Next Article in Journal
Adsorption and Sensing Properties of Ni-Modified InSe Monolayer Towards Toxic Gases: A DFT Study
Next Article in Special Issue
CO2 Interaction Mechanism of SnO2-Based Sensors with Respect to the Pt Interdigital Electrodes Gap
Previous Article in Journal
Tailoring Ruthenium(II) and Rhenium(I) Complexes for Turn-On Luminescent Sensing of Antimony(III)
Previous Article in Special Issue
Low-Drift NO2 Sensor Based on Polyaniline/Black Phosphorus Composites at Room Temperature
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Novel Mechanism Based on Oxygen Vacancies to Describe Isobutylene and Ammonia Sensing of p-Type Cr2O3 and Ti-Doped Cr2O3 Thin Films

1
Key Laboratory of Special Functional Materials for Ecological Environment and Information, Hebei University of Technology, Ministry of Education, Tianjin 300130, China
2
Department of Chemistry, University College London, London WC1H 0AJ, UK
3
School of Resources and Civil Engineering, Northeastern University, Shenyang 110819, China
4
School of Engineering, University of Warwick, Coventry CV4 7AL, UK
5
NosmoTech Ltd., Cambridge CB3 0AZ, UK
6
Advanced Sensing Technologies Ltd., London N12 8FA, UK
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Chemosensors 2024, 12(10), 218; https://doi.org/10.3390/chemosensors12100218
Submission received: 11 September 2024 / Revised: 15 October 2024 / Accepted: 16 October 2024 / Published: 18 October 2024
(This article belongs to the Special Issue Advanced Chemical Sensors for Gas Detection)

Abstract

:
Gas sensors based on metal oxide semiconductors (MOS) have been widely used for the detection and monitoring of flammable and toxic gases. In this paper, p-type Cr2O3 and Ti-doped Cr2O3 (CTO) thin films were synthesized using an aerosol-assisted chemical vapor deposition (AACVD) method. Detailed analysis of the thin films deposited, including structural information, their elemental composition, oxidation state, and morphology, was investigated using XRD, Raman analysis, SEM, and XPS. All the gas sensors based on pristine Cr2O3 and CTO exhibited a reversible response and good sensitivity to isobutylene (C4H8) and ammonia (NH3) gases. Doping Ti into the Cr2O3 lattice improves the response of the CTO-based sensors to C4H8 and NH3. We describe a novel mechanism for the gas sensitivity of p-type metal oxides based on variations in the oxygen vacancy concentration.

1. Introduction

Volatile organic compounds (VOCs), including ammonia and hydrocarbons, are ubiquitous, and their detection has been employed for the monitoring of food and beverages, agricultural produce, pharmaceuticals, and health [1]. Certain VOCs are highly toxic and/or carcinogenic and may cause both short- and long-term health issues (e.g., allergies or cancer) as well as impact our ecosystem [2]. Even though the human nose serves as a highly sensitive sensing system, it still fails when low-concentration or odorless toxic gases—which pose a serious threat to human health—need to be detected [3,4], so the development of a low-cost, simple-to-operate VOC sensor is of great current interest.
Gas sensors based on metal oxide semiconductors (MOSs) are commercially available and have been widely used in the detection and monitoring of flammable and toxic gases [5]. MOS-based sensors are cheap, reliable, consume relatively low power, and have good sensitivity, due to a wide conductivity range and a robust structure, making them an ideal gas-sensing material [6]. Compared to the commercially successful n-type SnO2 and WO3 sensors, p-type MOS-based sensors, such as Cr2O3, CuO, and NiO, have received relatively little attention [7]. As a typical p-type semiconductor, Cr2O3 has been used in gas sensing [8], catalysis [9], and electrochemical devices [10]. It is reported that doping titanium into Cr2O3 is an effective method to enhance the gas-sensing performance of pristine Cr2O3 [11]. Ti-doped Cr2O3 (CTO) MOSs are also good p-type materials for use in gas sensors, due to their tolerance towards humidity, good baseline stability, and reasonable sensitivity. CTO-based sensors are especially good for the detection of trace quantities of reducing gases in air (such as H2S, CO, or ethanol vapor), with operating temperatures ranging from 300 to 500 °C [12,13]. Shaw [14] reported the deposition of Cr2−xTixO3 thin films on sensor substrates using a variety of techniques—including screen printing, atmospheric pressure chemical vapor deposition (APCVD), and flame fusion methods—and investigated their gas response to CO and ethanol. Du [15] used electrostatic spray- assisted vapor deposition (ESAVD) at 650 °C to prepare Cr2−xTixO3 films for the detection of NH3. The results showed that the sensitivity of Cr1.8Ti0.2O3 films to 500 ppm NH3 at 500 °C was around 1.45. To further investigate the effect of different Ti-doping levels on Cr2−xTixO3, Du [16] carried out the synthesis and characterization of Cr2−xTixO3 with different nominal compositions using the ESAVD technique. Of all the CTO sensors—prepared with varying compositions—Cr1.7Ti0.3O3 exhibited the highest gas sensitivity. This sensor exhibited a gas response of 2.90 when exposed to 500 ppm NH3 at 200 °C and also revealed good sensitivity to ethanol vapor at 400 °C. Conde-Gallardo [17] found that the conductivity of Cr2−xTixO3 films deposited by aerosol-assisted CVD (AACVD), using copper acetylacetonate and titanium butoxide precursors dissolved in isopropanol, was dependent on the degree of Ti-substitution (conductivity decreased with increasing Ti), although they did not measure the gas-sensing properties.
AACVD has been widely implemented to prepare homogenous and uniform thin films and can effectively generate high-purity MOSs [18]. It has a key advantage over alternative methods of sensor material synthesis in that it combines both material synthesis and device integration in a single step. The AACVD process has attracted attention over other CVD processes due to its characteristics of easy operation, low cost, higher deposition rates, and flexibility in tuning thin-film micro/nanostructures [19].
This work focused on the synthesis of Cr2O3 and Ti-doped Cr2O3 thin films via the AACVD method and demonstrated the relative sensitivity of the materials to the gases ammonia (NH3) and C4H8. NH3 is a colorless gas with corrosive properties and a strong pungent odor. Severe damage is caused to the throat, lungs, eyes, and skin, even at low NH3 concentrations [20]. Moreover, it is reported that exposure to an NH3 atmosphere exceeding 50 ppm for a long time leads to serious pathological changes in organs such as the liver and kidneys [21] and also results in serious health issues, including respiratory distress, eye irritation, and skin problems [22]. NH3 is also a metabolite in breath exhaled from the human body, which can be used as a marker of end-stage renal disease (ESRD) for a non-destructive diagnosis in clinical medicine (average 4.88 ppm; range 0.82–14.7 ppm) [23]. As far as we know, this report is also the first time that CTO sensors have been used to detect C4H8, a model alkene VOC. The chemical composition and morphological structure of the Cr2O3 and CTO thin films were characterized using XRD, Raman analysis, SEM, and XPS. To understand how the conductivity of CTO may change upon exposure to these gases, we present the first description of a gas-sensing mechanism for p-type MOSs, based on varying concentrations of oxygen vacancies in the sensing material.

2. Experimental Section

2.1. Synthesis of Cr2O3 and CTO Thin Films

The deposition of Cr2O3 and CTO thin films was achieved via the AACVD technique in a homemade reactor equipped with a water-cooling system in the proximity of the inlet to minimize the pre-reaction of precursors. A schematic of the AACVD technique is shown in Figure 1.
To prepare the Cr2O3 films, alumina gas-sensor platforms (with interdigitated screen-printed gold electrodes) were first cleaned with acetone and deionized water before use and were then set onto a base and covered with a mask. These were then placed into the reactor and heated to 340 °C. A solution of chromium hexacarbonyl (0.1 g, 0.454 mmol) in methanol (40.0 mL) was prepared and aerosols of the solution were formed using an ultrasonic humidifier operated at 2 MHz, with the aerosols then transported to the reactor by a N2 gas carrier at a flow rate of 1000 standard cubic centimeters per minute (sccm). After ~10 cm3 of the precursor solution had been transported, the substrate and mask were rotated 90° and the deposition process continued; this was repeated four times to ensure homogeneous coverage of the thin films across the sensors. Once the precursor solution was exhausted, the heating temperature and flow rate were reduced to 100 °C and 300 sccm, respectively. After deposition, the dark green thin films obtained were annealed for 24 h at variable temperatures and, respectively, named as Cr2O3-1, Cr2O3-2, and Cr2O3-3, corresponding to the annealing temperatures of 500, 600, and 700 °C.
For the synthesis of CTO thin films, a solution of 0.22 g (0.775 mmol) titanium diisopropoxide bis-(acetylacetonate) (TDBAA) and 15.0 mL methanol was prepared, then 3.0 cm3 of this TDBAA–methanol solution was added to the chromium solution (as previously prepared) and stirred to obtain a precursor solution with a Cr:Ti molar ratio of Cr1.5Ti0.5. Aerosols of this solution were again generated via a humidifier and transported by a N2 gas carrier, but this time at a flow rate of 1500 sccm (to provide more homogeneous Ti-incorporation). For the CTO-based sensors, the annealing temperature was maintained at 600 °C and the annealing time varied; CTO-1 is the sample annealed for 6 h, while CTO-2 is the sample annealed for 30 h.

2.2. Characterization of the Samples

X-ray diffraction (XRD) patterns were analyzed using a Bruker, LinxEye D8 X-ray diffractometer in reflection mode, using Cu Kα radiation (λ = 1.5406 Å) and operated at 50.0 kV and 1.0 mA. Scans were performed using a detection angle ranging from 20.0° to 70.0°. Raman spectroscopy analysis was carried out using a Renishaw 1000 spectrometer equipped with a 532 nm laser. The Raman system was calibrated using a silicon reference. All films were placed in the spectrometer using an X-Y stage and analyzed in the ranges of 200 to 800 cm−1, with a laser power of 10%, an exposure time of 45 s, and an accumulation of three times per sample. Scanning electron microscopy (SEM) was conducted using a Jeol 6310F microscope. All sample images were collected using a secondary electron detector. The samples were coated with a thin layer of sputtered gold and connected to the metal stage with copper tape. Energy dispersive analysis of the X-rays was conducted using the same instrument to determine the sample composition (for the samples coated with carbon, not gold). X-ray photoelectron spectroscopy (XPS) measurements were performed using a Thermo Kα spectrometer with monochromatic Al Kα radiation, a dual-beam charge compensation system, and a constant pass energy of 50 eV. The binding energies were calibrated with respect to the C 1s peak at 284.6 eV.

2.3. Sensor Fabrication and Gas-Sensing Measurements

A photograph of the MEMS platform is shown in Figure 2a. Once the AACVD-deposited substrates were made, platinum wires were used to connect the alumina substrates to a pin stage, as shown in Figure 2b. Four platinum wires were used for the electrical connections; two were welded to the gold sensor trackpads on the top side of the sensor platform, and the other two were connected to the platinum heater trackpads on the bottom side. Gas-sensing measurements of the as-prepared Cr2O3 and CTO sensors were tested at Alphasense Ltd., UK. The working temperature and resistance of the sensors were measured and controlled by a Sensor Management System (SMS), which is designed to work with up to eight metal oxide gas sensors. It has circuits to accurately control the sensor heater temperature based on a constant-resistance setup with a digital control. The variance in the concentration of the tested gas was obtained by altering the flow rate of each gas using a mass flow controller (MFC, UFC 1100, Brooks), which was controlled by a computer program written in LabVIEW (National Instrument 2016). In a typical test cycle, the enclosed system containing the sensors was first purged with 50% humid air for 30 min; afterwards, a flow of concentrated gas analyte was passed through for 30 min, and then a flow of just humid air passed for 30 min again to clean the analyte.

3. Results and Discussion

3.1. Material Characterization

EDX analysis gave the composition of the samples as Cr1.95Ti0.05 (97.5% Cr, 2.5% Ti on a metals basis), indicating decreased incorporation of Ti into the film relative to the concentration in the precursor solution; this is commonly observed in multicomponent films deposited by AACVD and occurs due to differences in the thermal decomposition behaviour of the individual precursors. The XRD patterns of Cr2O3 and CTO-based sensors are shown in Figure 3, as shown in Figure 3a, the characteristic peaks at 2θ = 24.5°, 33.6°, 36.2°, 41.2°, 50.2°, 54.9°, and 57.1° can be assigned to the (012), (104), (110), (113), (024), (116), and (211) planes of hexagonal phase eskolaite (JCPDS # 38-1479) [24]. The XRD patterns of corundum indicate that all of the diffraction peaks match well with the hexagonal phase Al2O3 (JCPDS # 46-1212), which is expected as the substrates that are made from alumina. Notably, the sharp diffraction peaks of gold (JCPDS # 04-0784) were also detected that come from the ink used for the sensing electrodes. No obvious 2θ peak shift was observed when Cr2O3 thin films were annealed at different temperatures (500, 600, and 700 °C) on the surface of alumina sensing platform. Figure 3b exhibits the XRD patterns of CTO thin films and the results indexed to the crystal structure of eskolaite (Cr2O3), corundum (Al2O3), and gold. The CTO patterns are in close alignment with the deposited Cr2O3 thin films and the result shows that no measurable change in the crystal structure is observed and peaks of TiO2 are not present, which indicates that the full incorporation of Ti dopant into the lattice of Cr2O3 and no phase separation occurred within the limit of detection of XRD. Furthermore, no other impurity peaks were detected in all of the XRD patterns, demonstrating the high purity of the samples.
The average crystal sizes of the Cr2O3 and CTO thin films were calculated using Scherrer’s formula:
D = k λ / β cos θ .
Herein, D is the average crystal size, and all angles are in radians. For spherical crystallites, in most cases k is a constant of about 0.9. λ represents the X-ray wavelength (1.5406Å), and β is the full width at half maximum (FWHM) of the peaks of the Cr2O3 and CTO thin films. For the Cr2O3 thin films, the calculated values of the crystallite sizes are 26, 31, and 30 nm, corresponding to the different annealing temperatures of 500, 600, and 700 °C. From the results, it can be seen that the crystallite sizes may become a little bigger with the increase in annealing temperatures, but the difference is likely to be insignificant within the precision of the Scherrer estimate. In the CTO thin films, the crystallite sizes of CTO-1 and CTO-2 are 25 and 36 nm, respectively, demonstrating that the incorporation of Ti into Cr2O3 crystal causes no obvious influence on its crystallite size (CTO-1), but extending the annealing time from 6 h to 30 h appears to contribute to a growth in crystallite size (CTO-2). Table 1 summarizes the preparation conditions and Scherrer sizes of the different sensors.
The crystal structures were also investigated using Raman spectroscopy. Figure 4 displays the Raman spectra of the Cr2O3 and CTO thin films. In Cr2O3, there are a total of four Raman modes observed (three Eg modes and one A1g mode). Pristine Cr2O3 thin films exhibited three Eg vibration modes at 303, 351, and 612 cm−1 and one intense A1g mode at 553 cm−1. The absence of characteristic TiO2 Raman peaks for the CTO sample correlates well with the results from XRD characterization—namely, there was no phase separation on the film. The broad peak located at 721 cm−1 has been seen previously and has been ascribed to the existence of local vibration [25].
The morphologies of the as-prepared CTO thin films were characterized using SEM. As seen from the high-resolution picture (Figure 5a), the CTO thin films consisted of many uniform and well-dispersed sub-micron particles, which are spherical-like, with a diameter of 300–500 nm. It can be speculated that the particles observed in SEM are agglomerates of many smaller crystallites, due to the disparity between the particle size (SEM) and the crystallite size (Scherrer). As shown in the low-magnification SEM image (Figure 5b), the CTO particles are self-assembled to form a lamellar structure. SEM images of the Cr1.95T0.05O3 films previously deposited (at higher temperatures) by APCVD exhibited dense and spheroidal platelets sized 1–4 μm [14], whilst the Cr1.8Ti0.2O3 films deposited by the ESAVD technique had a more ordered porous structure, due to the introduction of polymeric additives [15].
The surface chemical composition of the as-synthesized CTO was analyzed by XPS and the results are shown in Figure 6. The XPS surface analysis was conducted using C 1s as calibration at 284.6 eV. It can be observed that the sharp peaks in the survey spectrum (Figure 6a) can be attributed to the elements of C, Ti, O, and Cr, demonstrating the successful doping of Ti into the Cr2O3 thin film. The high-resolution spectrum of Ti4+ 2p (shown in Figure 6b) is split into two peaks—of Ti4+ 2p1/2 and Ti4+ 2p3/2—centered at 464.5 and 458.6 eV. This is in contradiction with the work of Conde-Gallardo et al. [17], who found that their Cr2−xTixO3 films deposited by AACVD featured Ti3+—although substitution by Ti4+ would be consistent with the decrease in conductivity they observed with increasing Ti incorporation, due to electronic compensation through the formation of electrons (which are expected to recombine with the native p-type carriers in Cr2O3). The Cr 2p spectrum of CTO in Figure 6c could be divided into two peaks of Cr3+ 2p1/2 and Cr3+ 2p3/2, located at 586.6 and 576.7 eV, respectively. Figure 6d shows that the XPS spectrum of O 1s is deconvoluted into three peaks, corresponding to Ti–O–Ti (534.2 eV), surface oxygen species (-COx, -OH) (532.3 eV), and Cr–O–Cr (530.4 eV), respectively [26].

3.2. Gas Response

Figure 7a,b show the dynamic responses of the Cr2O3 and CTO sensors towards C4H8 gas, respectively. The test conditions were set at two different concentrations of C4H8 gas at a relative humidity of 50% and operating temperatures of 400 and 450 °C. Each temperature was tested with three pulses of C4H8 gas (20-5-20 ppm). Figure 8a,b show the sensors response (resistance in gas over resistance in air, Rg/Ra) towards the different concentrations of C4H8 at different temperatures. The dynamic response results (Figure 7) show that the response values remained stable after the continuous test with different gas concentrations at different working temperatures. Each pulse towards the analyte was 30 min, then followed by 30 min of humid air. It can be observed that resistance values are higher for CTO as compared to resistance values of Cr2O3, as expected for substitution of Cr3+ with Ti4+, again suggesting that Ti was fully incorporated into the host lattice rather than as a separate phase. The CTO sensors also exhibited a typical p-type response upon exposure to C4H8 gas. All the CTO sensors prepared by AACVD show a higher response than that of the Cr2O3 sensors (Figure 8), although the CTO-1 sensor annealed for 6 h displayed a lower response compared to the one annealed for 30 h (CTO-2).
All the sensors were then exposed to NH3, at the same relative humidity and operating temperature as the C4H8 gas but using different concentrations (75-25-75 ppm), and the results are shown in Figure 9. Obviously, the resistance values of the sensors also all increase upon exposure to NH3, again showing the typical p-type property associated with Cr2O3/CTO. As previously, the Cr2O3-2 film annealed at 600 °C shows the largest response of the three Cr2O3-based sensors (Figure 8a). After doping Ti into the Cr2O3 thin films, the as-prepared CTO sensors exhibited a higher response to NH3 compared to the Cr2O3-based sensors (Figure 8b). The highest response for Cr2O3 at 400 °C and 75 ppm NH3 is about 1.09, whereas the response for CTO under the same conditions is about 1.45. It is noteworthy that the annealing time influences the response of the CTO sensors to C4H8 and NH3 gases.

3.3. Sensing Mechanism

In general, electron depletion theory is the most widely accepted sensing mecha-nism for the majority of MOSs [27]. However, an alternative self-consistent description of electronic change mediated by oxygen vacancies, rather than ionosorbed oxygen species, has been postulated for n-type MOS [28,29,30]. A similar mechanism for p-type materials is currently missing, despite the commonly employed electron depletion ex-planations for p-type materials requiring the surface oxygen acceptor species to be at vastly different energy to those expected in n-type materials, which must be close to the conduction band maximum in order for their adsorption and desorption to be thermodynamically accessible, for the descriptions to work (in fact below the valence band maximum), despite there being no rationalization for such a difference to exist. However, in direct contrast to n-type materials, oxygen vacancies found in p-type ma-terials are typically ‘deep’ in the bandgap (further from the conduction band edge than in n-type). Consequently, they will not be ionized at normal sensor operating temper-atures (Equation (2), the electrons arising from formation of the oxygen vacancy are not ionized to free electrons but rather are ‘trapped’ at the vacancy). Due to the pres-ence of the (relatively) numerous (majority carrier) holes in p-type materials it is ex-pected to be thermodynamically favorable for the electrons trapped at oxygen vacan-cies (within the bandgap) to recombine with holes (in the valence band). This means formation of oxygen vacancies would reduce the concentration of primary carriers in a p-type material (Equation (3)), hence leading to a decrease in conductivity (increase in resistance) with increasing oxygen vacancy concentration (a similar explanation ex-plains the observed increase in Cr2O3 sensor resistance with Ti(IV)-doping (CTO); to maintain electroneutrality for every two Ti(IV) substituted onto a Cr(III) site an oxygen vacancy must be formed).
2 O O   O 2 + 2 V O
2 V O + 4 h · 2 V O x .
Therefore, under exposure to reducing conditions, the hole concentration is expected to decrease [31] and hence the resistance is expected to increase, at least at the surface of the material. As shown in Figure 10, exposure of Cr2O3 or CTO to NH3 or C4H8 would cause the surface lattice oxygen concentration (OO) to decrease as the lattice oxygen is consumed by oxidizing the gas species (R), i.e., the oxidation of the analyte gas causes the oxygen vacancy concentration in the sensing material to increase (VO), which annihilates holes (4) and causes the resistance to rise, as observed in our tests.
2 O O + 4 h · + 2 R 2 R O + 2 V O x
Exposure to pure air would allow the surface to re-oxidize, reducing the concentration of oxygen vacancies and hence decreasing the (surface) resistance.
Assuming that the increases in resistance observed upon introduction of the analyte gas are due to their reaction with lattice oxygen, rather than some form of direct donor (analyte)/acceptor (sensor) behavior—which seems reasonable given that the strength of any donor interaction at the elevated temperatures used here (400–450 °C) must be questionable—the relative sensitivity towards these different reducing gases would then relate either to the different thermodynamics (for a steady-state ‘saturated’ response) or kinetics (for a transient response) for the oxidation of the gas species by lattice oxygen, i.e., the reactivity of a given analyte species towards oxidation. Whilst we do not have a direct comparison available in our test data, a comparison of the change in resistance of our CTO sensors towards 20 ppm C4H8 and 25 ppm NH3 at 450 °C (closest to steady-state response) suggests that it is thermodynamically more favorable for C4H8 to be oxidized by CTO (either due to the ease of oxidation or due to a more complete oxidation) at an elevated temperature (higher resistance change observed for C4H8 at a similar concentration).
In addition, the difference in the sensing performance between Cr2O3 and CTO (e.g., Figure 8) may then be attributed to the difference in the relative change in oxygen vacancy concentration between the two materials (the lower initial hole concentration in CTO, displayed schematically in Figure 10, with each Ti inducing the loss of one hole, meaning a greater relative change is measured)—although without additional analysis techniques, it cannot be discounted that the presence of Ti has a direct (e.g., catalytic) effect on the reaction.

4. Conclusions

In this work, Cr2O3 and CTO thin films, from [Cr(CO)6] and TBDAA precursors in a methanol solution, were deposited on the surface of an alumina platform using AACVD. Except for the patterns coming from the platform, XRD patterns for the CTO sensors only displayed the crystalline phase of eskolaite (Cr2O3), which shows no indication of phase separation. XPS analysis confirmed the presence of Ti4+ in the CTO thin films. Doping with Ti4+ increased the resistance, characteristic of an n-type dopant. All the as-prepared sensors based on Cr2O3 and CTO thin films exhibited good sensitivity and reproducibility in response to isobutylene and ammonia gases at a humidity of 50% RH. We have postulated a new mechanism for gas sensitivity in p-type metal oxides based on a varying oxygen vacancy concentration.

Author Contributions

Conceptualization, P.Z. and J.-H.T.; methodology, C.B.; validation, J.A.C., J.S. and E.D.; formal analysis, J.L.; investigation, J.-H.T.; resources, C.B.; data curation, P.Z.; writing—original draft preparation, P.Z.; writing—review and editing, P.Z. and C.B.; visualization, Y.S.; supervision, C.B.; project administration, C.B.; funding acquisition, C.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Engineering and Physical Sciences Research Council (grant number EP/R512400/1), the China Scholarship Council (grant number 202006080083), and the Natural Science Foundation of Hebei Province (grant number E2023202041).

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 author.

Acknowledgments

The authors thank the Engineering and Physical Sciences Research Council (EP/R512400/1), the China Scholarship Council (CSC, No. 202006080083), and the Natural Science Foundation of Hebei Province (No. E2023202041) for providing financial support. The authors would also like to thank Alphasense Ltd. for packaging the sensors.

Conflicts of Interest

Author John Saffell employed by the company Nosmotech Ltd.; Author Ehsan Danesh employed by the company Advanced Sensing Technologies Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Ayyala, S.K.; Covington, J.A. Nickel-Oxide Based Thick-Film Gas Sensors for Volatile Organic Compound Detection. Chemosensors 2021, 9, 247. [Google Scholar] [CrossRef]
  2. Agbroko, S.O.; Covington, J. A novel, low-cost, portable PID sensor for the detection of volatile organic compounds. Sens. Actuators B 2018, 275, 10–15. [Google Scholar] [CrossRef]
  3. Ding, C.; Ma, Y.; Lai, X.; Yang, Q.; Xue, P.; Hu, F.; Geng, W. Ordered Large-Pore Mesoporous Cr2O3 with Ultrathin Framework for Formaldehyde Sensing. ACS Appl. Mater. Interfaces 2017, 9, 18170–18177. [Google Scholar] [CrossRef] [PubMed]
  4. Jin, H.; Huang, Y.; Jian, J. Plate-like Cr2O3 for highly selective sensing of nitric oxide. Sens. Actuators B 2015, 206, 107–110. [Google Scholar] [CrossRef]
  5. Zhang, M.; Sui, N.; Wang, R.; Zhang, T. The effect of shell thickness on gas sensing properties of core-shell fibers. Sens. Actuators B 2021, 332, 129456. [Google Scholar] [CrossRef]
  6. Yamazoe, N. Toward innovations of gas sensor technology. Sens. Actuators B Chem. 2005, 108, 2–14. [Google Scholar] [CrossRef]
  7. Moumen, A.; Kumarage, G.C.W.; Comini, E. P-Type Metal Oxide Semiconductor Thin Films: Synthesis and Chemical Sensor Applications. Sensors 2022, 22, 1359. [Google Scholar] [CrossRef]
  8. Song, B.-Y.; Liu, L.-H.; Yang, M.; Zhang, X.-F.; Deng, Z.-P.; Xu, Y.-M.; Huo, L.-H.; Gao, S. Biotemplate-directed synthesis of Cr2O3 mesoporous monotubes for enhanced sensing to trace H2S gas. Sens. Actuators B 2022, 369, 132294. [Google Scholar] [CrossRef]
  9. Liu, B.; Han, W.; Li, X.; Li, L.; Tang, H.; Lu, C.; Li, Y.; Li, X. Quasi metal organic framework with highly concentrated Cr2O3 molecular clusters as the efficient catalyst for dehydrofluorination of 1,1,1,3,3-pentafluoropropane. Appl. Catal. B 2019, 257, 117939. [Google Scholar] [CrossRef]
  10. Cheng, J.; Li, Y.; Zhong, J.; Lu, Z.; Wang, G.; Sun, M.; Jiang, Y.; Zou, P.; Wang, X.; Zhao, Q.; et al. Molecularly imprinted electrochemical sensor based on biomass carbon decorated with MOF-derived Cr2O3 and silver nanoparticles for selective and sensitive detection of nitrofurazone. Chem. Eng. J. 2020, 398, 125664. [Google Scholar] [CrossRef]
  11. Atkinson, A.; Levy, M.R.; Roche, S.; Rudkin, R.A. Defect properties of Ti-doped Cr2O3. Solid State Ion. 2006, 177, 1767–1770. [Google Scholar] [CrossRef]
  12. Henshaw, G.S.; Dawson, D.H.; Williams, D.E. Selectivity and composition dependence of response of gas-sensitive resistors. Part 2.—Hydrogen sulfide response of Cr2−xTixO3+y. J. Mater. Chem. 1995, 5, 1791–1800. [Google Scholar] [CrossRef]
  13. Blacklocks, A.N.; Atkinson, A.; Packer, R.J.; Savin, S.L.P.; Chadwick, A.V. An XAS study of the defect structure of Ti-doped α- Cr2O3. Solid State Ion. 2006, 177, 2939–2944. [Google Scholar] [CrossRef]
  14. Shaw, G.A.; Pratt, K.F.E.; Parkin, I.P.; Williams, D.E. Gas sensing properties of thin film (≤3 μm) Cr2−xTixO3 (CTO) prepared by atmospheric pressure chemical vapour deposition (APCVD), compared with that prepared by thick film screen-printing. Sens. Actuators B 2005, 104, 151–162. [Google Scholar] [CrossRef]
  15. Du, J.; Wu, Y.; Choy, K.-L. Controlled synthesis of gas sensing Cr2−xTixO3 films by electrostatic spray assisted vapour deposition and their structural characterisation. Thin Solid Films 2006, 497, 42–47. [Google Scholar] [CrossRef]
  16. Du, J.; Wu, Y.; Choy, K.-L.; Shipway, P.H. Structure, properties and gas sensing behavior of Cr2−xTixO3 films fabricated by electrostatic spray assisted vapour deposition. Thin Solid Films 2010, 519, 1293–1299. [Google Scholar] [CrossRef]
  17. Conde-Gallardo, A.; Cruz-Orea, A.; Zelaya-Angel, O.; Bartolo-Pèrez, P. Electrical and optical properties of Cr2−xTixO3 thin films. J. Phys. D Appl. Phys. 2008, 41, 205407. [Google Scholar] [CrossRef]
  18. Vallejos, S.; Di Maggio, F.; Shujah, T.; Blackman, C. Chemical Vapour Deposition of Gas Sensitive Metal Oxides. Chemosensors 2016, 4, 4. [Google Scholar] [CrossRef]
  19. Ehsan, M.A.; Shah, S.S.; Basha, S.I.; Hakeem, A.S.; Aziz, M.A. Recent Advances in Processing and Applications of Heterobimetallic Oxide Thin Films by Aerosol-Assisted Chemical Vapor Deposition. Chem. Rec. 2022, 22, e202100278. [Google Scholar] [CrossRef]
  20. Srirattanapibul, S.; Nakarungsee, P.; Issro, C.; Tang, I.M.; Thongmee, S. Enhanced room temperature NH3 sensing of rGO/Co3O4 nanocomposites. Mater. Chem. Phys. 2021, 272, 125033. [Google Scholar] [CrossRef]
  21. Dey, A. Semiconductor metal oxide gas sensors: A review. Mater. Sci. Eng. B 2018, 229, 206–217. [Google Scholar] [CrossRef]
  22. Natarajamani, G.S.; Kannan, V.P.; Madanagurusamy, S. Synergistically enhanced NH3 gas sensing of graphene oxide-decorated Nano-ZnO thin films. Mater. Chem. Phys. 2024, 316, 129036. [Google Scholar] [CrossRef]
  23. Li, D.; Han, D.; Chen, Y.; Hong, Y.; Duan, Q.; Wang, H.; He, X.; Zhao, L.; Wang, W.; Sang, S. GaN/rGO nanocomposite gas sensor for enhanced NH3 sensing performances at room temperature. Sens. Actuators B 2024, 403, 135209. [Google Scholar] [CrossRef]
  24. Song, B.-Y.; Zhang, X.-F.; Huang, J.; Cheng, X.-L.; Deng, Z.-P.; Xu, Y.-M.; Huo, L.-H.; Gao, S. Porous Cr2O3 Architecture Assembled by Nano-Sized Cylinders/Ellipsoids for Enhanced Sensing to Trace H2S Gas. ACS Appl. Mater. Interfaces 2022, 14, 22302–22312. [Google Scholar] [CrossRef]
  25. Dutta, S.; Pandey, A.; Leeladhar; Jain, K.K. Growth and characterization of ultrathin TiO2-Cr2O3 nanocomposite films. J. Alloys Compd. 2017, 696, 376–381. [Google Scholar] [CrossRef]
  26. Baraskar, P.; Chouhan, R.; Agrawal, A.; Choudhary, R.J.; Sen, P. Weak ferromagnetism at room temperature in Ti incorporated Cr2O3 thin film. Phys. B 2019, 571, 36–40. [Google Scholar] [CrossRef]
  27. Williams, D.E. Semiconducting oxides as gas-sensitive resistors. Sens. Actuators B Chem. 1999, 57, 1–16. [Google Scholar] [CrossRef]
  28. Blackman, C. Do We Need “Ionosorbed” Oxygen Species? (Or, “A Surface Conductivity Model of Gas Sensitivity in Metal Oxides Based on Variable Surface Oxygen Vacancy Concentration”). ACS Sens. 2021, 6, 3509–3516. [Google Scholar] [CrossRef]
  29. Kucharski, S.; Ferrer, P.; Venturini, F.; Held, G.; Walton, A.S.; Byrne, C.; Covington, J.A.; Ayyala, S.K.; Beale, A.M.; Blackman, C. Direct in situ spectroscopic evidence of the crucial role played by surface oxygen vacancies in the O2-sensing mechanism of SnO2. Chem. Sci. 2022, 13, 6089–6097. [Google Scholar] [CrossRef]
  30. Cox, D.F.; Fryberger, T.B.; Semancik, S. Oxygen vacancies and defect electronic states on the SnO2 (110)-1 × 1 surface. Phys. Rev. B 1988, 38, 2072–2083. [Google Scholar] [CrossRef]
  31. Gunkel, F.; Christensen, D.V.; Chen, Y.Z.; Pryds, N. Oxygen vacancies: The (in)visible friend of oxide electronics. Appl. Phys. Lett. 2020, 116, 120505. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of an aerosol-assisted CVD.
Figure 1. Schematic diagram of an aerosol-assisted CVD.
Chemosensors 12 00218 g001
Figure 2. (a) Photograph of sensor platform, and (b) schematic diagram of a sensor connected to nickel pins via platinum wires.
Figure 2. (a) Photograph of sensor platform, and (b) schematic diagram of a sensor connected to nickel pins via platinum wires.
Chemosensors 12 00218 g002
Figure 3. XRD patterns of (a) Cr2O3 thin film-based sensors exposed to different annealing temperatures and (b) CTO thin film-based sensors subjected to different annealing times.
Figure 3. XRD patterns of (a) Cr2O3 thin film-based sensors exposed to different annealing temperatures and (b) CTO thin film-based sensors subjected to different annealing times.
Chemosensors 12 00218 g003
Figure 4. Raman spectra of the Cr2O3 and CTO thin films.
Figure 4. Raman spectra of the Cr2O3 and CTO thin films.
Chemosensors 12 00218 g004
Figure 5. (a) High-resolution and (b) low-resolution SEM images of the CTO thin film.
Figure 5. (a) High-resolution and (b) low-resolution SEM images of the CTO thin film.
Chemosensors 12 00218 g005
Figure 6. XPS spectra of the CTO film: (a) survey scan, (b) Ti 2p, (c) Cr 2p, and (d) O 1s.
Figure 6. XPS spectra of the CTO film: (a) survey scan, (b) Ti 2p, (c) Cr 2p, and (d) O 1s.
Chemosensors 12 00218 g006
Figure 7. Change in resistance of the (a) Cr2O3-based sensors and (b) CTO-based sensors to C4H8 gas.
Figure 7. Change in resistance of the (a) Cr2O3-based sensors and (b) CTO-based sensors to C4H8 gas.
Chemosensors 12 00218 g007
Figure 8. Rg/Ra against C4H8 concentration for each Cr2O3 and CTO sensor at different temperatures: (a) 400 °C; (b) 450 °C.
Figure 8. Rg/Ra against C4H8 concentration for each Cr2O3 and CTO sensor at different temperatures: (a) 400 °C; (b) 450 °C.
Chemosensors 12 00218 g008
Figure 9. Change in resistance of the (a) Cr2O3-based sensors and (b) CTO-based sensors to NH3.
Figure 9. Change in resistance of the (a) Cr2O3-based sensors and (b) CTO-based sensors to NH3.
Chemosensors 12 00218 g009
Figure 10. Schematic illustration of the possible sensing mechanism for Cr2O3 and CTO thin films when exposed to C4H8 or NH3.
Figure 10. Schematic illustration of the possible sensing mechanism for Cr2O3 and CTO thin films when exposed to C4H8 or NH3.
Chemosensors 12 00218 g010
Table 1. Preparation conditions and Scherrer sizes of the different sensors.
Table 1. Preparation conditions and Scherrer sizes of the different sensors.
SamplesAnnealing Temperatures (°C)Annealing Times (h)Scherrer Size (nm)
Cr2O3-15002426
Cr2O3-26002431
Cr2O3-37002430
CTO-1600625
CTO-26003036
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

Zhou, P.; Tsang, J.-H.; Blackman, C.; Shen, Y.; Liang, J.; Covington, J.A.; Saffell, J.; Danesh, E. A Novel Mechanism Based on Oxygen Vacancies to Describe Isobutylene and Ammonia Sensing of p-Type Cr2O3 and Ti-Doped Cr2O3 Thin Films. Chemosensors 2024, 12, 218. https://doi.org/10.3390/chemosensors12100218

AMA Style

Zhou P, Tsang J-H, Blackman C, Shen Y, Liang J, Covington JA, Saffell J, Danesh E. A Novel Mechanism Based on Oxygen Vacancies to Describe Isobutylene and Ammonia Sensing of p-Type Cr2O3 and Ti-Doped Cr2O3 Thin Films. Chemosensors. 2024; 12(10):218. https://doi.org/10.3390/chemosensors12100218

Chicago/Turabian Style

Zhou, Pengfei, Jone-Him Tsang, Chris Blackman, Yanbai Shen, Jinsheng Liang, James A. Covington, John Saffell, and Ehsan Danesh. 2024. "A Novel Mechanism Based on Oxygen Vacancies to Describe Isobutylene and Ammonia Sensing of p-Type Cr2O3 and Ti-Doped Cr2O3 Thin Films" Chemosensors 12, no. 10: 218. https://doi.org/10.3390/chemosensors12100218

APA Style

Zhou, P., Tsang, J. -H., Blackman, C., Shen, Y., Liang, J., Covington, J. A., Saffell, J., & Danesh, E. (2024). A Novel Mechanism Based on Oxygen Vacancies to Describe Isobutylene and Ammonia Sensing of p-Type Cr2O3 and Ti-Doped Cr2O3 Thin Films. Chemosensors, 12(10), 218. https://doi.org/10.3390/chemosensors12100218

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