresistive response values in dependence on the test analytes.

3.3. Chemiresistive Characterization

Prior to measuring the gas response of the Ti0.2V1.8C MXene-based sensor elements, we have tested the I-V characteristics of the local layer confined between two measuring Pt electrodes at the chip. Figure 7a shows the curves recorded for five sensor elements under RT conditions. As one can see, the I(V) follows a linear function under various slopes depending on the MXene-layer resistance. The observed differences belong to a spatial variation between the elements, primarily in a volume of ink-deposited material as a prerequisite for the differentiation of the sensor elements for combining a sensor array, as discussed later. The resistance magnitudes lie in a sub-MOhm range which is easy to record and process with conventional read-out electronics.
The Ti0.2V1.8C MXene-based sensors were exposed to five volatile analytes of three alcohols (methanol, ethanol, and isopropanol), ammonia, and water vapors at a concentration, C, range of 500–16,000 ppm. The exposures to analytes at certain concentrations were repeated twice to ensure the reproducibility of the records. The typical resistance transients recorded for the exemplary sensor element operated under RT are drawn in Figure 7c regarding all five of the analytes. When the vapors appear, the resistance goes up with a magnitude defining by the analyte’s concentration; the higher the concentration, the larger the resistance gain. This behavior is quite characteristic to match the known one in MXenes chemiresistors [19,21] and fully match the quantum-chemical calculations given in Section 3.2. Still, we may note that the sensor response to methanol present at the concentrations below 2000 ppm is rather negligible, while other analytes deliver a clear chemiresistive response to all the test analytes in the whole concentration range under study. This response is reproducible and reversible. It is worth noting that we observe some residual steady enhancing of the resistance when the concentrations of the analytes exceed 1000 ppm, which indicates that not all the adsorbed species are fast desorbed from the surface under RT conditions. Normally, to accelerate these processes, some activation via a slight heating or ultra-violet irradiation is required [58].
We have estimated the recorded chemiresistive response, S, according to Formula (6) and plotted it versus an analyte concentration, as a major characteristic of the gas sensor, in Figure 7b. The curves go linear in a double log scale according to the S ~ C n function, which is consistent with the Freundlich isotherm. The slope of the curves depends on the power index, n, which is distinctive for various analytes. We have collected the observed values of n in Table 3. As one can see, these values lie in the range of 0.38–0.61 which are similar to the ones observed in previous experiments with MXene structures [21].
Further, we have taken the magnitude of the electrical noise of the sensor recorded upon background air exposures and estimated the signal-to-noise ratio (SNR) for finding a limit of detection (LOD). Thus, the LOD value is determined as the lowest analyte concentration which yields a × 3 magnitude of SNR [59]. The calculated LOD values for various analytes are collected in Table 3. It is worth noting that the weakest chemiresistive effect is observed in case of methanol vapors with an LOD value equal to ca. 181 ppm, while the water vapors influence the MXene resistance in most degrees, yielding an LOD of more than an order lower, to be about 14 ppm. The humidity exhibits also the highest value of n equal to ca. 0.61 and the higher magnitude of the response. Therefore, we have compared the observed response of Ti0.2V1.8C MXene-based sensors to H2O vapors with the ones previously published in the literature on titanium carbide MXenes as sensors (Figure 8a).
Here, we employ the S/C value because of the variations in the presentation of the data in the literature. This parameter is close to a sensitivity one, which is normally defined as ΔS/ΔC. It might be seen from Figure 8a that the Ti0.2V1.8C MXene slightly outperforms the literature data on the chemiresistive sensors based on titanium carbide MXenes in the range of water concentrations below 15,000 ppm. At a higher H2O, there are few reports on the capacitive-type and quartz microbalance sensors whose response is still higher but was, however, recorded in the narrower range of the concentrations. The finding on so high a response of the Ti0.2V1.8C MXene-based sensor to the water vapors fully corresponds to our DFT calculations. In order to match the experimental data with the DFT ones, we have plotted the observed chemiresistive response to the test analytes at the highest concentration of 16,000 ppm in Figure 8b. While comparing this plot to the DFT data present in Figure 6b, we should note a large alignment between the experimental and theoretical estimations of the response. Still, the largest difference is the response to methanol vapors which has to deliver the effect comparable to the ethanol ones according to the DFT, while the experiment shows its effect as the lowest one. These differences should be further clarified.
Furthermore, the same (positive) sign of the chemiresistive response of Ti0.2V1.8C MXene-based sensors to various analytes does not allow us to selectively distinguish the test analytes without a prior knowledge on the probe content similar to other researches [67]. Indeed, if we take only the response magnitude as a measure, it does not provide the selectivity as discussed in numerous works [30]. Say, the response equal to ca. 1% might be induced (Figure 7b) by water vapors at 317 ppm or ethanol at 1219 ppm, or ammonia at 2119 ppm, or isopropanol at 3793 ppm. Therefore, to expand the sensor’s performance, we have collected the resistances of the five sensor elements in the Ti0.2V1.8C MXene-based on-chip array recorded in the stationary conditions upon exposing to various analytes. These multisensory data were processed with a linear discriminant analysis (LDA) protocol to build an artificial space with a dimensionality equal to N-1, where N is the number of classes to be recognized [68]; in our case N = 5. For framing the classes, we have employed a normal distribution of the vector data within each class with a 0.9 confidence level. The obtained LDA plot is shown in Figure 8c in a 2D cross-section of the first two LDA components. While the physical meaning of the LDA components is unknown, they reflect the specific features of the classes under interest, in our case the analytes, in relation to the vector input composing of the resistances of Ti0.2V1.8C MXene-based sensors. In Figure 8c, the gravity centers of all the classes are well distanced from the center of the coordinate system, which means that the analytes yield some specific output which links to the variations in the charge exchange in the adsorbate/adsorbent system in the framework of the DFT consideration discussed in Section 3.2. Moreover, these differences are significant enough to separate the analyte-related classes in the LDA space; the average Mahalonobis distance here is 73.6 un. It shows the ability to selectively distinguish the analytes according to the known E-nose approach [69].

4. Conclusions

In this study, quantum-chemical calculations and experimental chemosensory measurements were carried out to evaluate the possibility of employing partially substituted TixV2-xC (where x = 0.2) MXenes as a receptor component of the gas sensor and gas-analytical multisensor array. For this purpose, the synthesis process of the precursor, the Ti0.2V1.8AlC MAX phase, was investigated, and the corresponding MXene (Ti0.2V1.8C) was obtained following a selective etching of aluminum. The multilayer MXene agglomerates were delaminated into individual two-dimensional structures using tetramethylammonium hydroxide, and the resulting stable disperse system was utilized as the functional ink for the microextrusion printing of Ti0.2V1.8C layers on the surface of the multielectrode chip. The formation of the target MAX-phase and MXene was confirmed by X-ray diffraction analysis, and the results of the energy dispersive X-ray microanalysis indicated the given ratio of the metals. The performed quantum-chemical calculations allowed us to estimate the contribution of various analytes adsorption on the surface of MXenes with a V2C, Ti2C, and Ti0.2V1.8C structure into changing their electronic structure and, accordingly, the chemiresistive response. It was demonstrated that the greatest change in the electrical resistance of the Ti0.2V1.8C MXene is observed upon the adsorption of H2O molecules when compared to other test analytes (methanol, ethanol, isopropanol, and ammonia). The theoretical and experimental results are in a good agreement, except for the methanol effect which needs a further clarification. To further improve the gas sensitivity, these structures could be combined into heterojunctions with oxides [70], including a partial oxidation of the pristine material [71].
The selectivity to the analytes has been approached via the multisensory processing of the array signal recorded from Ti0.2V1.8C MXene-based sensors. These findings show that designing complex MXene structures with a substitution of a major matrix with foreign atoms, here the V2C with some Ti ones, might be an effective strategy to optimize the functional properties of this layered material.

Author Contributions

Conceptualization, V.V.S. and N.P.S.; investigation, N.P.S., I.A.N., I.A.P., A.S.V., D.A.K., T.L.S. and E.P.S.; writing—original draft preparation, N.P.S., T.L.S. and D.A.K.; writing—review and editing, N.P.S. and V.V.S.; visualization, N.P.S., D.A.K., I.A.P. and A.S.V.; supervision, N.P.S., O.E.G., V.V.S. and N.T.K. 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 20 December 2022) (in the part of materials synthesis, quantum-chemical calculations, and chemosensory measurements) and the Ministry of Science and Higher Education of the Russian Federation, the contract 075-15-2021-709, unique identifier of the project RF-2296.61321X0037 (in the part of the reaction system preparation using a planetary ball mill).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding authors.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yuan, C.; Ma, J.; Zou, Y.; Li, G.; Xu, H.; Sysoev, V.V.; Cheng, X.; Deng, Y. Modeling Interfacial Interaction between Gas Molecules and Semiconductor Metal Oxides: A New View Angle on Gas Sensing. Adv. Sci. 2022, 9, 2203594. [Google Scholar] [CrossRef] [PubMed]
  2. Mirzaei, A.; Ansari, H.R.; Shahbaz, M.; Kim, J.-Y.; Kim, H.W.; Kim, S.S. Metal Oxide Semiconductor Nanostructure Gas Sensors with Different Morphologies. Chemosensors 2022, 10, 289. [Google Scholar] [CrossRef]
  3. Korotcenkov, G.; Brinzari, V.; Ham, M.H. Materials Acceptable for Gas Sensor Design: Advantages and Limitations. Key Eng. Mater. 2018, 780, 80–89. [Google Scholar] [CrossRef]
  4. Ou, L.-X.; Liu, M.-Y.; Zhu, L.-Y.; Zhang, D.W.; Lu, H.-L. Recent Progress on Flexible Room-Temperature Gas Sensors Based on Metal Oxide Semiconductor. Nano-Micro Lett. 2022, 14, 206. [Google Scholar] [CrossRef] [PubMed]
  5. Novoselov, K. Mind the Gap. Nat. Mater. 2007, 6, 720–721. [Google Scholar] [CrossRef] [PubMed]
  6. Meng, Z.; Stolz, R.M.; Mendecki, L.; Mirica, K.A. Electrically-Transduced Chemical Sensors Based on Two-Dimensional Nanomaterials. Chem. Rev. 2019, 119, 478–598. [Google Scholar] [CrossRef]
  7. Schedin, F.; Geim, A.K.; Morozov, S.V.; Hill, E.W.; Blake, P.; Katsnelson, M.I.; Novoselov, K.S. Detection of Individual Gas Molecules Adsorbed on Graphene. Nat. Mater. 2007, 6, 652–655. [Google Scholar] [CrossRef]
  8. Dral, A.P.; ten Elshof, J.E. 2D Metal Oxide Nanoflakes for Sensing Applications: Review and Perspective. Sens. Actuators B Chem. 2018, 272, 369–392. [Google Scholar] [CrossRef] [Green Version]
  9. Donarelli, M.; Ottaviano, L. 2D Materials for Gas Sensing Applications: A Review on Graphene Oxide, MoS2, WS2 and Phosphorene. Sensors 2018, 18, 3638. [Google Scholar] [CrossRef] [Green Version]
  10. Lee, E.; Yoon, Y.S.; Kim, D.-J. Two-Dimensional Transition Metal Dichalcogenides and Metal Oxide Hybrids for Gas Sensing. ACS Sens. 2018, 3, 2045–2060. [Google Scholar] [CrossRef]
  11. Anichini, C.; Czepa, W.; Pakulski, D.; Aliprandi, A.; Ciesielski, A.; Samorì, P. Chemical Sensing with 2D Materials. Chem. Soc. Rev. 2018, 47, 4860–4908. [Google Scholar] [CrossRef] [Green Version]
  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. Liu, X.; Ma, T.; Pinna, N.; Zhang, J. Two-Dimensional Nanostructured Materials for Gas Sensing. Adv. Funct. Mater. 2017, 27, 1702168. [Google Scholar] [CrossRef]
  14. Yang, S.; Jiang, C.; Wei, S. Gas Sensing in 2D Materials. Appl. Phys. Rev. 2017, 4, 021304. [Google Scholar] [CrossRef]
  15. Bhardwaj, R.; Hazra, A. MXene-Based Gas Sensors. J. Mater. Chem. C 2021, 9, 15735–15754. [Google Scholar] [CrossRef]
  16. Lee, E.; VahidMohammadi, A.; Prorok, B.C.; Yoon, Y.S.; Beidaghi, M.; Kim, D.-J. Room Temperature Gas Sensing of Two-Dimensional Titanium Carbide (MXene). ACS Appl. Mater. Interfaces 2017, 9, 37184–37190. [Google Scholar] [CrossRef] [PubMed]
  17. Chertopalov, S.; Mochalin, V.N. Environment-Sensitive Photoresponse of Spontaneously Partially Oxidized Ti3C2 MXene Thin Films. ACS Nano 2018, 12, 6109–6116. [Google Scholar] [CrossRef] [PubMed]
  18. 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] [PubMed] [Green Version]
  19. Zhao, Q.-N.; Zhang, Y.-J.; Duan, Z.-H.; Wang, S.; Liu, C.; Jiang, Y.-D.; Tai, H.-L. A Review on Ti3C2Tx-Based Nanomaterials: Synthesis and Applications in Gas and Humidity Sensors. Rare Met. 2021, 40, 1459–1476. [Google Scholar] [CrossRef]
  20. Yang, Z.; Liu, A.; Wang, C.; Liu, F.; He, J.; Li, S.; Wang, J.; You, R.; Yan, X.; Sun, P.; et al. Improvement of Gas and Humidity Sensing Properties of Organ-like MXene by Alkaline Treatment. ACS Sens. 2019, 4, 1261–1269. [Google Scholar] [CrossRef]
  21. Pazniak, H.; Varezhnikov, A.S.; Kolosov, D.A.; Plugin, I.A.; Vito, A.D.; 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]
  22. He, T.; Liu, W.; Lv, T.; Ma, M.; Liu, Z.; Vasiliev, A.; Li, X. MXene/SnO2 Heterojunction Based Chemical Gas Sensors. Sens. Actuators B Chem. 2021, 329, 129275. [Google Scholar] [CrossRef]
  23. Gogotsi, Y.; Huang, Q. MXenes: Two-Dimensional Building Blocks for Future Materials and Devices. ACS Nano 2021, 15, 5775–5780. [Google Scholar] [CrossRef] [PubMed]
  24. Lee, E.; VahidMohammadi, A.; Yoon, Y.S.; Beidaghi, M.; Kim, D.-J. Two-Dimensional Vanadium Carbide MXene for Gas Sensors with Ultrahigh Sensitivity Toward Nonpolar Gases. ACS Sens. 2019, 4, 1603–1611. [Google Scholar] [CrossRef] [PubMed]
  25. Xu, X.; Yin, M.; Li, N.; Wang, W.; Sun, B.; Liu, M.; Zhang, D.; Li, Z.; Wang, C. Vanadium-Doped Tin Oxide Porous Nanofibers: Enhanced Responsivity for Hydrogen Detection. Talanta 2017, 167, 638–644. [Google Scholar] [CrossRef]
  26. Wang, Y.-T.; Whang, W.-T.; Chen, C.-H. Hollow V2O5 Nanoassemblies for High-Performance Room-Temperature Hydrogen Sensors. ACS Appl. Mater. Interfaces 2015, 7, 8480–8487. [Google Scholar] [CrossRef]
  27. Gardner, J.W.; Bartlett, P.N. A Brief History of Electronic Noses. Sens. Actuators B Chem. 1994, 18, 210–211. [Google Scholar] [CrossRef]
  28. Sysoev, V.V.; Strelcov, V.V.; Kolmakov, A. Multisensor Micro-Arrays Based on Metal Oxide Nanowires for Electronic Nose Applications. In Metal Oxide Nanomaterials for Chemical Sensors; Carpenter, M.A., Mathur, S., Kolmakov, A., Eds.; Springer Science + Business Media: New York, NY, USA, 2013; pp. 425–502. [Google Scholar]
  29. 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]
  30. Hierlemann, A.; Gutierrez-Osuna, R. Higher-Order Chemical Sensing. Chem. Rev. 2008, 108, 563–613. [Google Scholar] [CrossRef]
  31. Khazaei, M.; Arai, M.; Sasaki, T.; Chung, C.-Y.; Venkataramanan, N.S.; Estili, M.; Sakka, Y.; Kawazoe, Y. Novel Electronic and Magnetic Properties of Two-Dimensional Transition Metal Carbides and Nitrides. Adv. Funct. Mater. 2013, 23, 2185–2192. [Google Scholar] [CrossRef]
  32. Anasori, B.; Lukatskaya, M.R.; Gogotsi, Y. 2D Metal Carbides and Nitrides (MXenes) for Energy Storage. Nat. Rev. Mater. 2017, 2, 16098. [Google Scholar] [CrossRef]
  33. Anasori, B.; Xie, Y.; Beidaghi, M.; Lu, J.; Hosler, B.C.; Hultman, L.; Kent, P.R.C.; Gogotsi, Y.; Barsoum, M.W. Two-Dimensional, Ordered, Double Transition Metals Carbides (MXenes). ACS Nano 2015, 9, 9507–9516. [Google Scholar] [CrossRef] [PubMed]
  34. Zhan, X.; Si, C.; Zhou, J.; Sun, Z. MXene and MXene-Based Composites: Synthesis, Properties and Environment-Related Applications. Nanoscale Horiz. 2020, 5, 235–258. [Google Scholar] [CrossRef]
  35. Khazaei, M.; Ranjbar, A.; Arai, M.; Sasaki, T.; Yunoki, S. Electronic Properties and Applications of MXenes: A Theoretical Review. J. Mater. Chem. C 2017, 5, 2488–2503. [Google Scholar] [CrossRef] [Green Version]
  36. Anasori, B.; Shi, C.; Moon, E.J.; Xie, Y.; Voigt, C.A.; Kent, P.R.C.; May, S.J.; Billinge, S.J.L.; Barsoum, M.W.; Gogotsi, Y. Control of Electronic Properties of 2D Carbides (MXenes) by Manipulating Their Transition Metal Layers. Nanoscale Horiz. 2016, 1, 227–234. [Google Scholar] [CrossRef]
  37. Khazaei, M.; Mishra, A.; Venkataramanan, N.S.; Singh, A.K.; Yunoki, S. Recent Advances in MXenes: From Fundamentals to Applications. Curr. Opin. Solid State Mater. Sci. 2019, 23, 164–178. [Google Scholar] [CrossRef] [Green Version]
  38. Gao, G.; Ding, G.; Li, J.; Yao, K.; Wu, M.; Qian, M. Monolayer MXenes: Promising Half-Metals and Spin Gapless Semiconductors. Nanoscale 2016, 8, 8986–8994. [Google Scholar] [CrossRef] [Green Version]
  39. Kumar, H.; Frey, N.C.; Dong, L.; Anasori, B.; Gogotsi, Y.; Shenoy, V.B. Tunable Magnetism and Transport Properties in Nitride MXenes. ACS Nano 2017, 11, 7648–7655. [Google Scholar] [CrossRef]
  40. Weng, H.; Ranjbar, A.; Liang, Y.; Song, Z.; Khazaei, M.; Yunoki, S.; Arai, M.; Kawazoe, Y.; Fang, Z.; Dai, X. Large-Gap Two-Dimensional Topological Insulator in Oxygen Functionalized MXene. Phys. Rev. B 2015, 92, 075436. [Google Scholar] [CrossRef] [Green Version]
  41. He, J.; Lyu, P.; Nachtigall, P. New Two-Dimensional Mn-Based MXenes with Room-Temperature Ferromagnetism and Half-Metallicity. J. Mater. Chem. C 2016, 4, 11143–11149. [Google Scholar] [CrossRef]
  42. Si, C.; Zhou, J.; Sun, Z. Half-Metallic Ferromagnetism and Surface Functionalization-Induced Metal–Insulator Transition in Graphene-like Two-Dimensional Cr2C Crystals. ACS Appl. Mater. Interfaces 2015, 7, 17510–17515. [Google Scholar] [CrossRef] [PubMed]
  43. Zha, X.-H.; Luo, K.; Li, Q.; Huang, Q.; He, J.; Wen, X.; Du, S. Role of the Surface Effect on the Structural, Electronic and Mechanical Properties of the Carbide MXenes. EPL (Europhys. Lett.) 2015, 111, 26007. [Google Scholar] [CrossRef]
  44. 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]
  45. Dash, A.; Vaßen, R.; Guillon, O.; Gonzalez-Julian, J. Molten Salt Shielded Synthesis of Oxidation Prone Materials in Air. Nat. Mater. 2019, 18, 465–470. [Google Scholar] [CrossRef] [PubMed]
  46. Rabchinskii, M.K.; Varezhnikov, A.S.; Sysoev, V.V.; Solomatin, M.A.; Ryzhkov, S.A.; Baidakova, M.V.; Stolyarova, D.Y.; Shnitov, V.V.; Pavlov, S.S.; Kirilenko, D.A.; et al. Hole-Matrixed Carbonylated Graphene: Synthesis, Properties, and Highly-Selective Ammonia Gas Sensing. Carbon 2021, 172, 236–247. [Google Scholar] [CrossRef]
  47. Fedorov, F.S.; Varezhnikov, A.S.; Kiselev, I.; Kolesnichenko, V.V.; Burmistrov, I.N.; Sommer, M.; Fuchs, D.; Kübel, C.; Gorokhovsky, A.V.; Sysoev, V.V. Potassium Polytitanate Gas-Sensor Study by Impedance Spectroscopy. Anal. Chim. Acta 2015, 897, 81–86. [Google Scholar] [CrossRef] [PubMed]
  48. Naqvi, S.R.; Shukla, V.; Jena, N.K.; Luo, W.; Ahuja, R. Exploring Two-Dimensional M2NS2 (M = Ti, V) MXenes Based Gas Sensors for Air Pollutants. Appl. Mater. Today 2020, 19, 100574. [Google Scholar] [CrossRef]
  49. Mehdi Aghaei, S.; Aasi, A.; Panchapakesan, B. Experimental and Theoretical Advances in MXene-Based Gas Sensors. ACS Omega 2021, 6, 2450–2461. [Google Scholar] [CrossRef]
  50. Nahirniak, S.; Saruhan, B. MXene Heterostructures as Perspective Materials for Gas Sensing Applications. Sensors 2022, 22, 972. [Google Scholar] [CrossRef]
  51. Akkuş, Ü.Ö.; Balcı, E.; Berber, S. Mo2TiC2O2 MXene-based nanoscale pressure sensor. Phys. E Low-Dimens. Syst. Nanostruct. 2020, 116, 113762. [Google Scholar] [CrossRef]
  52. Soler, J.M.; Artacho, E.; Gale, J.D.; García, A.; Junquera, J.; Ordejón, P.; Sánchez-Portal, D. The SIESTA Method for Ab Initio Order- N Materials Simulation. J. Phys. Condens. Matter 2002, 14, 2745–2779. [Google Scholar] [CrossRef] [Green Version]
  53. García, A.; Papior, N.; Akhtar, A.; Artacho, E.; Blum, V.; Bosoni, E.; Brandimarte, P.; Brandbyge, M.; Cerdá, J.I.; Corsetti, F.; et al. Siesta: Recent Developments and Applications. J. Chem. Phys. 2020, 152, 204108. [Google Scholar] [CrossRef] [PubMed]
  54. Monkhorst, H.J.; Pack, J.D. Special Points for Brillouin-Zone Integrations. Phys. Rev. B 1976, 13, 5188–5192. [Google Scholar] [CrossRef]
  55. Johnson, D.D. Modified Broyden’s Method for Accelerating Convergence in Self-Consistent Calculations. Phys. Rev. B 1988, 38, 12807–12813. [Google Scholar] [CrossRef] [PubMed]
  56. Büttiker, M.; Imry, Y.; Landauer, R.; Pinhas, S. Generalized Many-Channel Conductance Formula with Application to Small Rings. Phys. Rev. B 1985, 31, 6207–6215. [Google Scholar] [CrossRef] [Green Version]
  57. Lu, H.; Dai, D.; Yang, P.; Li, L. Atomic Orbitals in Molecules: General Electronegativity and Improvement of Mulliken Population Analysis. Phys. Chem. Chem. Phys. 2006, 8, 340–346. [Google Scholar] [CrossRef]
  58. 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]
  59. D’Amico, A.; Di Natale, C. A Contribution on Some Basic Definitions of Sensors Properties. IEEE Sens. J. 2001, 1, 183–190. [Google Scholar] [CrossRef]
  60. Muckley, E.S.; Naguib, M.; Wang, H.-W.; Vlcek, L.; Osti, N.C.; Sacci, R.L.; Sang, X.; Unocic, R.R.; Xie, Y.; Tyagi, M.; et al. Multimodality of Structural, Electrical, and Gravimetric Responses of Intercalated MXenes to Water. ACS Nano 2017, 11, 11118–11126. [Google Scholar] [CrossRef]
  61. Liu, L.; Chen, W.; Zhang, H.; Wang, Q.; Guan, F.; Yu, Z. Flexible and Multifunctional Silk Textiles with Biomimetic Leaf-Like MXene/Silver Nanowire Nanostructures for Electromagnetic Interference Shielding, Humidity Monitoring, and Self-Derived Hydrophobicity. Adv. Funct. Mater. 2019, 29, 1905197. [Google Scholar] [CrossRef]
  62. Muckley, E.S.; Naguib, M.; Ivanov, I.N. Multi-Modal, Ultrasensitive, Wide-Range Humidity Sensing with Ti3C2 Film. Nanoscale 2018, 10, 21689–21695. [Google Scholar] [CrossRef] [PubMed]
  63. Li, N.; Jiang, Y.; Zhou, C.; Xiao, Y.; Meng, B.; Wang, Z.; Huang, D.; Xing, C.; Peng, Z. High-Performance Humidity Sensor Based on Urchin-Like Composite of Ti3C2 MXene-Derived TiO2 Nanowires. ACS Appl. Mater. Interfaces 2019, 11, 38116–38125. [Google Scholar] [CrossRef] [PubMed]
  64. Zhao, X.; Wang, L.-Y.; Tang, C.-Y.; Zha, X.-J.; Liu, Y.; Su, B.-H.; Ke, K.; Bao, R.-Y.; Yang, M.-B.; Yang, W. Smart Ti3C2Tx MXene Fabric with Fast Humidity Response and Joule Heating for Healthcare and Medical Therapy Applications. ACS Nano 2020, 14, 8793–8805. [Google Scholar] [CrossRef] [PubMed]
  65. An, H.; Habib, T.; Shah, S.; Gao, H.; Patel, A.; Echols, I.; Zhao, X.; Radovic, M.; Green, M.J.; Lutkenhaus, J.L. Water Sorption in MXene/Polyelectrolyte Multilayers for Ultrafast Humidity Sensing. ACS Appl. Nano Mater. 2019, 2, 948–955. [Google Scholar] [CrossRef]
  66. Li, Z.; Zhang, B.; Fu, C.; Tao, R.; Li, H.; Luo, J. Ultra-Fast and Sensitive Hydrophobic QCM Humidity Sensor by Sulfur Modified Ti3C2Tx MXene. IEEE Sens. J. 2022. [Google Scholar] [CrossRef]
  67. Li, Q.; Li, Y.; Zeng, W. Preparation and Application of 2D MXene-Based Gas Sensors: A Review. Chemosensors 2021, 9, 225. [Google Scholar] [CrossRef]
  68. Kiselev, I.; Sysoev, V.; Kaikov, I.; Koronczi, I.; Adil Akai Tegin, R.; Smanalieva, J.; Sommer, M.; Ilicali, C.; Hauptmannl, M. On the Temporal Stability of Analyte Recognition with an E-Nose Based on a Metal Oxide Sensor Array in Practical Applications. Sensors 2018, 18, 550. [Google Scholar] [CrossRef] [Green Version]
  69. Persaud, K.; Dodd, G. Analysis of Discrimination Mechanisms in the Mammalian Olfactory System Using a Model Nose. Nature 1982, 299, 352–355. [Google Scholar] [CrossRef]
  70. Zhang, H.-F.; Xuan, J.-Y.; Zhang, Q.; Sun, M.-L.; Jia, F.-C.; Wang, X.-M.; Yin, G.-C.; Lu, S.-Y. Strategies and Challenges for Enhancing Performance of MXene-Based Gas Sensors: A Review. Rare Met. 2022, 41, 3976–3999. [Google Scholar] [CrossRef]
  71. 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]
Figure 1. The technology route to form the chemiresistor/on-chip multisensor array from Ti0.2V1.8C MXene. The denotings are, 1—fabrication of Ti, V, Al, C, and KBr powders, 2—grinding of the powders, 3—molding of the intermediate products by pressing, 4—heat-treating the intermediate products, 5—water-washing of samples, 6—obtaining the Ti0.2V1.8AlC MAX phase, 7—a chemical etching of Al, 8—a preparing Ti0.2V1.8C MXene multilayer agglomerates, 9—a delamination of the agglomerates by ultrasonic treatment, 10—a separation/purification of layered MXenes via the centrifugation, 11—placing the MXene structures from inks over the multielectroded chip, and 12—the output chip wired into holder.
Figure 1. The technology route to form the chemiresistor/on-chip multisensor array from Ti0.2V1.8C MXene. The denotings are, 1—fabrication of Ti, V, Al, C, and KBr powders, 2—grinding of the powders, 3—molding of the intermediate products by pressing, 4—heat-treating the intermediate products, 5—water-washing of samples, 6—obtaining the Ti0.2V1.8AlC MAX phase, 7—a chemical etching of Al, 8—a preparing Ti0.2V1.8C MXene multilayer agglomerates, 9—a delamination of the agglomerates by ultrasonic treatment, 10—a separation/purification of layered MXenes via the centrifugation, 11—placing the MXene structures from inks over the multielectroded chip, and 12—the output chip wired into holder.
Chemosensors 11 00007 g001
Figure 2. The experimental setup to measure the chemiresistive response of Ti0.2V1.8C MXene-based sensor chip which consists of the gas-delivery part based onmixing the dried lab air with few test analyte vapors, including H2O, yielded via bubbling to the gas line, and read-out electronics employing resistance measuring of number of on-chip sensor elements via a multiplex card.
Figure 2. The experimental setup to measure the chemiresistive response of Ti0.2V1.8C MXene-based sensor chip which consists of the gas-delivery part based onmixing the dried lab air with few test analyte vapors, including H2O, yielded via bubbling to the gas line, and read-out electronics employing resistance measuring of number of on-chip sensor elements via a multiplex card.
Chemosensors 11 00007 g002
Figure 3. XRD patterns of the synthesized Ti0.2V1.8AlC MAX phase and Ti0.2V1.8C MXene samples.
Figure 3. XRD patterns of the synthesized Ti0.2V1.8AlC MAX phase and Ti0.2V1.8C MXene samples.
Chemosensors 11 00007 g003
Figure 4. The characterization of Ti0.2V1.8C MXene layer placed from ink over the glass substrate by SEM under various magnifications: (a) ×2000, (b) ×100,000.
Figure 4. The characterization of Ti0.2V1.8C MXene layer placed from ink over the glass substrate by SEM under various magnifications: (a) ×2000, (b) ×100,000.
Chemosensors 11 00007 g004
Figure 5. The DFT modeling of charge state for interaction between Ti0.2V1.8C and analyte molecules: (a) case of methanol; (b) case of ethanol; (c) case of isopropanol; (d) case of ammonia; (e) case of water; (f) density of electronic states in Ti0.2V1.8C under the analyte adsorption relative to Fermi level.
Figure 5. The DFT modeling of charge state for interaction between Ti0.2V1.8C and analyte molecules: (a) case of methanol; (b) case of ethanol; (c) case of isopropanol; (d) case of ammonia; (e) case of water; (f) density of electronic states in Ti0.2V1.8C under the analyte adsorption relative to Fermi level.
Chemosensors 11 00007 g005
Figure 6. The DFT modeling of interaction between Ti0.2V1.8C MXene and analyte molecules under study: (a) DOS without smearing characterizing the supercell upon adsorption of the analytes; (b) dependence of the supercell resistance and its chemiresistive response upon the analyte adsorption.
Figure 6. The DFT modeling of interaction between Ti0.2V1.8C MXene and analyte molecules under study: (a) DOS without smearing characterizing the supercell upon adsorption of the analytes; (b) dependence of the supercell resistance and its chemiresistive response upon the analyte adsorption.
Chemosensors 11 00007 g006
Figure 7. The gas sensing performance of Ti0.2V1.8C MXene-based chemiresistive sensor under RT conditions: (a) I-V curves for five sensor elements; (b) the chemiresistive response, S(C), in dependence on volatile analyte concentration, the points are experimental data, the curves are fitting with S ~ C n function; (c) the typical sensor R(t) transient upon exposing to five volatile analytes in the wide range of concentrations, C = 500–16,000 ppm.
Figure 7. The gas sensing performance of Ti0.2V1.8C MXene-based chemiresistive sensor under RT conditions: (a) I-V curves for five sensor elements; (b) the chemiresistive response, S(C), in dependence on volatile analyte concentration, the points are experimental data, the curves are fitting with S ~ C n function; (c) the typical sensor R(t) transient upon exposing to five volatile analytes in the wide range of concentrations, C = 500–16,000 ppm.
Chemosensors 11 00007 g007
Figure 8. The gas selectivity of Ti0.2V1.8c MXene-based sensors to volatile analyte molecules under study: (a) [20,60,61,62,63,64,65,66] the comparison of chemiresistive response to H2O vapors with literature data dealing with sensors based on titanium carbide MXenes structures; (b) the response to the analytes present at C = 16,000 ppm; (c) results of LDA processing the vector signals of 5-element sensor array to the analytes; 2D cross-section of 4D LDA space is present, points are vector signals to analytes present at 16,000 ppm of concentration, ellipses are built around gravity centers of analyte-related classes with 0.9 confidence under normal distribution.
Figure 8. The gas selectivity of Ti0.2V1.8c MXene-based sensors to volatile analyte molecules under study: (a) [20,60,61,62,63,64,65,66] the comparison of chemiresistive response to H2O vapors with literature data dealing with sensors based on titanium carbide MXenes structures; (b) the response to the analytes present at C = 16,000 ppm; (c) results of LDA processing the vector signals of 5-element sensor array to the analytes; 2D cross-section of 4D LDA space is present, points are vector signals to analytes present at 16,000 ppm of concentration, ellipses are built around gravity centers of analyte-related classes with 0.9 confidence under normal distribution.
Chemosensors 11 00007 g008
Table 1. Mulliken charges, e, characterizing the interaction of analyte molecules with MXene surfaces.
Table 1. Mulliken charges, e, characterizing the interaction of analyte molecules with MXene surfaces.
MXeneAnalyte
MethanolEthanolIsopropanolAmmoniaWater
Ti0.2V1.8C−0.01e−0.11e−0.006e0.007e0.003e
Ti2C−0.028e−0.061e−0.016e0.019e0.012e
V2C−0.006e−0.009e−0.011e0.015e0.004e
Table 2. The calculated binding energy, eV, characterizing the interaction of analyte molecules with MXene surfaces.
Table 2. The calculated binding energy, eV, characterizing the interaction of analyte molecules with MXene surfaces.
MXeneAnalyte
MethanolEthanolIsopropanolAmmoniaWater
Ti0.2V1.8C−0.051−0.044−0.041−0.058−0.062
Ti2C−0.101−0.128−0.112−0.118−0.082
V2C−0.084−0.098−0.073−0.106−0.063
Table 3. The characteristics of Ti0.2V1.8C MXene-based sensor versus volatile test analytes: power index, n, characterizing S~Cn fitting of experimentally observed chemiresistive response and calculated limit of detection (LOD).
Table 3. The characteristics of Ti0.2V1.8C MXene-based sensor versus volatile test analytes: power index, n, characterizing S~Cn fitting of experimentally observed chemiresistive response and calculated limit of detection (LOD).
AnalyteMethanolEthanolIsopropanolAmmoniaWater
n0.38 ± 0.10.55 ± 0.010.49 ± 0.020.51 ± 0.020.61 ± 0.03
LOD, ppm18124392714
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, 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 2023, 11, 7. https://doi.org/10.3390/chemosensors11010007

AMA Style

Simonenko NP, Glukhova OE, Plugin IA, Kolosov DA, Nagornov IA, Simonenko TL, Varezhnikov AS, Simonenko EP, Sysoev VV, Kuznetsov NT. The Ti0.2V1.8C MXene Ink-Prepared Chemiresistor: From Theory to Tests with Humidity versus VOCs. Chemosensors. 2023; 11(1):7. https://doi.org/10.3390/chemosensors11010007

Chicago/Turabian Style

Simonenko, Nikolay P., Olga E. Glukhova, Ilya A. Plugin, Dmitry A. Kolosov, Ilya A. Nagornov, Tatiana L. Simonenko, Alexey S. Varezhnikov, Elizaveta P. Simonenko, Victor V. Sysoev, and Nikolay T. Kuznetsov. 2023. "The Ti0.2V1.8C MXene Ink-Prepared Chemiresistor: From Theory to Tests with Humidity versus VOCs" Chemosensors 11, no. 1: 7. https://doi.org/10.3390/chemosensors11010007

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