Next Article in Journal
Energy-Efficient Neuromorphic Architectures for Nuclear Radiation Detection Applications
Previous Article in Journal
Hybrid Precision Floating-Point (HPFP) Selection to Optimize Hardware-Constrained Accelerator for CNN Training
Previous Article in Special Issue
Ultrathin Platinum Film Hydrogen Sensors with a Twin-T Type Notch Filter Circuit
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Sensing Response of Magnesium Antimoniate Oxide (MgSb2O6) in the Presence of Propane Atmospheres at Different Operating Voltages

by
Héctor Guillén-Bonilla
1,
José Trinidad Guillén-Bonilla
2,
Verónica-María Rodríguez-Betancourtt
3,
Jorge Alberto Ramírez-Ortega
3,
Juan Pablo Morán Lázaro
4 and
Alex Guillén-Bonilla
4,*
1
Departament of Project Engineer, CUCEI, Universidad de Guadalajara, M. García Barragán 1421, Guadalajara 44410, Mexico
2
Departament of Electro-Photonics, CUCEI, Universidad de Guadalajara, M. García Barragán 1421, Guadalajara 44410, Mexico
3
Departament of Physics, CUCEI, Universidad de Guadalajara, Guadalajara 44410, Mexico
4
Department of Computer Science and Engineering, CUVALLES, Universidad de Guadalajara, Carretera Guadalajara-Ameca Km 45.5, Ameca 46600, Mexico
*
Author to whom correspondence should be addressed.
Sensors 2024, 24(7), 2147; https://doi.org/10.3390/s24072147
Submission received: 22 February 2024 / Revised: 18 March 2024 / Accepted: 26 March 2024 / Published: 27 March 2024
(This article belongs to the Collection Gas Sensors)

Abstract

:
Nanoparticles of MgSb2O6 were synthesized using a microwave-assisted wet chemistry method, followed by calcination at 700 °C. Their ability to detect different concentrations of propane gas (C3H8) at various operating voltages was evaluated. The material’s crystalline phase was identified using X-ray powder diffraction (XRD). The morphology was analyzed by scanning electron microscopy (SEM), finding bar- and polyhedron-type geometries. Through transmission electron microscopy (TEM), we found particle sizes of 8.87–99.85 nm with an average of ~27.63 nm. Employing ultraviolet–visible (UV-Vis) spectroscopy, we found a band gap value of ~3.86 eV. Thick films made with MgSb2O6 powders were exposed to atmospheres containing 150, 300, 400, and 600 ppm of propane gas for dynamic testing. The time-dependent sensitivities were ~61.09, ~88.80, ~97.65, and ~112.81%. In addition, tests were carried out at different operating voltages (5–50 V), finding very short response and recovery times (~57.25 and ~18.45 s, respectively) at 50 V. The excellent dynamic response of the MgSb2O6 is attributed mainly to the synthesis method because it was possible to obtain nanometric-sized particles. Our results show that the trirutile-type oxide MgSb2O6 possesses the ability, efficiency, and thermal stability to be applied as a gas sensor for propane.

1. Introduction

Extensive research is being conducted on semiconductor materials for their potential use as gas sensors [1]. These materials are known for their exceptional ability to change their electrical resistance when subjected to varying concentrations of certain gases (such as oxidants or toxins) in the presence of oxygen [1,2]. The electrical response of these semiconductors is based on the adsorption and desorption of oxygen ions on the surface [3,4] and the mobility of the charge carriers (holes or electrons) due to the temperature employed in the tests [1,2,3,4,5,6]. The variation in the electrical signal (resistance) shown by the semiconductor oxides depends on the detected gas and the type of semiconductor (p or n) [6,7]. In the gas sensor field, n-type semiconductors are the most studied [8] because they show excellent electrical response, thermal stability, and efficiency toward almost any gas (CO, O2, CO2, SO2, CH4, etc.) [8,9]. In addition, the advantages of using semiconductor oxides (whether p- or n-type) as gas sensors lie in the low production cost, the ease of use, the high capacity to detect various gases, and the simple manufacture of detection devices [7,8,10]. According to the literature, the ideal semiconductors for gas detection allow evaluation of their capacity and performance through parameters such as sensitivity, selectivity, response time, reversibility, and recovery time [11]. These parameters are strongly related to the material’s microstructural features, such as particle size, agglomeration, morphology, surface area, and porosity [12]. Other aspects that affect the detection properties are temperature, the thickness of the sensor surface [12], and its shape (which can be thin or thick films or pellets) [13,14,15]. Such features improve performance and efficiency in detecting low and high concentrations of gases.
Since the ability of a semiconductor to detect gases was discovered [8,13], n-type binary oxides (such as ZnO and SnO2 [16]) have been the most studied materials due to their fast response, high sensitivity, thermal stability, selectivity, reproducibility, short response and recovery times, and high reliability [8,16,17]. Other p- and n-type binary oxides have also been intensively studied for their potential application as gas detectors [1]. Among them are: TiO2, WO3, α-Fe2O3, In2O3, Cu2O, Co3O4, Cr2O3, Mn3O4, and NiO [4,17]. Some authors have reported that the ternary semiconductors LaCoO3 [18], NdCoO3 [19], ZnAl2O4 [1], ZnMn2O4 [20], and CoTa2O6 [21] can be considered for use as gas sensors due to their good electrical response and thermal stability in different atmospheres. Additionally, several research groups have found that antimonates with different divalent cations (with formula MSb2O6, where M can be any cation of Ni, Co, Zn, Mn, etc.) could be used as gas sensors [7,22]. Some of the antimonates that have been studied include CoSb2O6 [22,23,24], MnSb2O6 [25], and ZnSb2O6 [26]. At room temperature, CobSb2O6 showed an excellent response to liquefied petroleum gas (LPG) concentrations of 5000 ppm [22]. Similarly, MnSb2O6 detected propane levels from 50 ppm at 100 °C [25], while ZnSb2O6 demonstrated good sensitivity to LPG concentrations from 5000 ppm. According to a recent study [7], nickel antimonate (NiSb2O6) showed an excellent response in static propane atmospheres, which led to the development of a gas detection prototype based on pellets of the oxide. These studies found that materials with a trirutile-type structure have good response and recovery times, stability, and reproducibility in propane atmospheres.
In general, the gases in which trirutile-type antimoniates have been tested the most are O2, CO, CO2, C3H8, and LPG [22,23,25,26], showing excellent electrical response and thermal stability, which is attributed to the morphology, as well as the nanometric particle size obtained when the oxides are synthesized [14,26]. As mentioned above, the ability, efficiency, and stability of gas sensors are closely related to the nanometric particle size of the semiconductor material [11]. This is because when the particle size is between 1 and 100 nm, the specific surface area increases, favoring the improvement in the catalytic activity to adsorb gases on the semiconductor’s surface and the electrical response [2,11,27]. Studies have shown that various metallic semiconductor nanostructures such as nanowires, nanotubes, core–shells, nanofibers, nanoflowers, nanosheets, and random figured nanoparticles [2,18] exhibit lower operating temperatures, increased sensitivity, improved selectivity, and lower response and recovery times [27]. In a previous study, we synthesized magnesium antimonate (MgSb2O6) nanorods to measure its detection properties in static CO and C3H8 atmospheres, achieving high sensitivity (~245.75 and ~61.66, respectively) when increasing gas concentration and operating temperature (from 23 to 300 °C) [28].
In this work, we synthesized magnesium antimonate by applying a wet chemistry process assisted with microwave radiation. This method is simple and cheap and allows good control of the material’s microstructure. Thick films were fabricated with powders of the oxide to study its ability to detect propane atmospheres at different concentrations, voltages, and operating temperatures. The material showed high sensitivity, thermal stability, efficiency, and reproducibility, which are desirable features of a gas sensor.

2. Materials and Methods

2.1. Synthesis

The wet chemistry method reported by Casillas-Zamora et al. [25] was used to synthesize MgSb2O6 nanoparticles with a trirutile-type structure. The reactants were magnesium nitrate hexahydrate (Mg(NO3)2·6H2O, Sigma-Aldrich, Jalmek, Guadalajara, Mexico, 99%), antimony trichloride (SbCl3, Sigma-Aldrich, Jalmek, Guadalajara, Mexico, 99%), ethylenediamine (C2H8N2, Sigma-Aldrich, 99%), and ethyl alcohol (C2H5OH, Jalmek, Guadalajara, Mexico, 99.5%). Three solutions were prepared using 1.28 g of Mg(NO3)2·6H2O, 2.80 g of SbCl3, and 5 mL of ethylenediamine separately. To each solution, 5 mL of ethyl alcohol was added, except for the C2H8N2 solution, to which 10 mL of the alcohol was added. All solutions were left under constant stirring at 300 rpm for 20 min at 25 °C. Then, the Mg(NO3)2·6H2O and C2H8N2 solutions were mixed. During the synthesis, the two amino groups forming ethylenediamine’s structure favored the formation of metal complexes by capturing the Mg2+ ions with the nitrogen’s free electrons, generating a large molecular mesh. Next, that solution was mixed with the SbCl3 solution, which joined the ethylenediamine-Mg2+ metal complex to form the Mg2Sb2O6. The final mixture was kept under constant stirring for 24 h at 25 °C. After the stirring, the ethyl alcohol was evaporated by applying microwave radiation (18 irradiations of 60 s) in a domestic oven (General Electric, model JES769WK) at a power of 140 W. The total energy applied to the solution was 151.2 kJ. The purpose of applying microwave radiation in steps of 60 s was to maintain the colloidal solution below 70 °C and thus avoid material loss due to splashes. After evaporation, the resulting material was dried at 200 °C for 8 h and then calcined at 700 °C in a Novatech muffle. The heating rate to reach 700 °C was 100 °C/h. Then, it was left at that temperature for 5 h.

2.2. Physical Characterization

To analyze the crystalline structure of the oxide calcined at 700 °C, a diffractometer (Panalytical Empyrean, Guadalajara, Mexico) coupled to a deuterium/tungsten-halogen lamp was used, applying Cuα radiation with a wavelength (λ) of 1.5406 Å and a continuous scan (2θ) from 10 to 90° using 0.026 ° steps at a rate of 1 s/step. The value of the semiconductor’s forbidden bandgap was found by UV-Vis–NIR spectroscopy (UV-3600 Plus, Mexico City, Mexico). The absorbance spectrum was taken in a shift range from 200 to 700 nm. The calcined oxide’s microstructure was analyzed by field-emission scanning electron microscopy (FE-SEM) using a Tescan MIRA 3 LMU (Mexico City, Mexico) system with an acceleration voltage of 10 kV in a high vacuum. Transmission electron microscopy (TEM) was employed to study nanoparticles’ morphology and size using a Jeol system (model JEM-2010, Mexico City, Mexico) with an acceleration voltage of 100 kV. For this, the MgSb2O6 powders were dispersed in alcohol by ultrasound and drop-deposited onto Formvar-coated microgrids.

2.3. Dynamic Tests in C3H8 Atmospheres

Dynamic tests were carried out in air–C3H8 flows using thick films manufactured with powders of the MgSb2O6 calcined at 700 °C. A ceramic base with a central circular cavity and four ceramic millicolumns around the cavity was used. Each millicolumn had a small hole at half its length through which high-purity platinum wires (0.006 in diameter) were inserted to form the electrodes connected to the detection system. Then, 0.4 g of MgSb2O6 was dispersed in ethyl alcohol and placed dropwise into the ceramic base’s cavity to form a film ~500 µm thick and ~300 µm in diameter. Subsequently, the thick film was dried at 300 °C using a heating ramp of 100 °C/h for 4 h in a programmable muffle (Vulcan, model 5–550). The device was then placed in a quartz tube and inside a tubular furnace (Lindberg/blue) with programmable temperature control (Figure 1). The concentration of the air–C3H8 flows was controlled utilizing Brooks Instruments’, Cleveland, OH, USA, GF100CXXC-SH452.6L (2600 cm3/min) and GF100CXXC-SH40010C (10 cm3/min) mass flow regulators. When the material was exposed to the gas flows, the variation of its electrical resistance was quantified using a multimeter. The gas detection system was controlled with LabView 8.6 software (National Instruments, Cleveland, OH, USA). A schematic of the system used for the experiments is depicted in Figure 1.

3. Results

3.1. XRD Analysis

Figure 2 shows the diffractogram of the MgSb2O6 calcined at 700 °C. Using PDF file No. 88-1725, it was possible to identify the peaks associated with the oxide’s phase at 2θ = 19.17°, 21.41°, 27.15°, 28.85°, 33.46°, 34.91°, 38.78°, 40.10°, 43.61°, 44.72°, 48.10°, 53.16°, 56.01°, 59.95°, 63.33°, 67.06°, 67.72°, 73.73°, 76.78°, 80.87°, 83.22°, 86.57°, and 89.55°, corresponding to the crystalline planes (002), (101), (110), (111), (112), (103), (200), (113), (202), (211), (114), (213), (220), (006), (310), (116), (303), (206), (314), (305), (400), (226), and (412), respectively. According to that, the MgSb2O6 belongs to the family of trirutile-type materials, with a tetragonal crystalline structure, cell parameters a = 4.64 Å and c = 9.25 Å, and a space group P42/mnm (136) [28,29]. High crystallinity and purity are also inferred from the diffractogram, as reported in other works [25,29,30]. To determine the crystallite size, Scherrer’s equation [24] was used:
t = 0.9 λ β c o s θ ,
where λ is the wavelength (1.5406 Å), β is the full width at half maximum of Bragg’s peak, and θ is Bragg’s angle. The most intense peak (110) corresponding to the last calcination was considered for the calculation. A crystallite size of ~43.48 nm was obtained.
Comparing our results (shown in Figure 2) with the literature, we confirmed that the MgSb2O6’s crystalline phase was obtained without secondary phases by the heat treatment (at 700 °C for 5 h). Other authors synthesized MgSb2O6 using the solid-state reaction method, subjecting the powder to 1000 °C for 48 h [29]. Nagarajan and Naraginti [30] reported that the oxide was synthesized by the solid-state reaction, calcining at 600 and 900 °C for 12 and 6 h, respectively. In a previous study, we synthesized MgSb2O6 nanoparticles at 800 °C using a chemical method, identifying a secondary phase associated with carbon (C) [28]. Thus, the synthesis route and the applied calcination temperature employed in this work are the best method to obtain MgSb2O6’s crystalline phase.

3.2. UV-Vis Analysis

A UV-Vis absorption spectrum of the MgSb2O6 calcined at 700 °C is shown in Figure 3a, measured in a wavelength range of 200 to 700 nm (2.00 to 6.00 eV). In the 200–300 nm range, bands characteristic of materials with a trirutile-type structure [30,31,32,33] were identified. To evaluate the MgSb2O6’s forbidden energy bandgap, Tauc’s formula was used: ( α h υ ) n = A ( h v E g ) , where α is the absorption coefficient, h ν is the discrete energy, A is the band parameter, E g is the band’s energy gap, and n depends on the semiconductor transition type, where n = 2 for the direct transition [30,31,32]. The estimated value was ~3.86 eV (Figure 3b).
This result is consistent with previous studies [31,32]. For example, Arunkumar and Naraginti [31] found that when exchanging the divalent cation of the trirutile structure (CoSb2O6, CuSb2O6, NiSb2O6, and FeSb2O6), the bandgap value ranged from 2.10–3.83 eV. Nagarajan and Naraginti [30] obtained values in the range of 3.30–4.05 eV for the MgSb2O6. Our value of ~3.86 eV is within those reported for trirutile-type semiconductors [30,31]. The band gap value in a trirutile-type semiconductor is strongly related to the synthesis method and the cation incorporated into the crystal lattice [2,17,31].

3.3. SEM Analysis

Figure 4 depicts six typical photomicrographs at magnifications of 9.42 kx 15 kx, 20.00 kx, 80 kx, 80 kx, and 80 kx of the microstructure of the MgSb2O6 calcined at 700 °C. Figure 4a shows that the oxide’s surface comprised very fine filament-type particles. Some appear like different-sized microneedles evenly distributed throughout the material’s surface. Figure 4b–d show polyhedron- and bar-shaped particles composed of smaller particles of different sizes (~0.1 μm). Figure 4e,f depict the growth in all directions of rods made up of assembled smaller particles (average size ~90 nm). It is worth mentioning that the particles agglomerated on the entire surface due to the material’s heat treatment.
To estimate the length and diameter of the microrods, several SEM images of different areas of the surface were required (Figure 5). The average size of the microrods was calculated in the range of 50–350 nm, with a mean of ~161.31 nm and a standard deviation of ~±54.48 nm (Figure 5a). The diameter of the microrods was calculated in a range of 10–55 nm, with a mean of ~30.22 nm and a standard deviation of ~±8.54 nm (Figure 5b). The measurements were carried out where the particles were clearly identifiable.
According to Figure 4, the formation of microrods, microneedles, and other similar morphologies is strongly related to the synthesis method [2,10,11,18]. Lamer and Dinegar proposed a possible mechanism for the nucleation and growth of particles, like those obtained here, using chemical methods. The rationale lies in increasing species’ concentration quickly until reaching a critical concentration. As a result, nuclei begin to form until a supersaturation concentration is reached. The nucleation process will end, and the particle growth will continue until the solubility reaches an equilibrium [34]. One of the drawbacks of Lamer and Dinegar’s model is the redissolution and precipitation of smaller particles (“Ostwald ripening”). To remedy that, the use of organic complexes such as ethylenediamine (considered as a stabilizer), surfactant molecules, or bases with nanometric structures that help maintain greater control in the particles’ nucleation and growth processes has been proposed [35,36], which gives rise to the formation of polyhedral and bar structures, such as those obtained in this work.

3.4. TEM Analysis

Figure 6 shows typical TEM images of the microstructure of the MgSb2O6 powders calcined at 700 °C. It is essential to mention that the powders were previously dispersed in ethyl alcohol using a sonifier to be able to analyze the particles individually. In Figure 6a–c, an agglomeration of differently sized particles (in the order of nanometers) is observed. These nanoparticles are linked together by necks formed by the heat treatment, the coalescence of the particles, and the material’s residence time in the muffle. Figure 6d,e show the formation of hexagonal structures and differently oriented agglomerated polyhedral shapes. The nanostructures’ morphology is attributed to the heat treatment and the effect of the chelating agent (ethylenediamine) [7,14,23,37]. According to the literature, applying chelating agents, such as ethylenediamine, in synthesizing materials like the one studied here favors the creation of organometallic complexes that form a template for particle growth. By calcining the MgSb2O6 powders, the organic material is eliminated, giving rise to octahedral and tetrahedral structures, nanorods, nanowires, and randomly shaped nanoparticles [25,26,28,37]. The nanoparticles’ size was estimated at 8.87–99.85 nm, with a mean of ~27.63 nm and a standard deviation of ~±17.69 nm (Figure 7). The calculation of the particles’ average size was carried out in areas where they were clearly visible.

3.5. Gas-Sensing Properties

To evaluate the ability of the MgSb2O6 nanoparticles obtained at 700 °C to detect C3H8 atmospheres, experiments were carried out using, first, 560 ppm of the gas (Figure 8). Later, propane concentrations of 150, 300, 400, and 600 ppm were injected into the measuring chamber (Figure 9). In these tests, the MgSb2O6 films were always at 400 °C in a direct current (DC) of 200 μA. Films’ sensitivity was estimated with the formula S = ( R a R g ) / R g × 100, where R a is air resistance and R g is test gas resistance. During the measurements, a constant flow of 1500 cm3/min of extra-dry air (21% O2, 79% N2) was employed to stabilize the thick films’ surface. For the 560 ppm sensing tests, the thick film surface was supersaturated by a flow of extra-dry air for the first 8 min. Then, the 560 ppm of C3H8 was injected for 8 min. Subsequently, the propane flow was stopped, observing that the thick films’ electrical resistance returned to its original value when the films were subjected to extra-dry air flows, thus corroborating the reproducibility of the material. This process was repeated cyclically until the test’s end (lasting approximately 55 min). The results are shown in Figure 8, where the change in electrical resistance (Figure 8a) and the sensitivity percentage (Figure 8b) at 400 °C can be observed.
As expected, when injecting the C3H8 at a constant current of 200 μA, the material’s electrical resistance decreased, while the sensitivity percentage increased considerably. The excellent dynamic response, the high sensitivity, and the good reproducibility shown by the MgSb2O6 films [14] were verified with the number of uniform cycles carried out during the test. The curves’ behavior shown in Figure 8a,b commonly occurs in an n-type semiconductor when exposed to atmospheres like the one studied here [2]. The variation in electrical resistance ranged from 34.22 to 79.74 kΩ, with an average of 45.52 kΩ. In contrast, the dynamic sensitivity range was 0.36 to 128.25%, with an average of 127.97%. To estimate the response and recovery times of the MgSb2O6 films, we considered 90% of the variation in electrical resistance in propane and 10% when exposed to air atmospheres [38]. Therefore, the calculated average response and recovery times were 1.898 and 5.63 min, respectively.
The excellent response shown in Figure 8 is mainly attributed to the chemical reactions between the test gas and the oxygen species ( O and O 2 -ionic forms [2,11]) previously adsorbed on the material’s surface due to the temperature (in our case, at 400 °C) [4]. These oxygen species are more reactive than those below 200 °C ( O 2 ) [11]. It means that the C3H8 chemisorption on the film’s surface increased due to the temperature, favoring the mobility of the charge carriers (electrons) [14,23], which provoked variations in the electrical resistance and an increase in the material’s dynamic sensitivity.
Figure 9 shows the dynamic tests’ results at different C3H8 concentrations (150, 300, 400, and 600 ppm) at constant temperature (400 °C) using a current of 200 μA. The experiments were carried out by injecting the propane concentrations into the measurement chamber, first increasing them (150–600 ppm) and then decreasing them (600–150 ppm). By varying the C3H8 concentrations, the MgSb2O6 films showed a drop in electrical resistance and increased dynamic sensitivity percentage as the gas concentration rose and vice versa (Figure 9a–c). With these experiments, we could test the reproducibility, stability, efficiency, and ability to detect low and high concentrations of C3H8. The average values of the electrical resistance in both directions for concentrations of 150, 300, 400, and 600 ppm were 52.50, 65.95, 68.90, and 70.59 kΩ, respectively. The respective sensitivity values were 61.09, 88.80, 97.65, and 112.81%. Using information from Kida et al. [38], we calculated the response and recovery times considering only the results for the increasing propane concentration (Figure 9d). The results are summarized in Table 1.
On the other hand, calibration curves were obtained (Figure 9e) fitting to the equation S(C) = mC + b, where S is the response of the thick films, m is the response coefficient in C3H8, and b is power law’s constant. A correlation coefficient of 0.9691 indicates a good fit. This result suggests that a sensor made of MgSb2O6 can be employed for C3H8 concentrations within the range of 150 to 600 ppm, with 150 being the detection limit.
Then, according to Figure 9, thick films’ response improved when the test gas concentration increased. This means that the increase in the material’s dynamic response was strongly related to the rise in the gas concentration, which reacted with the available oxygen ( O ) [11,15,17] on the films’ surface, causing greater charge carriers’ mobility due to the operating temperature’s effect [2,11] (400 °C). This increase in electrons’ kinetics was attributed to the rising concentration of the test gas, leading to stronger chemical reactions. This resulted in an increase in activation energy, enhancing electron mobility on the thick films’ surface due to the operating temperature [15,17]. This contributed to the observed variation in electrical resistance and sensitivity [14,20,23], significantly improving the material’s detection properties. It has been reported that such improvements in semiconductors allow an increase in their response if the test gas concentration is increased [15]. Wang et al. [39] and Ramírez-Ortega et al. [40] reported that the operating temperature is the most probable cause of the rise in the semiconductor’s response. Other studies report that the response depends on the test gas, its concentration, the material’s microstructure, and the operating temperature [5,6,11]. All these conditions favor a better diffusion of the gas molecules on the material’s surface, causing an increase in its response. Furthermore, the literature suggests that metallic oxides like ours exhibit faster response and recovery times due to the temperature effect and the high level of interaction between the thick films’ surface and the test gas, implying that the speed of oxygen adsorption and desorption increases, resulting in faster response and recovery times.
Experiments were carried out at a concentration of 560 ppm at 400 °C to know the ability of the n-type MgSb2O6 to detect C3H8 atmospheres. A direct current (DC) signal was used for this, applying different voltages: 5, 10, 15, 20, 25, 30, 35, 40, 45, and 50 V. As for the results shown in Figure 8b and Figure 9b, sensitivity was calculated using the formula S = ( R a R g ) / R g × 100, where R a is air resistance and R g is propane resistance. The results are shown in Figure 10. The data were graphed considering the variation in electrical resistance and the sensitivity percentage as a function of time (Figure 10a–c). A graph depicting the response and recovery times calculated as a function of the applied voltage (Figure 10d) is also shown. According to these results, the electrical resistance decreases with increasing voltage. This decrease in resistance is more obvious at higher voltages (30–50 V, Figure 10b). As can be observed, the rise in voltage and operating temperature favors the increase in energy, which causes greater mobility of the charge carriers on the material’s surface [40]. We observed that, when increasing the voltage at a constant temperature and constant flow of extra-dry air (21% O2), the thick films’ sensitivity percentage increased. This is associated with the enrichment of the oxygen species on the films’ surface due to the flow of extra-dry air, which caused a higher concentration of highly reactive oxygen species ( O ) [15,17] that reacted vigorously when the voltage and the temperature (400 °C) increased, favoring changes in the electrical resistance and, therefore, the increase in the material’s dynamic sensitivity percentage. We observed that with increasing operating temperature and voltage, oxidation of the test gas by oxygen species occurred, thus contributing to greater mobility of the charge carriers as well as an increase in the conductivity of the films [2,11], causing the material’s sensitivity to increase significantly [11]. For example, at 50 V, an increase in charge energy and an improvement in the oxygen adsorption and desorption on the surface are promoted, increasing the sensitivity of the MgSb2O6. According to the literature, the chemisorption of oxygen species as a function of temperature is the most likely cause of the films’ increase in dynamic response and sensitivity [13,14]. Additionally, our results indicate that increasing voltage also improves the material’s ability to detect C3H8 atmospheres.
The electrical resistance at 400 °C was 23.73, 9.34, 9.03, 6.67, 5.13, 4.18, 3.34, 2.65, 2.21, and 1.78 MΩ respective to the voltages 5, 10, 15, 20, 25, 30, 35, 40, 45, and 50 V. For the same voltages, the sensitivity percentages were 20.35, 19.09, 30.74, 33.72, 35.93, 39.50, 40.97, 42.71, 45.35, and 49.01%, respectively. The response and recovery times were calculated as in the previous case (according to reference [38]), considering 90% of the oxide’s electrical resistance variation in C3H8 and 10% in air. Thus, the response times respective to each voltage were 80.40, 78.53, 76.67, 74.83, 71.09, 69.24, 65.63, 63.70, 60.93, and 57.25 s. Recovery times were 38.77, 36.10, 34.15, 31.38, 29.54, 27.69, 25.85, 22.16, 20.31, and 18.45 s (see Table 2). Response and recovery times decreased considerably as the voltage increased while keeping the air flow and operating temperature constant. As previously mentioned, this is attributed to the fact that when the voltage changed, a vigorous chemical reaction (i.e., a high reaction rate) occurred between the test gas and the oxygen available on the thick films’ surface due to the temperature (400 °C) [40,41], causing a rapid response (i.e., a decrease in the material’s response and recovery times).
From Figure 10e, it is evident that the sensor’s electrical response is correlated with the applied voltage and the optimal operating temperature that triggers the reaction of oxygen present on the sensor’s surface with the test gas at 400 °C. This correlation is supported by Figure 10a,b, which show the relationship between the decrease in electrical resistance and the increase in sensitivity percentage as a function of the applied voltage. A linear calibration curve corresponding to the sensor’s sensitivity percentage was plotted using the equation S(V) = mV + b, where S is the sensitivity, m is the sensitivity coefficient in propane, and b is the power law’s constant. The calculated R2 value of 0.9654 indicates a good fit, implying that the material can show a sensitivity percentage at voltages ranging from 5 to 50 V, consistently showing response increments as the voltage rises.
A possible chemical mechanism in propane detection for semiconductors like the one studied here has yet to be fully established. However, different authors have reported that the C3H8 detection process is effectively related to the operating temperature, the gas concentration, and the presence of oxygen [4,5,6]. When the material’s surface comes into contact with C3H8 molecules, the gas dissociates before reacting with oxygen, causing changes in the semiconductor’s electrical resistance [2,3,17]. When the propane molecules are adsorbed, they react with O oxygen species on the surface, producing CO2, water vapor, and a release of electrons to the semiconductor material [7,14,42]. A possible chemical reaction that occurs in the detection of the C3H8 is the following [42]:
C3H8 + 10O2− → 3CO2 + 4H2O + e
It means that when the propane is injected into the measurement chamber and comes into contact with the thick films’ surface, the gas molecules dissociate before reacting with the ionosorbed oxygen species [14,42], causing an electron transfer during the gas chemisorption [40,42]. This leads to a decrease in electrical resistance and an increase in the MgSb2O6’s sensitivity percentage. Another factor that significantly favors the excellent response of our oxide films is the use of the Sb cation in the synthesis process [29,31,32]. That is because by combining Sb3+ or Sb5+ with divalent cations such as Mg2+ (or Co2+, Zn2+, Ni2+), trirutile-type compounds (CoSb2O6, ZnSb2O6, NiSb2O6, respectively) are produced, which are thermally stable when exposed to toxic atmospheres [22,26,40]. It is reported in the literature that an advantage of using materials composed of the Sb cation (such as the trirutile-types) is that they substantially improve the catalytic activity and, with it, the ability to detect gases like CO, CO2, liquefied petroleum gas (LPG), and propane [14,22,26], among others. Also, a high chemical reactivity is favored for semiconductors containing Sb, which improves their catalytic properties (see Table 3) [7,14,22,29]. This offers a tremendous advantage for detecting toxic (CO, CO2) and flammable (C3H8) gases compared to other semiconductor oxides that have been studied as gas sensors [26,28,42].
For instance, in a study [14], it was found that ZnAl2O4 took 176 s to respond and 205 s to recover when exposed to 1000 ppm of propane. Another study [22] found that CoSb2O6 took 21 s to respond and 234 s to recover when exposed to 5000 ppm of propane. Lastly, MgSb2O6 showed a response when exposed to 500 ppm of propane [28]. In this work, we found that MgSb2O6 had a response time of 57.2 s and a recovery time of 18.5 s when exposed to low concentrations of propane gas (560 ppm). So, our results suggest that MgSb2O6 can be considered an excellent gas sensor.
In summary, using Sb and Mg cations to form the trirutile-like structure of the MgSb2O6 contributed to improving its catalytic activity and the diffusion of the test gas on the thick films’ surface at 400 °C. It was verified that the excellent compound’s detection ability, reproducibility, and efficiency are mainly due to the microstructural characteristics and the nanometric particle size (in our case, ~27.63 nm). Additionally, some studies suggest that oxygen adsorption and desorption, as well as the microstructure of the material, are additional factors that affect the detection properties of the sensor [4,11,15]. Specifically, if the particle size is less than half the thickness of the outer layer (LS) [17,27], then the adsorbed oxygen species are responsible for the variation in detection properties and an increase in sensitivity. Furthermore, if the crystal or particle size is D < 2L, the crystals participate in electronic transport during sensing [17,27]. This leads to variations in electrical resistance, resulting in an increase in the sensor’s response, high electrical sensitivity, thermal stability, and high efficiency. By reducing the MgSb2O6 particles’ dimensions, their surface area increased, thus improving the diffusion of C3H8 on the films’ surface due to the operating temperature. This caused the oxygen adsorption and desorption process to improve on the surface and, with it, changes in the material’s conductivity (or electrical resistance) [20,23,26]. In agreement with the literature, the improvement of the gas-sensing ability of our material, as well as its thermal stability and electrical response, occurred when its particle size was brought to a nanometric scale [2,3,11]. Similarly, with the use of the Sb cation in preparing the MgSb2O6, its response and recovery times, as well as the dynamic response, were optimized. This can be corroborated by the results shown in Figure 8, Figure 9 and Figure 10.

4. Conclusions

The synthesis of MgSb2O6 nanoparticles was achieved using a chemical method aided by microwave radiation, resulting in a crystalline phase without secondary phases after heating at 700 °C for five hours. SEM analysis revealed polyhedral and bar-shaped morphologies, and TEM estimated the particle size to be 8.87–99.85 nm, with a mean of ~27.63 nm and a standard deviation of ±17.69 nm. The material’s band gap was determined to be ~3.86 eV using UV-Vis spectroscopy. Dynamic detection tests in propane atmospheres revealed a good response of the oxide at various gas concentrations (150, 300, 400, and 600 ppm) and operating voltages (5–50 V) with a maximum sensitivity of ~112.81 at 600 ppm and 400 °C. The material also showed a sensitivity of ~61.09 at lower gas concentrations (150 ppm) with response and recovery times of 6.15 and 3.22 min. These excellent results can be attributed to the synthesis method, operating temperature, and the material’s microstructure (bars and polyhedrons). Therefore, MgSb2O6 shows excellent potential as a propane sensor at concentrations as low as 150 ppm.

Author Contributions

Formal analysis, J.T.G.-B. and J.P.M.L.; investigation, J.A.R.-O. and V.-M.R.-B.; methodology, J.A.R.-O., J.T.G.-B., A.G.-B., V.-M.R.-B., J.P.M.L. and H.G.-B.; project administration, H.G.-B.; resources, A.G.-B. and J.P.M.L.; supervision, H.G.-B.; validation, A.G.-B.; visualization, V.-M.R.-B.; writing—original draft, J.A.R.-O., J.T.G.-B., A.G.-B., V.-M.R.-B., J.P.M.L., V.-M.R.-B. and H.G.-B.; writing—review and editing, J.A.R.-O. and H.G.-B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon request.

Acknowledgments

The authors thank Mexico’s National Council of Science and Technology (CONACyT), and the University of Guadalajara for the support granted. Likewise, we thank M. de la Luz Olvera-Amador and Miguel-Ángel Luna-Arias for their technical assistance. This research was carried out following the line of research “Nanostructured Semiconductor Oxides” of the academic group UDG-CA-895 “Nanostructured Semiconductors” of CUCEI, University of Guadalajara.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Ayesh, A.I. Metal/metal-Oxide nanoclusters for gas sensor applications. J. Nanomater. 2016, 2016, 2359019. [Google Scholar] [CrossRef]
  2. Fazio, E.; Spadaro, S.; Corsaro, C.; Neri, G.; Gianluca Leonardi, S.; Neri, F.; Lavanya, N.; Sekar, C.; Donato, N.; Neri, G. Metal-Oxide based nanomaterials: Synthesis, characterization and their applications in electrical and electrochemical sensors. Sensors 2021, 21, 2494. [Google Scholar] [CrossRef]
  3. Comini, E.; Baratto, C.; Concina, I.; Faglia, G.; Falasconi, M.; Ferroni, M.; Galstyan, V.; Gobbi, E.; Ponzoni, A.; Vomiero, A. Metal oxide nanoscience and nanotechnology for chemical sensors. Sens. Actuators B Chem. 2013, 179, 3–20. [Google Scholar] [CrossRef]
  4. Korotcenkov, G. Metal oxides for solid-state gas sensors: What determines our choice? Mater. Sci. Eng. B 2007, 139, 1–23. [Google Scholar] [CrossRef]
  5. Zappa, D.; Galstyan, V.; Kaur, N.; Munasinghe Arachchige, H.M.M.; Sisman, O.; Comini, E. “Metal oxide -based heterostructures for gas sensors”—A review. Anal. Chim. Acta 2018, 1039, 1–23. [Google Scholar] [CrossRef]
  6. Shingange, K.; Swart, H.C.; Mhlongo, G.H. Design of porous p-type LaCoO3 nanofibers with remarkable response and selectivity to ethanol at low operating temperature. Sens. Actuators B Chem. 2020, 308, 127670. [Google Scholar] [CrossRef]
  7. Guillen Bonilla, J.T.; Guillen Bonilla, H.; Rodríguez-Betancourtt, V.M.; Guillen Bonilla, A.; Casillas Zamora, A.; Blanco Alonso, O.; Ramírez Ortega, J.A. A gas sensor for application as a propane leak detector. J. Sens. 2021, 2021, 8871166. [Google Scholar] [CrossRef]
  8. Vesna Nikolic, M.; Milovanovic, V.; Vasiljevic, Z.Z.; Stamenkovic, Z. Semiconductor gas sensors: Materials, technology, design, and application. Sensors 2020, 20, 6694. [Google Scholar] [CrossRef]
  9. Hua, Z.; Tian, C.; Huang, D.; Yuan, W.; Zhang, C.; Tian, X.; Wang, M.; Li, E. Power-law response of metal oxide semiconductor gas sensors to oxygen in presence of reducing gases. Sens. Actuators B Chem. 2018, 267, 510–518. [Google Scholar] [CrossRef]
  10. Kim, H.J.; Lee, J.H. Highly sensitive and selective gas sensors using p-type oxide semiconductors: Overview. Sens. Actuators B Chem. 2014, 192, 607–627. [Google Scholar] [CrossRef]
  11. Dey, A. Semiconductor metal oxide gas sensors: A review. Mater. Sci. Eng. B 2018, 229, 206–217. [Google Scholar] [CrossRef]
  12. Wetchakun, K.; Samerjai, T.; Tamaekong, N.; Liewhiran, C.; Siriwong, C.; Kruefu, V.; Wisitsoraat, A.; Tuantranont, A.; Phanichphant, S. Semiconducting metal oxides as sensors for environmentally hazardous gases. Sens. Actuator 2011, 160, 580–591. [Google Scholar] [CrossRef]
  13. Jerome McAleer, F.; Patrick Moseley, T.; Norris John, O.W.; Williams, D.E. Tin dioxide gas sensors. J. Chem. Soc. Faraday Trans. I 1987, 83, 1323–1346. [Google Scholar]
  14. Huízar-Padilla, E.; Guillén-Bonilla, H.; Guillén-Bonilla, A.; Rodríguez-Betancourtt, V.M.; Sánchez-Martínez, A.; Guillen-Bonilla, J.T.; Gildo-Ortiz, L.; Reyes-Gómez, J. Synthesis of ZnAl2O4 and Evaluation of the Response in Propane Atmospheres of Pellets and Thick Films Manufactured with Powders of the Oxide. Sensors 2021, 21, 2362. [Google Scholar] [CrossRef] [PubMed]
  15. Gao, X.; Zhang, T. An overview: Facet-dependent metal oxide semiconductor gas sensors. Sens. Actuators B Chem. 2018, 277, 604–633. [Google Scholar] [CrossRef]
  16. Yang, X.; Zhang, S.; Yu, Q.; Zhao, L.; Sun, P.; Wang, T.; Liu, F.; Yan, X.; Gao, Y.; Liang, X.; et al. One step synthesis of branched SnO2/ZnO heterostructures and their enhanced gas-sensing properties. Sens. Actuators B Chem. 2018, 2019, 415–423. [Google Scholar] [CrossRef]
  17. Saruhan, B.; Lontio Fomekong, R.; Nahirniak, S. Review: Influences of semiconductor metal oxide properties on gas sensing characteristics. Front. Sens. 2021, 2, 1–24. [Google Scholar]
  18. Qin, W.; Yuan, Z.; Gao, H.; Zhang, R.; Meng, F. Perovskite-structured LaCoO3 modified ZnO gas sensor and investigation on its gas sensing mechanism by first principle. Sens. Actuators B Chem. 2021, 341, 130015. [Google Scholar] [CrossRef]
  19. Ateia, E.E.; Arman, M.M.; Morsy, M. Synthesis, characterization of NdCoO3 perovskite and its uses as humidity sensor. Appl. Phys. 2019, 125, 883. [Google Scholar] [CrossRef]
  20. Guillén-Lopez, E.S.; Lopez-Urías, F.; Munoz-Sandoval, E.; Courel-Piedrahita, M.; Sanchez-Tizapa, M.; Guillén-Bonilla, H.; Rodríguez-Betancourtt, V.M.; Blanco-Alonso, O.; Guillén-Bonilla, A.; Moran-Lázaro, J.P. High performance isopropanol sensor based on spinel ZnMn2O4 nanoparticles. Mater. Today Commun. 2021, 26, 102138. [Google Scholar] [CrossRef]
  21. Liu, F.; Wang, B.; Yang, X.; Guan, Y.; Wang, Q.; Lianga, X.; Sun, P.; Wang, Y.; Lu, G. High-temperature NO2 gas sensor based on stabilized zirconia and CoTa2O6 sensing electrode. Sens. Actuators B Chem. 2017, 240, 148–157. [Google Scholar] [CrossRef]
  22. Singh, S.; Singh, A.; Singh, A.; Rathore, S.; Yadav, B.C.; Tandon, P. Nanostructured cobalt antimonate: A fast responsive and highly stable sensing material for liquefied petroleum gas detection at room temperature. RSC Adv. 2020, 10, 33770–33781. [Google Scholar] [CrossRef] [PubMed]
  23. Michel, C.R.; Martínez, A.H.; Jiménez, S. Gas sensing response of nanostructured trirutile-type CoSb2O6 synthesized by solution-polymerization method. Sens. Actuators B Chem. 2008, 132, 45–51. [Google Scholar] [CrossRef]
  24. Jamal, A.; Rahman, M.M.; Khan, S.B.; Faisal, M.; Akhtar, K.; Rub, M.A.; Asiri, A.M.; Al-Youbi, A.O. Cobalt doped antimony oxide nano-particles based chemical sensor and photo-catalyst for environmental pollutants. Appl. Surf. Sci. 2012, 261, 52–58. [Google Scholar] [CrossRef]
  25. Casillas-Zamora, A.; Guillen-Bonilla, J.T.; Guillén-Bonilla, A.; Rodríguez-Betancourtt, M.V.; Casallas-Moreno, Y.L.; Gildo-Ortiz, L.; Olvera-Amador, M.L.; Tomás, S.A.; Guillen-Bonilla, H. Synthesis of MnSb2O6 powders through a simple low-temperature method and their test as a gas sensor. J. Mater. Sci. Mater. Electron. 2020, 31, 7359–7372. [Google Scholar] [CrossRef]
  26. Singh, S.; Singh, A.; Singh, A.; Tandon, P. A stable and highly sensitive room-temperature liquefied petroleum gas sensor based on nanocubes/cuboids of zinc antimonate. RSC Adv. 2020, 10, 20349–20357. [Google Scholar] [CrossRef]
  27. 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]
  28. Guillén-Bonilla, H.; Flores-Martínez, M.; Rodríguez-Betancourtt, V.M.; Guillen-Bonilla, A.; Reyes-Gómez, J.; Gildo-Ortiz, L.; Olvera Amador, M.L.; Santoyo-Salazar, J. A Novel Gas Sensor Based on MgSb2O6 Nanorods to Indicate Variations in Carbon Monoxide and Propane Concentrations. Sensors 2016, 16, 177. [Google Scholar] [CrossRef]
  29. Roper, A.; Leverett, P.; Murphy, T.; Williams, P. Stabilities of byströmite, MgSb2O6, ordoñezite, ZnSb2O6 and rosiaite, PbSb2O6, and their possible roles in limiting antimony mobility in the supergene zone. Mineral. Mag. 2015, 79, 537–544. [Google Scholar] [CrossRef]
  30. Nagarajan, A.; Naraginti, S. Facile synthesis of N-MgSb2O6 trirutile antimonate and its enhanced photocatalytic performance. Int. J. Environ. Anal. Chem. 2020, 102, 7938–7952. [Google Scholar] [CrossRef]
  31. Arunkumar, N.; Naraginti, S. Facile synthesis of nanostructured trirutile antimonates M(II)Sb2O6 (M = Co, Cu, Ni, Fe) and its visible photocatalytic studies. Inorg. Nano-Met. Chem. 2022, 52, 151–160. [Google Scholar] [CrossRef]
  32. Singh, J.; Bhardwaj, N.; Uma, S. Single step hydrothermal based synthesis of M(II)Sb2O6 (M = Cd and Zn) type antimonates and their photocatalytic properties. Bull. Mater. Sci. 2013, 36, 287–291. [Google Scholar] [CrossRef]
  33. Litong, H.; Qiang, Z.; Fangfei, L.; Liang, L. Optical properties of trirutile structure MgTa2O6 single crystals grown by optical floating zone method. Mod. Phys. Lett. B 2020, 34, 2050281. [Google Scholar]
  34. LaMer, V.K.; Dinegar, R.H. Theory, Production and Mechanism of Formation of Monodispersed Hydrosols. J. Am. Chem. Soc. 1950, 72, 4847–4854. [Google Scholar] [CrossRef]
  35. Voorhees, P.W. The theory of Ostwald ripening. J. Stat. Phys. 1985, 38, 231–252. [Google Scholar] [CrossRef]
  36. Yu, H.; Regulacio, D.; Ye, M.D.; Han, M.Y. Chemical routes to top-down nanofabrication. Chem. Soc. Rev. 2013, 42, 6006–6018. [Google Scholar] [CrossRef]
  37. Esposito, S. “Traditional” sol-gel chemistry as a powerful tool for the preparation of supported metal and metal oxide catalysts. Materials 2019, 12, 668. [Google Scholar] [CrossRef] [PubMed]
  38. Kida, T.; Kuroiwa, T.; Yuasa, M.; Shimanoe, K.; Yamazoe, N. Study on the response and recovery properties of semiconductor gas sensors using a high-speed gas-switching system. Sens. Actuators B Chem. 2008, 134, 928–933. [Google Scholar] [CrossRef]
  39. Wang, C.; Yin, L.; Zhang, L.; Xiang, D.; Gao, R. Metal oxide gas sensors: Sensitivity and influencing factors. Sensors 2010, 10, 2088–2106. [Google Scholar] [CrossRef]
  40. Ramírez-Ortega, J.; Guillén-Bonilla, H.; Guillén-Bonilla, A.; Rodríguez-Betancourtt, V.M.; Sánchez-Martínez, A.; Guillén-Bonilla, J.T.; Gildo-Ortiz, L.; Huizar, E.; Reyes-Gómez, J. Synthesis of the oxide NiSb2O6 and its electrical characterization in toxic atmospheres for its application as a gas sensor. J. Mater. Sci. Mater. Electron. 2022, 33, 18268–18283. [Google Scholar] [CrossRef]
  41. Kim, J.H.; Mirzaei, A.; Kim, H.W.; Kim, S.S. Pd-functionalized core-shell composite nanowires for self-heating, sensitive, and benzene-selective gas sensors. Sens. Actuator A Phys. 2020, 308, 112011. [Google Scholar] [CrossRef]
  42. Avila-García, A.; Chaudhary, A.; Rojas-Chávez, H. Iridium oxide films as propane sensors. Thin Solid Films 2021, 724, 138617. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the system used in the dynamic tests in air–C3H8 flows at controlled concentrations and temperatures.
Figure 1. Schematic representation of the system used in the dynamic tests in air–C3H8 flows at controlled concentrations and temperatures.
Sensors 24 02147 g001
Figure 2. Diffractogram of powders of the MgSb2O6 calcined at 700 °C in air.
Figure 2. Diffractogram of powders of the MgSb2O6 calcined at 700 °C in air.
Sensors 24 02147 g002
Figure 3. (a) A characteristic spectrum of the MgSb2O6, (b) value of the oxide’s forbidden band.
Figure 3. (a) A characteristic spectrum of the MgSb2O6, (b) value of the oxide’s forbidden band.
Sensors 24 02147 g003
Figure 4. SEM images of powders of the MgSb2O6 calcined at 700 °C at magnifications of: (a) 9.42 kx, (b) 15 kx, (c) 20.00 kx, (d) 80 kx, (e) 80 kx, (f) 80 kx.
Figure 4. SEM images of powders of the MgSb2O6 calcined at 700 °C at magnifications of: (a) 9.42 kx, (b) 15 kx, (c) 20.00 kx, (d) 80 kx, (e) 80 kx, (f) 80 kx.
Sensors 24 02147 g004
Figure 5. Distribution of (a) length and (b) diameter of microrods of the MgSb2O6 calcined at 700 °C.
Figure 5. Distribution of (a) length and (b) diameter of microrods of the MgSb2O6 calcined at 700 °C.
Sensors 24 02147 g005
Figure 6. TEM images showing the nanostructured morphologies of the MgSb2O6 powders calcined at 700 °C: (a,b) agglomerated particles, (c,d) hexagonal morphology and (e,f) agglomerated polyhedral morphology.
Figure 6. TEM images showing the nanostructured morphologies of the MgSb2O6 powders calcined at 700 °C: (a,b) agglomerated particles, (c,d) hexagonal morphology and (e,f) agglomerated polyhedral morphology.
Sensors 24 02147 g006
Figure 7. Particle size distribution of the MgSb2O6 calcined at 700 °C.
Figure 7. Particle size distribution of the MgSb2O6 calcined at 700 °C.
Sensors 24 02147 g007
Figure 8. Dynamic response in C3H8 of the MgSb2O6 as a function of (a) the variation in electrical resistance, and (b) the dynamic sensitivity percentage.
Figure 8. Dynamic response in C3H8 of the MgSb2O6 as a function of (a) the variation in electrical resistance, and (b) the dynamic sensitivity percentage.
Sensors 24 02147 g008
Figure 9. Dynamic response of the MgSb2O6 in C3H8 at 400 °C. (a) Variation of electrical resistance with gas concentration, (b) sensitivity vs. gas concentration, (c) variation of sensitivity and electrical resistance as a function of gas concentration, (d) response and recovery times of MgSb2O6 thick films, and (e) linear fitting of the response as a function of C3H8 concentration.
Figure 9. Dynamic response of the MgSb2O6 in C3H8 at 400 °C. (a) Variation of electrical resistance with gas concentration, (b) sensitivity vs. gas concentration, (c) variation of sensitivity and electrical resistance as a function of gas concentration, (d) response and recovery times of MgSb2O6 thick films, and (e) linear fitting of the response as a function of C3H8 concentration.
Sensors 24 02147 g009
Figure 10. Dynamic response of the MgSb2O6 in C3H8 at 400 °C. (a,b) Variation in electrical resistance as a function of time, (c) dynamic sensitivity percentage as a function of time, (d) response and recovery times as a function of voltage, and (e) linear fitting of the response as a function of voltage.
Figure 10. Dynamic response of the MgSb2O6 in C3H8 at 400 °C. (a,b) Variation in electrical resistance as a function of time, (c) dynamic sensitivity percentage as a function of time, (d) response and recovery times as a function of voltage, and (e) linear fitting of the response as a function of voltage.
Sensors 24 02147 g010
Table 1. Mgsb2O6 thick films at different C3H8 concentrations.
Table 1. Mgsb2O6 thick films at different C3H8 concentrations.
Concentration
(ppm)
ΔR
(kΩ)
Sensitivity
(%)
Response Time
(min)
Recovery Time
(min)
15052.5061.096.153.22
30065.9588.803.103.08
40068.9097.652.40 2.97
60070.59112.812.132.85
Table 2. Variation of electrical resistance, sensitivity, and response and recovery times of the MgSb2O6 sensor at different operating voltages.
Table 2. Variation of electrical resistance, sensitivity, and response and recovery times of the MgSb2O6 sensor at different operating voltages.
Voltage
(V)
ΔR
(MΩ)
Sensitivity (%)Response Time (s)Recovery Time
(s)
523.7320.3580.4038.77
109.3419.0978.5336.10
159.0330.7476.6734.15
206.6733.7274.8331.38
255.1335.9371.0929.54
304.1839.5069.2427.69
353.3440.9765.6325.85
402.6542.7163.7022.16
452.2145.3560.9320.31
501.7849.0157.2518.45
Table 3. Performance comparison of propane sensors based on metal oxides.
Table 3. Performance comparison of propane sensors based on metal oxides.
MaterialGasConcentration
(ppm)
Sensitivity (%)Response Time (s)Recovery Time
(s)
Reference
ZnAl2O4Propane1000 ppm-176.0205.0[14]
CoSb2O6LPG5000 ppm1.9621.0234.0[22]
ZnSb2O6LPG5000 ppm 1.7341.095.0[26]
MgSb2O6Propane500 ppm61.66--[28]
IrO2-1000 ppm---[42]
MgSb2O6
MgSb2O6
Propane
Propane
600 ppm
560 ppm
112.81
49.01
127.8
57.2
171
18.5
This work
This work
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Guillén-Bonilla, H.; Guillén-Bonilla, J.T.; Rodríguez-Betancourtt, V.-M.; Ramírez-Ortega, J.A.; Morán Lázaro, J.P.; Guillén-Bonilla, A. Synthesis and Sensing Response of Magnesium Antimoniate Oxide (MgSb2O6) in the Presence of Propane Atmospheres at Different Operating Voltages. Sensors 2024, 24, 2147. https://doi.org/10.3390/s24072147

AMA Style

Guillén-Bonilla H, Guillén-Bonilla JT, Rodríguez-Betancourtt V-M, Ramírez-Ortega JA, Morán Lázaro JP, Guillén-Bonilla A. Synthesis and Sensing Response of Magnesium Antimoniate Oxide (MgSb2O6) in the Presence of Propane Atmospheres at Different Operating Voltages. Sensors. 2024; 24(7):2147. https://doi.org/10.3390/s24072147

Chicago/Turabian Style

Guillén-Bonilla, Héctor, José Trinidad Guillén-Bonilla, Verónica-María Rodríguez-Betancourtt, Jorge Alberto Ramírez-Ortega, Juan Pablo Morán Lázaro, and Alex Guillén-Bonilla. 2024. "Synthesis and Sensing Response of Magnesium Antimoniate Oxide (MgSb2O6) in the Presence of Propane Atmospheres at Different Operating Voltages" Sensors 24, no. 7: 2147. https://doi.org/10.3390/s24072147

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