Next Article in Journal
Nanoarchitectonics of Spherical Nucleic Acids with Biodegradable Polymer Cores: Synthesis and Evaluation
Next Article in Special Issue
Crystal Structure, Raman, FTIR, UV-Vis Absorption, Photoluminescence Spectroscopy, TG–DSC and Dielectric Properties of New Semiorganic Crystals of 2-Methylbenzimidazolium Perchlorate
Previous Article in Journal
Investigation of Soft Magnetic Material Fe-6.5Si Fracture Obtained by Additive Manufacturing
Previous Article in Special Issue
Recrystallization of Si Nanoparticles in Presence of Chalcogens: Improved Electrical and Optical Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Electrical and Gas Sensor Properties of Nb(V) Doped Nanocrystalline β-Ga2O3

1
Department of Chemistry, Lomonosov Moscow State University, Leninskie Gory 1/3, 119234 Moscow, Russia
2
A.V. Topchiev Institute of Petrochemical Synthesis, Russian Academy of Sciences, Leninsky Prospect 29, 119991 Moscow, Russia
3
Udmurt Federal Research Center of the Ural Branch of the Russian Academy of Sciences, Tatyana Baramzina St. 34, 426067 Izhevsk, Russia
4
Scientific-Manufacturing Complex «Technological Centre», Shokina Square, House 1, Bld. 7 Off. 7237, 124498 Zelenograd, Moscow, Russia
*
Author to whom correspondence should be addressed.
Materials 2022, 15(24), 8916; https://doi.org/10.3390/ma15248916
Submission received: 21 October 2022 / Revised: 5 December 2022 / Accepted: 8 December 2022 / Published: 13 December 2022
(This article belongs to the Special Issue New Trends in Crystalline Materials)

Abstract

:
A flame spray pyrolysis (FSP) technique was applied to obtain pure and Nb(V)-doped nanocrystalline β-Ga2O3, which were further studied as gas sensor materials. The obtained samples were characterized with XRD, XPS, TEM, Raman spectroscopy and BET method. Formation of GaNbO4 phase is observed at high annealing temperatures. Transition of Ga(III) into Ga(I) state during Nb(V) doping prevents donor charge carriers generation and hinders considerable improvement of electrical and gas sensor properties of β-Ga2O3. Superior gas sensor performance of obtained ultrafine materials at lower operating temperatures compared to previously reported thin film Ga2O3 materials is shown.

1. Introduction

Metal oxide semiconductor gas sensors find their applications in many important fields: industrial safety [1], ecological monitoring [2], indoor air quality assessment [3], non-invasive medical diagnostics [4] and others [5,6]. Significant drawback of this type of sensors—low selectivity of response (high cross-sensitivity)—can be overcome by utilization of machine learning algorithms, which have been proposed on either single sensor, or multi-sensor systems [7,8,9]. Additionally, machine learning techniques allow to partially compensate response instability and drift effects [10]. However, the stability of intrinsic electrical and chemical characteristics of gas sensor material is still very important, otherwise more or less frequent recalibration MOX gas sensor operating devices is required [11]. Insufficient sensor stability is caused by diffusion-related inter-grain mass transfer, phase transitions, surface poisoning and variable environmental conditions, such as moisture level in working media [12].
As a semiconductor material Ga2O3 materials were first proposed by Fleischer et al. to detect oxygen concentration in atmosphere [13,14]. However, gallium oxide β-Ga2O3 shows n-type semiconductor properties only at high temperatures due to band gap width about 4.9 eV, while exact value varies depending on material physical state [15]. As a result, thin gas sensing films of Ga2O3 usually show their optimum performance towards either oxidizing or reducing gases, among which are O2, NO, O3, CO, CH4 and other hydrocarbons, ammonia and volatile organic compounds, only at high temperatures over 450 °C [16]. Efforts to decrease operating temperatures of gallia-based gas sensors are directed at composite sensitive materials formation [17,18,19], decoration with noble metals, possessing catalytic properties [20,21] or exploitation of size effect of nanocrystalline high surface area Ga2O3-based materials [22,23,24]. The latter strategy may be considered as the most beneficial as the introduction of the second phase, either metal oxide or noble metal, in the materials structure can compromise the desired long-term stability of material and reliability of manufacturing process [25,26]. Doping of Ga2O3 with n-type donor dopants—Sn, Ti and Si—has been reported to improve its electrical and gas sensing properties towards both oxidating and reducing gases due to activation of oxygen surface chemisorption [27,28,29,30]. Other n-type dopants have been reported in experimental and theoretical studies to drastically improve electrical properties of Ga2O3, valuable in semiconductor industry [31,32,33]. Among them Nb(V) has been proposed as a perspective and effective n-type dopant, being a shallow donor with small ionization energy and close ionic radii to Ga(III)in both octahedral and tetrahedral positions of monoclinic β-Ga2O3 lattice [34]. Further experimental studies have supported this concept [31,35,36].
Thus, the Nb-doped β-Ga2O3 may be suggested as a very promising highly stable material for semiconductor gas sensors; however, to the best of the authors’ knowledge, the ability to the date neither thin film materials nor various ultrafine or nano-dimensional moieties has not been studied in this regard. The objective of the present study is to investigate the structural, electrical and gas sensor properties of β-Ga2O3 in ultrafine state and describe the effects of Nb(V) doping on the properties of β-Ga2O3 nano-crystals.

2. Materials and Methods

2.1. Materials Synthesis

Ga2O3/Nb materials were obtained via flame spray pyrolysis technique, thoroughly described in previous paper [37]. This synthetic approach has been chosen as it allows to obtain metal oxide materials in ultrafine state with homogeneous distribution of additional components—whether doping or surface modifying ones [38]. Toluene (99.9%) was used as a fuel. Following metalorganic precursors were used: gallium acetylacetonate and niobium 2-ethylhexanoate (V) (GELEST INC, 95%). Gallium acetylacetonate was prepared by mixing potassium acetylacetonate with gallium nitrate Ga(NO3)3·8H2O (Rare Metals Plant, Russia, 99.4%) in stochiometric ratio followed by purification with dichloromethane solution.
An amount of 0.2 M solutions of precursors in toluene were prepared with intended concentration of Nb (V) cation: 0, 1, 2 and 4 mol.% of common cation quantity. The liquid mixture was supplied to spray nozzle at 3 mL/min rate. Mixture was sprayed with oxygen (99.9%) flow of 1.5 L/min at pressure drop of 3 atm. To ignite the aerosol a circular methane/oxygen flame was used. The materials were collected on glass fiber filters (GE Whatman, Sigma-Aldrich, St. Louis, MO, USA), located 80 cm above nozzle with the aid of vacuum pump ISP 250 C (Anest Iwata, Yokohama, Japan). Powders were collected from filter manually and heated in tubular furnace at specified temperatures (500–1000 °C) for 24 h in dry clean air flow (300 mL/min). Annealed samples were characterized with the set of methods and used gas for sensor fabrication. Obtained materials (Table 1) were labeled as Ga2O3/Nb_x_y, where “x” and “y” represent given Nb(V) content and annealing temperature (°C) respectively (absence of “y” corresponds to as prepared sample).

2.2. Materials Characterization

Phase composition was studied by X-ray diffraction (XRD) using DRON-4M (Burevestnik, St. Petersburg, Russia) diffractometer with Cu Kα radiation (λ = 1.5406 Å) in 2θ = 5°–80° with 0.05° 2θ step. Particle sizes were calculated from Scherrer formula in spherical particles assumption. Specific surface area was measured by low temperature N2 adsorption (BET model) using Micromeritics Chemisorb 2750 (Micromeritics, Norcross, GA, USA). The morphology of the samples was characterized by using a JEOL JEM-2100 F/Cs/GIF/EDS transmission electron microscope (JEOL, Tokyo, Japan). Raman spectra were collected on i-Raman Plus spectrometer (BW Tek, Plainsboro Township, NJ, USA). Electronic structure was investigated on Perkin-Elmer Lambda-950 spectrometer in diffusion reflectance mode. Finally, samples composition was determined by X-Ray photoelectron spectroscopy (XPS) K-Alpha (Thermo Fisher Scientific, Waltham, MA, USA) and SPECS (SPECS, Berlin, Germany). Charge neutralization was applied, providing the C 1 s peak position of 285.0 eV.

2.3. Surface Reactivity Assessment

Oxidation reaction was conducted in the catalytic mode with the use of USGA-1 (Unisit, Russia) setup in order to measure surface activity of obtained materials. Carbon monoxide oxidation with oxygen was used as a model process. Samples were preliminarily heated in atmosphere of oxygen (99.9%) during 30 min at temperature of 500 °C, then cooled down to room temperature. Then, a gas mixture containing 79.6% N2, 15.9% O2 and 4.5% CO, respectively, was passed through reaction tube with samples heated with constant rate of 10 °C per minute till temperature of 900 °C. Mass-spectra during experiment were registered by means of MS7-100 mass-spectrometer (IAP RAN, St. Petersburg, Russia). Temperature of half conversion t_50 was used as a quantitative measure of surface reactivity and it was defined as:
p(t_50) = (pmax − pmin)/2,
where pmax and pmin are values, which correspond to maximum and minimum pressures of CO2 during experiment respectively.

2.4. Gas Sensor Measurements

Gas sensors were prepared by deposition of powder dispersion in binding substance α-terpineol on a heated substrate—aluminum polycrystalline wafer 2 × 2 × 0.15 mm with platinum heater and sensor electrodes as described before [37]. The binding substance was removed by the heating of the substrate at 500 °C for 24 h in the laboratory ambient air.
The experiment was conducted on laboratory made setup with ability to both control sensor temperature and measure sensor resistance [10]. Sensors were placed in a gastight flow-through PTFE chamber. Gas mixtures were prepared by diluting certified reference gas mixtures with pure dry air from a clean air generator, GCHV-2.0 (Himelectronica, Russia). To measure sensor response, gas mixture and pure air were consequently flown through sensor chamber and sensor response was calculated according to equation:
S = (R0 − R)/R0,
where R—is a sensing element resistance in the target gas presence, R0—sensor resistance in clean air. To investigate sensor properties several gas mixtures with fixed target gas concentrations were used: H2—20 ppm, CH3OH—20 ppm, acetone—20 ppm, and CH4—10,000 ppm. Three consequent experiments with one analyte concentration were conducted to get statistically significant results.

3. Results and Discussion

3.1. Materials Structure and Chemical Composition

XRD patterns (Figure 1) of synthesized materials show that β-Ga2O3 (ICDD PDF2 database card number 43-1012) is the primary phase in all samples. Admixture of GaNbO4 phase (ICDD PDF2 database card number 72-1666) is detected in samples annealed at temperatures over 900 °C with more than 2 mol.% Nb content. Solid solution boundaries were estimated by additional annealing at 800 °C, 850 °C and 1000 °C, respectively. For samples with less than 2 mol.% Nb content, no formation of an X-ray detectable GaNbO4 phase was shown up to 1000 °C.
The value of grain size derived from Scherrer formula decreases with growth of Nb content (Table 1). The obtained grain sizes correlate with particle dimensions measured on the basis of TEM images (Figure 2). TEM micrographs also indicate some agglomeration of nanometer size particulate matter. Elevation of annealing temperature leads to increase in particle size.
It can be generalized that the biggest specific surface area corresponds to samples with 1 mol.% Nb(V) content at every applied annealing temperature (Table 1). It should be noted that at 1000 °C annealing temperature, specific surface area value sharply decreases. This effect is probably connected with formation of GaNbO4 phase on particle surface: lower melting temperature of GaNbO4 (below 1700 K) [39] in comparison to Ga2O3 (above 2000 K) [40] leads to higher ion mobility and enhancement of diffusion processes.
Formation of GaNbO4 phase is supported by Raman spectroscopy data. Raman spectra (Figure 3) are given only for samples annealed at temperature over 900 °C because of intensive luminescence, due to high defect concentration in the crystal structure of materials calcined at lower temperatures.
Data obtained contains informative peaks in Raman shift range from 200 to 1200 cm−1. Peaks highlighted with red zones corresponds to valence and deformation oscillation modes in GaO4 and GaO6 polyhedrons. Peaks at 262 cm−1, 431 cm−1, 913 cm−1 should be attributed to O–Ga–O, O–Nb–O oscillation frequencies in GaNbO4 structure [39]. No presence of Nb2O5 was shown using Raman spectroscopy [41]. It is known that GaO6 octahedra distortion results in 913 cm−1 peak shift to lower frequency area [39]. Thus, it can be assumed that 830 cm−1 peak originated from NbGa replacement defect formation in strongly distorted oxygen coordination.
The chemical composition data obtained from XPS measurements are given in Table 2. As prepared samples and materials annealed at relatively low temperature of 500 °C demonstrate Nb content, which is corresponding to initial Nb loading. Some deviations may be caused by accuracy restrictions of XPS-quantitative analysis Nb and inaccuracy of commercially available precursor description as it was used without any pretreatment or additional analysis of precise Nb content. Table 2 follows the increase in the annealing temperature from 500 °C to 900 °C, which leads to the Nb content growth in the same material—Ga2O3/Nb-4. It may be speculated upon these results, that higher annealing temperature causes migration of Nb(V) cations to crystal surface with formation of GaNbO4 phase. This conclusion can be done keeping in mind that according to XRD analysis for samples annealed at temperatures below 800 °C particle radius is estimated to be less than 3 nm, therefore X-Ray photoelectron spectra for these materials describe composition of nearly whole particle for these samples. For the samples, annealed at higher temperatures, XPS data collected from surface layer is about 2 nm thick.
It is also possible to investigate fine chemical state of Ga cations using XPS data (Figure 4). Though the main Ga3d component (Eb = 20.4 eV) peak is overlapped with the O2s peak (Eb = 23.5 eV), a third component corresponding to the Ga(I) cation state is observable for samples containing niobium [42]. The ratio of XPS peak areas of Ga(I) to Ga(III) is close to one to eight. Hence formation of solid solution occurs with charge compensation due to transformation from Ga(III) to Ga(I) and without the rise of additional free charge carriers. Usually, the O1s XPS peak is used to estimate quantity of lattice and surface O-ions [43]. In this study, the quality of O1s peaks were insufficient to clearly distinguish two different forms of O2-contributions.

3.2. Optical Band Gap

Spectra of optical adsorption (Figure 5) of obtained materials consist of one or two linear regions in coordinates (F(R)·E)2 to E, usually used for direct bandgap semiconductors.
High-energy linear region corresponds to bandgap width of Ga2O3 phase, while low-energy linear region should be attributed to inner bandgap transition. Notably, a low energy linear region is not observed in the case of materials, for which the GaNbO4 phase formation with Eg equal to 3.41 eV [44] is confirmed either by XRD or Raman spectroscopy. The band gap width decreases with the growth of Nb content due to formation of the NbGa substitution defects, which create either shallow donor levels beneath the conduction band or additional oxygen vacancies levels. The modern theoretical studies indicate that the former case is more likely to be the cause of band gap narrowing [34]. Considering the gas sensing properties shallow donor levels, associated with the NbGa defects, should positively affect the free charge carrier concentration, which might be beneficial for oxygen chemisorption on the materials surface and, as a result, for the gas sensor response itself [27,28,29,30]. The increase in the annealing temperature causes the growth of the band gap to increase the levels, corresponding to the bulk β-Ga2O3 crystal.

3.3. Surface Reactivity

Increase in Nb content in the obtained materials is accompanied by the growth of the temperature of CO half-conversion (Figure 6).
The observed decrease in the materials reactivity can be connected with the decrease of the CO chemisorption active sites concentration on the surface of β-Ga2O3 grains, which are mainly Ga3+ cations with unsaturated coordination environment [45]. The access of gas molecules to these centers is blocked by newly formed Nb-rich GaNbO4 phase particles. There is scarce evidence of this phase’s reactivity in the heterogeneous chemical interaction with gas molecules; however, its poor activity in photochemical processes is reported [46].

3.4. Gas Sensor Properties

The typical plot of an electrical resistance variation during a gas sensor response measurement experiment of gas sensor material is given in Figure 7.
All obtained materials become more conductive as a result of interaction with reducing gas molecules. It indicates that the sensor response is formed as a result of free electrons release during a process of chemical oxidation on the surface of metal oxide grains:
R + [O2−] → RO + 2e
where R is a reducing gas molecule, [O2−] is a mobile oxygen ion in the lattice of metal oxide nanocrystal at the surface, RO is a molecular oxidation product, e is a free electron. The Equation (3) describes the oxidation by lattice oxygen as it was reported previously for heterogeneous oxidation for gallia-based materials [47].
The gas sensing properties of the obtained series of β-Ga2O3-based materials are generalized in Figure 8.
Low concentrations of the reducing gases cause the maximum of the sensor response, in the case of the materials annealed at 900 °C with the decrease in response after further annealing at higher temperature of 1000 °C. This response maximum is based on the counteraction of two factors connected with materials structure and morphology. The first one is associated with high concentration of defects in the crystal structure of materials grains, which lead to the formation of numerous levels inside the band gap acting as an electron trap and diminishing electrical and, as a result, sensor properties of the materials [48]. These defects are proven by the intense luminescence, observed during Raman spectroscopy investigation and mentioned above. The increase of annealing temperature according to XRD data allows to improve grains crystallinity, which is accompanied by gas sensory properties enhancement. On the other hand, the increase in the annealing temperature leads to the shrinkage of the effective surface area of the materials (Table 1), negatively affecting the intensity of the “solid-gas” chemical interaction. The second factor becomes dominating, when high concentrations of reducing gases are concerned. It is illustrated by the temperature dependence of sensor response towards 1000 ppm of methane, calcined below 800 °C, which shows the maximum for materials.
The influence of Nb content on the sensor response is less obvious; however, it can be generalized that those materials with 1% mol of Nb possess the highest sensor response towards reducing gases. This phenomenon can be connected to the maximum of effective surface area in the case of these materials in each series of materials obtained at the same annealing temperature. Table 3 compares the observed resistive type sensor response to most relevant reports in literature including Ga2O3 nanowires (NWs), thin films (TF) and thick films (ThF), including composites where gallia is a main component.
Table 3 leads to the conclusion that ultrafine β-Ga2O3 possesses a lower temperature of maximum sensor response compared to other forms of this material; however, in some cases, this comparison is hard to effectuate due to very different sensor response measurement conditions. Another notion is that β-Ga2O3, either described in this study or in reports of other scholars, is better suited for VOCs detection rather than simple gases, such as hydrogen or methane. The reason is that the mechanism of sensor response relies on the mobility of lattice oxygen ions at the surface of β-Ga2O3 grains, rather than chemisorbed oxygen species [16]. This mechanism requires stronger adsorption of reducing gas molecule, which is better fulfilled in case of bigger organic molecules with polar functional groups.
As the Nb-doping performed in this study does not allow for significant altering of electrical and chemical properties of the β-Ga2O3 material due to Ga(III) to Ga(I) transition, it does not affect the sensing mechanism as well. The pure and Nb-doped materials have similar dependence of sensor response not only on the working temperature, but on the gas concentration and air humidity as well (Figure 9, Figure 10 and Figure 11).
The Nb-doped material performs slightly better in the terms of absolute sensor response; however, it has similar sensitivity—an increase in the sensor response as a result of detected gas concentration increment. The observed dependence of gas sensor response on the acetone concentration slightly deviates from the linearity in the studied concentration range in accordance with power law, established for metal oxide gas sensors, operating in a DC mode [56,57]. Additionally, it suffers from the negative humidity influence the same way the pure β-Ga2O3 does. Once humidity is introduced to the analyzed media, even in relatively small amounts,—only 2%—the sensor response drops down two times. However, a further increase in the humidity does not lead to such dramatic changes in the response value, while the working temperature of maximum sensor response slightly shifts to higher values. Keeping in mind the data on the catalytic performance of the materials reported in Section 3.3, one should conclude the higher sensor response of Ga2O3/Nb-1-900 sample to be the result of grain size and effective surface area effect.
However, considering the long-term stability of obtained materials, it should be noted that they indeed show no degradation of response during continuous exploitation (Figure 12 and Figure 13). Instead, the opposite trend is observed—an improvement in response after many days of operation.
This effect should be explained as the improvement of the gas sensing thick film layer consistency and intergrain contact number increase over the time of operation. As the XRD and TEM data has shown, the 300–500 °C temperature range is not enough to cause Ga2O3 grain growth, which may diminish the size effect and sensor response as a result. However, some minute surface diffusion processes were present during repetitious surface partial reduction and further reoxidation in the course of gas sensor response measurement towards reducing gases and VOCs. These processes led to formation of new “necks” between Ga2O3 grains, the electron transport through which is controlled by the presence of reducing gases and, thus, contributed to the sensor response growth over time [58,59]. Additional long term gas sensor measurements are needed to investigate whether these slow diffusion related processes may lead to grain and intergrain contact growth, resulting in gas sensor response degradation or the sensing porous film stabilization at some point, which is crucial for the practical application.

4. Conclusions

The use of flame spray pyrolysis process allows for β-Ga2O3 to be obtained in an ultrafine state with 6–12 nm grain size and the effective surface area more than 100 m2/g. Doping by Nb(V) during the synthetic procedure leads to both Nb incorporation in the lattice of β-Ga2O3 grains and formation of separate GaNbO4 phase if the temperature of the post-synthetic annealing goes above 800 °C. Doping of β-Ga2O3 by Nb(V) is accompanied by the charge compensation of the cationic sublattice by means of Ga(III) to Ga (I) transition. The combination of these phenomena does not allow to significantly improve electrical and gas sensor properties of ultrafine β-Ga2O3 by doping with electron-donor cation Nb(V). The observed enhancement of the β-Ga2O3-based gas sensor performance can be connected to the increase in materials surface area due to the hampering of diffusion related grains growth and agglomeration at elevated working temperatures. The highest sensor response to low concentrations of reducing gases is observed in the case of Ga2O3 doped with 1% mol Nb(V) and passed through 900 °C 24 h postsynthetic annealing. The obtained ultrafine nanocrystalline samples are most suited to detection of VOCs and allowed the working temperature to decrease and sensor response to increase compared to other forms of Ga2O3—thin films, nanowires or thick films. The manufactured sensing elements demonstrated improvement of gas sensor response during long term exploitation due to enhanced intergrain contacts and porous film integrity. Modification of the synthetic procedure is required in order to obtain Nb(V) doped Ga2O3 in ultrafine nanocrystalline state without Ga(III) to Ga (I) transition.

Author Contributions

Conceptualization, V.K. and M.A.; methodology, V.K., M.A. and A.S. (Andrei Smirnov); software, M.A.; validation, M.R., A.B. and S.M.; formal analysis, M.A.; investigation, M.A., M.T., A.A., A.S. (Alina Sagitova), A.S. (Andrei Smirnov), V.A. and V.K.; resources, M.T., A.A., A.S. (Andrei Smirnov) and V.A.; data curation, M.A.; writing—original draft preparation, M.A.; writing—review and editing, V.K. and M.R.; visualization, M.A.; supervision, M.R.; project administration, V.K.; funding acquisition, V.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Russian Science Foundation, grant number 22-19-00703, https://rscf.ru/project/22-19-00703/.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

This work was performed using the equipment of the Shared Research Center “Analytical center of deep oil processing and petrochemistry of TIPS RAS”. This work is supported by MSU, Shared Use Center “Nanochemistry and nanomaterials” and Shared Use Centre “Centre of Physical and Physicochemical Methods of Analysis and Study of the Properties and Surface Characteristics of Nanostructures, Materials, and Products” UdmFRC UB RAS”.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Tonezzer, M.; Le, D.T.T.; Iannotta, S.; Van Hieu, N. Selective discrimination of hazardous gases using one single metal oxide resistive sensor. Sens. Actuators B Chem. 2018, 277, 121–128. [Google Scholar] [CrossRef]
  2. Güntner, A.T.; Weber, I.C.; Schon, S.; Pratsinis, S.E.; Gerber, P.A. Monitoring rapid metabolic changes in health and type-1 diabetes with breath acetone sensors. Sens. Actuators B Chem. 2022, 367, 132182. [Google Scholar] [CrossRef]
  3. Connolly, R.E.; Yu, Q.; Wang, Z.; Chen, Y.-H.; Liu, J.Z.; Collier-Oxandale, A.; Papapostolou, V.; Polidori, A.; Zhu, Y. Long-term evaluation of a low-cost air sensor network for monitoring indoor and outdoor air quality at the community scale. Sci. Total Environ. 2022, 807, 150797. [Google Scholar] [CrossRef] [PubMed]
  4. Schultealbert, C.; Amann, J.; Baur, T.; Schütze, A. Measuring Hydrogen in Indoor Air with a Selective Metal Oxide Semiconductor Sensor. Atmosphere 2021, 12, 366. [Google Scholar] [CrossRef]
  5. Pineau, N.J.; Magro, L.; van den Broek, J.; Anderhub, P.; Güntner, A.T.; Pratsinis, S.E. Spirit Distillation: Monitoring Methanol Formation with a Hand-Held Device. ACS Food Sci. Technol. 2021, 1, 839–844. [Google Scholar] [CrossRef]
  6. Fonollosa, J.; Solórzano, A.; Marco, S. Chemical Sensor Systems and Associated Algorithms for Fire Detection: A Review. Sensors 2018, 18, 553. [Google Scholar] [CrossRef] [Green Version]
  7. Thai, N.X.; Tonezzer, M.; Masera, L.; Nguyen, H.; Duy, N.V.; 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. Krivetskiy, V.; Efitorov, A.; Arkhipenko, A.; Vladimirova, S.; Rumyantseva, M.; Dolenko, S.; Gaskov, A. Selective detection of individual gases and CO/H-2 mixture at low concentrations in air by single semiconductor metal oxide sensors working in dynamic temperature mode. Sens. Actuat. B Chem. 2018, 254, 502–513. [Google Scholar] [CrossRef]
  9. Liu, H.; Meng, G.; Deng, Z.; Nagashima, K.; Wang, S.; Dai, T.; Li, L.; Yanagida, T.; Fang, X. Discriminating BTX Molecules by the Nonselective Metal Oxide Sensor-Based Smart Sensing System. ACS Sens. 2021, 6, 4167–4175. [Google Scholar] [CrossRef]
  10. Krivetskiy, V.V.; Andreev, M.D.; Efitorov, A.O.; Gaskov, A.M. Statistical shape analysis pre-processing of temperature modulated metal oxide gas sensor response for machine learning improved selectivity of gases detection in real atmospheric conditions. Sens. Actuators B Chem. 2021, 329, 129187. [Google Scholar] [CrossRef]
  11. Robin, Y.; Amann, J.; Baur, T.; Goodarzi, P.; Schultealbert, C.; Schneider, T.; Schütze, A. High-Performance VOC Quantification for IAQ Monitoring Using Advanced Sensor Systems and Deep Learning. Atmosphere 2021, 12, 1487. [Google Scholar] [CrossRef]
  12. Korotcenkov, G.; Cho, B.K. Instability of metal oxide-based conductometric gas sensors and approaches to stability improvement (short survey). Sens. Actuators B Chem. 2011, 156, 527–538. [Google Scholar] [CrossRef]
  13. Fleischer, M.; Hanrieder, W.; Meixner, H. Stability of semiconducting gallium oxide thin films. Thin Solid Film. 1990, 190, 93–102. [Google Scholar] [CrossRef]
  14. Fleischer, M.; Meixner, H. Gallium Oxide Thin-Films—A New Material for High-Temperature Oxygen Sensors. Sens. Actuators B Chem. 1991, 4, 437–441. [Google Scholar] [CrossRef]
  15. Fleischer, M.; Meixner, H. Sensing Reducing Gases at High-Temperatures Using Long-Term Stable Ga2o3 Thin-Films. Sens. Actuators B Chem. 1992, 6, 257–261. [Google Scholar] [CrossRef]
  16. Zhu, J.; Xu, Z.; Ha, S.; Li, D.; Zhang, K.; Zhang, H.; Feng, J. Gallium Oxide for Gas Sensor Applications: A Comprehensive Review. Materials 2022, 15, 7339. [Google Scholar] [CrossRef]
  17. Bagheri, M.; Khodadadi, A.A.; Mahjoub, A.R.; Mortazavi, Y. Strong effects of gallia on structure and selective responses of Ga2O3–In2O3 nanocomposite sensors to either ethanol, CO or CH4. Sens. Actuators B Chem. 2015, 220, 590–599. [Google Scholar] [CrossRef]
  18. Balasubramani, V.; Ahamed, A.N.; Chandraleka, S.; Kumar, K.K.; Kuppusamy, M.R.; Sridhar, T.M. Highly Sensitive and Selective H2S Gas Sensor Fabricated with β-Ga2O3/rGO. ECS J. Solid State Sci. Technol. 2020, 9, 055009. [Google Scholar] [CrossRef]
  19. Almaev, A.V.; Chernikov, E.V.; Novikov, V.V.; Kushnarev, B.O.; Yakovlev, N.N.; Chuprakova, E.V.; Oleinik, V.L.; Lozinskaya, A.D.; Gogova, D.S. Impact of Cr2O3 additives on the gas-sensitive properties of β-Ga2O3 thin films to oxygen, hydrogen, carbon monoxide, and toluene vapors. J. Vac. Sci. Technol. A 2021, 39, 023405. [Google Scholar] [CrossRef]
  20. Schwebel, T.; Fleischer, M.; Meixner, H.; Kohl, C.D. CO-sensor for domestic use based on high temperature stable Ga2O3 thin films. Sens. Actuators B Chem. 1998, 49, 46–51. [Google Scholar] [CrossRef]
  21. Tsai, J.-H.; Niu, J.-S.; Shao, W.-C.; Liu, W.-C. Characteristics of chemiresistive-type ammonia sensor based on Ga2O3 thin film functionalized with platinum nanoparticles. Sens. Actuators B Chem. 2022, 371, 132589. [Google Scholar] [CrossRef]
  22. Afzal, A. β-Ga2O3 nanowires and thin films for metal oxide semiconductor gas sensors: Sensing mechanisms and performance enhancement strategies. J. Mater. 2019, 5, 542–557. [Google Scholar] [CrossRef]
  23. Weng, T.-F.; Ho, M.-S.; Sivakumar, C.; Balraj, B.; Chung, P.-F. VLS growth of pure and Au decorated β-Ga2O3 nanowires for room temperature CO gas sensor and resistive memory applications. Appl. Surf. Sci. 2020, 533, 147476. [Google Scholar] [CrossRef]
  24. Dela Torre, H.M.B.; Santos, G.N.C. Synthesis and Characterization of Monoclinic Gallium Oxide Nanomaterials for High-Concentration Ethanol Vapor Detection. IOP Conf. Ser. Mater. Sci. Eng. 2020, 739, 012031. [Google Scholar] [CrossRef]
  25. Korotcenkov, G.; Cho, B.K. Metal oxide composites in conductometric gas sensors: Achievements and challenges. Sens. Actuators B Chem. 2017, 244, 182–210. [Google Scholar] [CrossRef]
  26. Korotcenkov, G.; Brinzari, V.; Cho, B.K. Conductometric gas sensors based on metal oxides modified with gold nanoparticles: A review. Microchim. Acta 2016, 183, 1033–1054. [Google Scholar] [CrossRef]
  27. Frank, J.; Fleischer, M.; Meixner, H.; Feltz, A. Enhancement of sensitivity and conductivity of semiconducting Ga2O3 gas sensors by doping with SnO2. Sens. Actuators B Chem. 1998, 49, 110–114. [Google Scholar] [CrossRef]
  28. Vorobyeva, N.; Rumyantseva, M.; Platonov, V.; Filatova, D.; Chizhov, A.; Marikutsa, A.; Bozhev, I.; Gaskov, A. Ga2O3(Sn) Oxides for High-Temperature Gas Sensors. Nanomaterials 2021, 11, 2938. [Google Scholar] [CrossRef]
  29. Manandhar, S.; Battu, A.K.; Devaraj, A.; Shutthanandan, V.; Thevuthasan, S.; Ramana, C.V. Rapid Response High Temperature Oxygen Sensor Based on Titanium Doped Gallium Oxide. Sci. Rep. 2020, 10, 178. [Google Scholar] [CrossRef] [Green Version]
  30. Almaev, A.V.; Chernikov, E.V.; Kushnarev, B.O.; Yakovlev, N.N. Effect of oxygen on the properties of Ga2O3:Si thin films. J. Phys. Conf. Ser. 2019, 1410, 012201. [Google Scholar] [CrossRef]
  31. Zhou, W.; Xia, C.; Sai, Q.; Zhang, H. Controlling n-type conductivity of β-Ga2O3 by Nb doping. Appl. Phys. Lett. 2017, 111, 242103. [Google Scholar] [CrossRef]
  32. Saleh, M.; Bhattacharyya, A.; Varley, J.B.; Swain, S.; Jesenovec, J.; Krishnamoorthy, S.; Lynn, K. Electrical and optical properties of Zr doped β-Ga2O3 single crystals. Appl. Phys. Express 2019, 12, 085502. [Google Scholar] [CrossRef]
  33. Alema, F.; Seryogin, G.; Osinsky, A.; Osinsky, A. Ge doping of β-Ga2O3 by MOCVD. APL Mater. 2021, 9, 091102. [Google Scholar] [CrossRef]
  34. Peelaers, H.; Van de Walle, C.G. Doping of Ga2O3 with transition metals. Phys. Rev. B 2016, 94, 195203. [Google Scholar] [CrossRef] [Green Version]
  35. Chen, J.-X.; Li, X.-X.; Tao, J.-J.; Cui, H.-Y.; Huang, W.; Ji, Z.-G.; Sai, Q.-L.; Xia, C.-T.; Lu, H.-L.; Zhang, D.W. Fabrication of a Nb-Doped β-Ga2O3 Nanobelt Field-Effect Transistor and Its Low-Temperature Behavior. ACS Appl. Mater. Inter. 2020, 12, 8437–8445. [Google Scholar] [CrossRef]
  36. Li, R.; Deng, J.; Kong, L.; Meng, J.; Luo, J.; Zhang, Q.; Gao, H.; Yang, Q.; Wang, G.; Wang, X. Influence of Substrate Temperature on Structure and Properties of Nb-Doped β-Ga2O3 Films. J. Electron. Mater. 2022, 51, 2390–2395. [Google Scholar] [CrossRef]
  37. Krivetskiy, V.; Zamanskiy, K.; Beltyukov, A.; Asachenko, A.; Topchiy, M.; Nechaev, M.; Garshev, A.; Krotova, A.; Filatova, D.; Maslakov, K.; et al. Effect of AuPd Bimetal Sensitization on Gas Sensing Performance of Nanocrystalline SnO2 Obtained by Single Step Flame Spray Pyrolysis. Nanomaterials 2019, 9, 728. [Google Scholar] [CrossRef] [Green Version]
  38. Schimmoeller, B.; Pratsinis, S.E.; Baiker, A. Flame Aerosol Synthesis of Metal Oxide Catalysts with Unprecedented Structural and Catalytic Properties. ChemCatChem 2011, 3, 1234–1256. [Google Scholar] [CrossRef]
  39. Perfler, L.; Kahlenberg, V.; Többens, D.; Schaur, A.; Tribus, M.; Orlova, M.; Kaindl, R. Mechanical Properties, Quantum Mechanical Calculations, and Crystallographic/Spectroscopic Characterization of GaNbO4, Ga(Ta,Nb)O4, and GaTaO4. Inorg. Chem. 2016, 55, 5384–5397. [Google Scholar] [CrossRef]
  40. Yoshida, K.; Kumagai, H.; Yamane, T.; Hayashi, A.; Koyama, C.; Oda, H.; Ito, T.; Ishikawa, T. Thermophysical properties of molten Ga2O3 by using the electrostatic levitation furnace in the International Space Station. Appl. Phys. Express 2022, 15, 085503. [Google Scholar] [CrossRef]
  41. Pittman, R.M.; Bell, A.T. Raman studies of the structure of niobium oxide/titanium oxide (Nb2O5.TiO2). J. Phys. Chem. 1993, 97, 12178–12185. [Google Scholar] [CrossRef]
  42. Carli, R.; Bianchi, C.L. XPS analysis of gallium oxides. Appl. Surf. Sci. 1994, 74, 99–102. [Google Scholar] [CrossRef]
  43. 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] [PubMed]
  44. Mascarenhas, D.; Shah, V. A comparative study of optical properties of site specific doping in GaNbO4. AIP Conf. Proc. 2019, 2115, 030617. [Google Scholar] [CrossRef]
  45. Delgado, M.R.; Areán, C.O. Infrared Spectroscopic Studies on the Surface Chemistry of High-Surface-Area Gallia Polymorphs. Z. Für Anorg. Und Allg. Chem. 2005, 631, 2115–2120. [Google Scholar] [CrossRef]
  46. Devi, S.; Kelkar, S.; Kashid, V.; Salunke, H.G.; Gupta, N.M. Preparation-method-dependent morphological, band structural, microstructural, and photocatalytic properties of noble metal–GaNbO4 nanocomposites. Rsc. Adv. 2013, 3, 16817–16828. [Google Scholar] [CrossRef]
  47. Kohl, D.; Ochs, T.; Geyer, W.; Fleischer, M.; Meixner, H. Adsorption and decomposition of methane on gallium oxide films. Sens. Actuators B Chem. 1999, 59, 140–145. [Google Scholar] [CrossRef]
  48. Zimmermann, C.; Rønning, V.; Kalmann Frodason, Y.; Bobal, V.; Vines, L.; Varley, J.B. Primary intrinsic defects and their charge transition levels in Ga2O3. Phys. Rev. Mater. 2020, 4, 074605. [Google Scholar] [CrossRef]
  49. Krawczyk, M.; Suchorska-Woźniak, P.; Szukiewicz, R.; Kuchowicz, M.; Korbutowicz, R.; Teterycz, H. Morphology of Ga2O3 Nanowires and Their Sensitivity to Volatile Organic Compounds. Nanomaterials 2021, 11, 456. [Google Scholar] [CrossRef]
  50. Pohle, R.; Fleischer, M.; Meixner, H. In situ infrared emission spectroscopic study of the adsorption of H2O and hydrogen-containing gases on Ga2O3 gas sensors. Sens. Actuators B Chem. 2000, 68, 151–156. [Google Scholar] [CrossRef]
  51. Bene, R.; Pintér, Z.; Perczel, I.V.; Fleischer, M.; Réti, F. High-temperature semiconductor gas sensors. Vacuum 2001, 61, 275–278. [Google Scholar] [CrossRef]
  52. Trinchi, A.; Wlodarski, W.; Li, Y.X. Study of Ga2O3 thin film gas sensors. In Proceedings of the SENSORS 2003, Toronto, ON, Canada, 22–24 October 2003; Volume 131, pp. 133–136. [Google Scholar] [CrossRef]
  53. Fleischer, M.; Kornely, S.; Weh, T.; Frank, J.; Meixner, H. Selective gas detection with high-temperature operated metal oxides using catalytic filters. Sens. Actuators B Chem. 2000, 69, 205–210. [Google Scholar] [CrossRef]
  54. Fleischer, M.; Meixner, H. Sensitive, Selective and Stable Ch4 Detection Using Semiconducting Ga2o3 Thin-Films. Sens. Actuators B Chem. 1995, 26, 81–84. [Google Scholar] [CrossRef]
  55. Frank, J.; Fleischer, M.; Meixner, H. Gas-sensitive electrical properties of pure and doped semiconducting Ga2O3 thick films. Sens. Actuators B Chem. 1998, 48, 318–321. [Google Scholar] [CrossRef]
  56. Yamazoe, N.; Shimanoe, K. Theory of power laws for semiconductor gas sensors. Sens. Actuators B Chem. 2008, 128, 566–573. [Google Scholar] [CrossRef]
  57. Potyrailo, R.A.; Go, S.; Sexton, D.; Li, X.; Alkadi, N.; Kolmakov, A.; Amm, B.; St-Pierre, R.; Scherer, B.; Nayeri, M.; et al. Extraordinary performance of semiconducting metal oxide gas sensors using dielectric excitation. Nat. Electron. 2020, 3, 280–289. [Google Scholar] [CrossRef]
  58. Barsan, N.; Weimar, U. Conduction model of metal oxide gas sensors. J. Electroceram. 2001, 7, 143–167. [Google Scholar] [CrossRef]
  59. Chwieroth, B.; Patton, B.R.; Wang, Y. Conduction and Gas–Surface Reaction Modeling in Metal Oxide Gas Sensors. J. Electroceram. 2001, 6, 27–41. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of synthesized materials with highest Nb content and analysis of phase composition.
Figure 1. XRD patterns of synthesized materials with highest Nb content and analysis of phase composition.
Materials 15 08916 g001
Figure 2. TEM images of samples: (a) Ga2O3/Nb-0-500; (b) Ga2O3/Nb-0-900; (c) Ga2O3/Nb-1-900.
Figure 2. TEM images of samples: (a) Ga2O3/Nb-0-500; (b) Ga2O3/Nb-0-900; (c) Ga2O3/Nb-1-900.
Materials 15 08916 g002
Figure 3. Raman spectra of Ga2O3/Nb-x-900 materials. Dash line marked peaks correspond to lattice oscillations in GaNbO4 phase. Remaining peaks are attributed to valence and deformation lattice oscillations in β-Ga2O3 phase.
Figure 3. Raman spectra of Ga2O3/Nb-x-900 materials. Dash line marked peaks correspond to lattice oscillations in GaNbO4 phase. Remaining peaks are attributed to valence and deformation lattice oscillations in β-Ga2O3 phase.
Materials 15 08916 g003
Figure 4. Deconvolution of Ga3d XPS spectra of materials with different Nb content and same annealing temperature.
Figure 4. Deconvolution of Ga3d XPS spectra of materials with different Nb content and same annealing temperature.
Materials 15 08916 g004
Figure 5. (a) Kubelka–Munk plot of optical absorption of synthesized materials (b) dependence of optical band gap width on Nb content and annealing temperature.
Figure 5. (a) Kubelka–Munk plot of optical absorption of synthesized materials (b) dependence of optical band gap width on Nb content and annealing temperature.
Materials 15 08916 g005
Figure 6. Dependency of partial CO2 pressure at the reaction chamber outlet on sample temperature during CO oxidation in the presence of oxygen. Vertical lines indicate temperature of CO2 half conversion and are 503 °C, 526 °C, 564 °C and 695 °C for pure β-Ga2O3 and 1,2,4 mol.% Nb containing materials, respectively.
Figure 6. Dependency of partial CO2 pressure at the reaction chamber outlet on sample temperature during CO oxidation in the presence of oxygen. Vertical lines indicate temperature of CO2 half conversion and are 503 °C, 526 °C, 564 °C and 695 °C for pure β-Ga2O3 and 1,2,4 mol.% Nb containing materials, respectively.
Materials 15 08916 g006
Figure 7. Dependence of Ga2O3/Nb-1-500 sensor material resistance on experiment time during the process of response measurement towards 20 ppm of methanol in dry air at (a) fixed working temperature 470 °C (b) different sensor working temperatures. Designations “R” and “R0” on the Figure 7a indicate the resistance of sensitive layer in the presence of methanol 20 ppm and in the flow of clean air, respectively, which are used to calculate sensor response according to Equation (2).
Figure 7. Dependence of Ga2O3/Nb-1-500 sensor material resistance on experiment time during the process of response measurement towards 20 ppm of methanol in dry air at (a) fixed working temperature 470 °C (b) different sensor working temperatures. Designations “R” and “R0” on the Figure 7a indicate the resistance of sensitive layer in the presence of methanol 20 ppm and in the flow of clean air, respectively, which are used to calculate sensor response according to Equation (2).
Materials 15 08916 g007
Figure 8. Dependence of sensor response of Ga2O3-based materials towards reducing gases and volatile organic compounds (VOCs) on temperature in relation to materials annealing temperature and Nb content—(a) sensor response of Ga2O3-based materials doped with 1% mol Nb and annealed at different temperatures towards 20 ppm of methanol in dry air (b) sensor response of Ga2O3-based materials doped with various mol content of Nb and annealed at 900 °C temperature towards 20 ppm of methanol in dry air (c) sensor response of Ga2O3-based materials doped with 1% mol Nb and anneald at different temperatures towards 20 ppm of methanol in dry air (d) sensor response of Ga2O3-based materials doped with various mol content of Nb and annealed at 900 °C temperature towards 20 ppm of acetone in dry air (e) sensor response of Ga2O3-based materials doped with 1% mol Nb and anneald at different temperatures towards 20 ppm of hydrogen in dry air (f) sensor response of Ga2O3-based materials doped with various mol content of Nb and annealed at 900 °C temperature towards 20 ppm of hydrogen in dry air (g) sensor response of Ga2O3-based materials doped with 1% mol Nb and anneald at different temperatures towards 10,000 ppm of methane in dry air (h) sensor response of Ga2O3-based materials doped with various mol content of Nb and annealed at 900 °C temperature towards 10,000 ppm of methane in dry air.
Figure 8. Dependence of sensor response of Ga2O3-based materials towards reducing gases and volatile organic compounds (VOCs) on temperature in relation to materials annealing temperature and Nb content—(a) sensor response of Ga2O3-based materials doped with 1% mol Nb and annealed at different temperatures towards 20 ppm of methanol in dry air (b) sensor response of Ga2O3-based materials doped with various mol content of Nb and annealed at 900 °C temperature towards 20 ppm of methanol in dry air (c) sensor response of Ga2O3-based materials doped with 1% mol Nb and anneald at different temperatures towards 20 ppm of methanol in dry air (d) sensor response of Ga2O3-based materials doped with various mol content of Nb and annealed at 900 °C temperature towards 20 ppm of acetone in dry air (e) sensor response of Ga2O3-based materials doped with 1% mol Nb and anneald at different temperatures towards 20 ppm of hydrogen in dry air (f) sensor response of Ga2O3-based materials doped with various mol content of Nb and annealed at 900 °C temperature towards 20 ppm of hydrogen in dry air (g) sensor response of Ga2O3-based materials doped with 1% mol Nb and anneald at different temperatures towards 10,000 ppm of methane in dry air (h) sensor response of Ga2O3-based materials doped with various mol content of Nb and annealed at 900 °C temperature towards 10,000 ppm of methane in dry air.
Materials 15 08916 g008
Figure 9. Dependence of sensor response of (a) Ga2O3-based materials without Nb doping annealed at 900 °C and (b) Ga2O3-based materials with 1% mol Nb doping annealed at 900 °C at different working temperatures towards acetone on concentration in dry air.
Figure 9. Dependence of sensor response of (a) Ga2O3-based materials without Nb doping annealed at 900 °C and (b) Ga2O3-based materials with 1% mol Nb doping annealed at 900 °C at different working temperatures towards acetone on concentration in dry air.
Materials 15 08916 g009
Figure 10. Dependence of sensor response of Ga2O3-based materials at 470 °C working temperature towards acetone on its concentration in dry air.
Figure 10. Dependence of sensor response of Ga2O3-based materials at 470 °C working temperature towards acetone on its concentration in dry air.
Materials 15 08916 g010
Figure 11. Dependence of sensor response of (a) pure Ga2O3-based materials annealed at 900 °C and (b) 1% mol Nb-doped Ga2O3-based materials annealed at 900 °C at different working temperatures towards acetone 20 ppm in dry air and in air with relative humidity (RH) 2%, 30% and 60%.
Figure 11. Dependence of sensor response of (a) pure Ga2O3-based materials annealed at 900 °C and (b) 1% mol Nb-doped Ga2O3-based materials annealed at 900 °C at different working temperatures towards acetone 20 ppm in dry air and in air with relative humidity (RH) 2%, 30% and 60%.
Materials 15 08916 g011
Figure 12. Dependence of sensor response of (a) pure Ga2O3-based materials annealed at 900 °C and (b) 1% mol Nb-doped Ga2O3-based materials annealed at 900 °C at different working temperatures towards acetone 20 ppm in dry air at different days after manufacturing and start of gas sensor response study. The sensors were busy in the 300–500 °C working temperature range almost every day since the first day of operation.
Figure 12. Dependence of sensor response of (a) pure Ga2O3-based materials annealed at 900 °C and (b) 1% mol Nb-doped Ga2O3-based materials annealed at 900 °C at different working temperatures towards acetone 20 ppm in dry air at different days after manufacturing and start of gas sensor response study. The sensors were busy in the 300–500 °C working temperature range almost every day since the first day of operation.
Materials 15 08916 g012
Figure 13. Dependence of sensor response of Ga2O3-based materials at fixed working temperature 470 °C towards acetone 20 ppm in dry air at different days after manufacturing and start of gas sensor response study.
Figure 13. Dependence of sensor response of Ga2O3-based materials at fixed working temperature 470 °C towards acetone 20 ppm in dry air at different days after manufacturing and start of gas sensor response study.
Materials 15 08916 g013
Table 1. Morphological parameters of the obtained materials—grain size, nm/surface area—m2/g.
Table 1. Morphological parameters of the obtained materials—grain size, nm/surface area—m2/g.
Annealing Temperature, °C (y) Nb Content, mol.% (x)
0124
-6/976/1116/926/109
5006/1046/1276/1026/103
7007/937/987/956/88
8009/638/737/657/67
90012/2411/4910/4111/34
100015/1119/1--
Table 2. Nb content in the synthesized samples according to XPS.
Table 2. Nb content in the synthesized samples according to XPS.
SampleNb Content, mol.% (x)
Ga2O3/Nb-0-5000
Ga2O3/Nb-1-5001.6
Ga2O3/Nb-2-5002.6
Ga2O3/Nb-44.2
Ga2O3/Nb-4-5003.9
Ga2O3/Nb-4-8005.4
Ga2O3/Nb-4-9008.7
Table 3. Sensor response of various Ga2O3-based materials towards gases and VOCs *.
Table 3. Sensor response of various Ga2O3-based materials towards gases and VOCs *.
MaterialGasTw, °CS, a.u.Reference
Ga2O3/Nb-1-900Acetone, 20 ppm4709.8This work
Ga2O3 NWsAcetone, 100 ppm60040[49]
Ga2O3 ThFAcetone, 200 ppm48035[50]
Ga2O3 TFAcetone, 1600 ppm530100[51]
Ga2O3/Nb-1-900H2, 20 ppm5002.5This work
Ga2O3-Cr2O3 TFH2, 2500 ppm50060[19]
Ga2O3 ThFH2, 5000 ppm5001[52]
H2, 20 ppmH2, 20 ppm7002[53] **
Ga2O3/Nb-1-500CH4, 10,000 ppm5004This work
Ga2O3 TFCH4, 5000 ppm840100[54] **
Ga2O3 ThFCH4, 10,000 ppm57010[55] **
* No methanol sensing data with Ga2O3 resistive sensors were found by authors in the literature. ** The most relevant studies of many publications by M. Fleischer et al. are cited considering used H2 and CH4 concentrations, measurement conditions and studies similarity.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Andreev, M.; Topchiy, M.; Asachenko, A.; Beltiukov, A.; Amelichev, V.; Sagitova, A.; Maksimov, S.; Smirnov, A.; Rumyantseva, M.; Krivetskiy, V. Electrical and Gas Sensor Properties of Nb(V) Doped Nanocrystalline β-Ga2O3. Materials 2022, 15, 8916. https://doi.org/10.3390/ma15248916

AMA Style

Andreev M, Topchiy M, Asachenko A, Beltiukov A, Amelichev V, Sagitova A, Maksimov S, Smirnov A, Rumyantseva M, Krivetskiy V. Electrical and Gas Sensor Properties of Nb(V) Doped Nanocrystalline β-Ga2O3. Materials. 2022; 15(24):8916. https://doi.org/10.3390/ma15248916

Chicago/Turabian Style

Andreev, Matvei, Maxim Topchiy, Andrey Asachenko, Artemii Beltiukov, Vladimir Amelichev, Alina Sagitova, Sergey Maksimov, Andrei Smirnov, Marina Rumyantseva, and Valeriy Krivetskiy. 2022. "Electrical and Gas Sensor Properties of Nb(V) Doped Nanocrystalline β-Ga2O3" Materials 15, no. 24: 8916. https://doi.org/10.3390/ma15248916

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