Next Article in Journal
Hierarchical Visual Place Recognition Based on Semantic-Aggregation
Previous Article in Journal
Stimuli-Responsive Polymeric Nanosystems for Controlled Drug Delivery
Previous Article in Special Issue
Electrical Response of the Spinel ZnAl2O4 and Its Application in the Detection of Propane Gas
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Preparation of Powders Containing Sb, Ni, and O for the Design of a Novel CO and C3H8 Sensor

by
Jorge Alberto Ramírez-Ortega
1,
José Trinidad Guillén-Bonilla
2,
Alex Guillén-Bonilla
3,
Verónica María Rodríguez-Betancourtt
4,*,
Lorenzo Gildo-Ortiz
5,
Oscar Blanco-Alonso
5,
Víctor Manuel Soto-García
4,
Maricela Jiménez-Rodríguez
6 and
Héctor Guillén-Bonilla
1,*
1
Departamento de Ingeniería de Proyectos, CUCEI, Universidad de Guadalajara, M. García Barragán 1421, Guadalajara 44410, Jalisco, Mexico
2
Departamento de Electrónica, CUCEI, Universidad de Guadalajara, M. García Barragán 1421, Guadalajara 44410, Jalisco, Mexico
3
Departamento de Ciencias Computacionales e Ingenierías, CUVALLES, Universidad de Guadalajara, Carretera Guadalajara-Ameca Km 45.5, Ameca 46600, Jalisco, Mexico
4
Departamento de Química, CUCEI, Universidad de Guadalajara, M. García Barragán 1421, Guadalajara 44410, Jalisco, Mexico
5
Departamento de Física, CUCEI, Universidad de Guadalajara, M. García Barragán 1421, Guadalajara 44410, Jalisco, Mexico
6
División de Desarrollos Biotecnologicos, Centro Universitario de la Ciénega (CUCienéga), Universidad de Guadalajara (U. de G.), Av. Universidad No. 1115, Ocotlán 47810, Jalisco, Mexico
*
Authors to whom correspondence should be addressed.
Appl. Sci. 2021, 11(20), 9536; https://doi.org/10.3390/app11209536
Submission received: 6 September 2021 / Revised: 30 September 2021 / Accepted: 11 October 2021 / Published: 14 October 2021

Abstract

:
In this work, powders of NiSb2O6 were synthesized using a simple and economical microwave-assisted wet chemistry method, and calcined at 700, 800, and 900 °C. It was identified through X-ray diffraction that the oxide is a nanomaterial with a trirutile-type structure and space group P42/mnm (136). UV–Vis spectroscopy measurements showed that the bandgap values were at ~3.10, ~3.14, and ~3.23 eV at 700, 800, and 900 °C, respectively. Using scanning electron microscopy (SEM), irregularly shaped polyhedral microstructures with a size of ~154.78 nm were observed on the entire material’s surface. The particle size was estimated to average ~92.30 nm at the calcination temperature of 900 °C. Sensing tests in static atmospheres containing 300 ppm of CO at 300 °C showed a maximum sensitivity of ~72.67. On the other hand, in dynamic atmospheres at different CO flows and at an operating temperature of 200 °C, changes with time in electrical resistance were recorded, showing a high response, stability, and repeatability, and good sensor efficiency during several operation cycles. The response times were ~2.77 and ~2.10 min to 150 and 200 cm3/min of CO, respectively. Dynamic tests in propane (C3H8) atmospheres revealed that the material improved its response in alternating current signals at two different frequencies (0.1 and 1 kHz). It was also observed that at 360 °C, the ability to detect propane flows increased considerably. As in the case of CO, NiSb2O6’s response in propane atmospheres showed very good thermal stability, efficiency, a high capacity to detect C3H8, and short response and recovery times at both frequencies. Considering the great performance in propane flows, a sensor prototype was developed that modulates the electrical signals at 360 °C, verifying the excellent functionality of NiSb2O6.

Graphical Abstract

1. Introduction

In recent years, atmospheric pollution due to CO, CO2, NO2, SO2, and H2S has caused a decrease in air quality [1,2]. Emissions of toxic gases into the atmosphere by industries and internal combustion engines are considered the main sources that cause air pollution and the emergence of various respiratory diseases [2,3]. An alternative to somehow counteract these problems is to constantly control and monitor the concentration of toxic gases by designing simple electronic devices that meet the features of high-performance environmental gas sensors [4,5]. Several research groups and sensor manufacturing companies have worked on a wide variety of materials, mainly semiconductors, that could act as detectors for different polluting gases [4,6]. With such materials, toxic gas (such as CH4, LPG, and H2) leak detectors have been manufactured to monitor the emission of pollutants thrown by automobiles, etc. [6].
Nowadays, there is a great variety of gas sensors that can be classified according to their electrolytic, catalytic, electrochemical, and semiconductor properties [7,8]. The latter have been the most studied due to their excellent detection properties, having a high degree of sensitivity, good efficiency, wide detection of different gases (selectivity), as well as good performance at high operating temperatures (200–400 °C) in unsuitable humidity conditions [4,5,6]. To determine if a material can be used as a gas detector, it must detect low gas concentrations, detect a specific gas (selectivity), obtain a signal in the presence of the test gas in a relatively short time, and return to its initial state when finishing the detection process [9,10]. It has been reported in the literature that the most effective materials that can be applied as gas detectors are semiconductor metal oxides because they possess great advantages, such as thermal stability and high sensitivity [9,11]. It has been recognized that the detection ability of semiconductor oxides is based on reduction–oxidation reactions that occur between the material’s surface and the test gas [11,12]. As a result of these reactions, an electronic variation is produced on the surface that can be verified by the change in capacitance, electrical resistance, work function, or changes in its optical properties [9]. Among the wide variety of investigations on gas sensors, the binary compounds SnO2 and ZnO [13] and the ternaries SmCoO3 and ZnAl2O4 have been the most studied [14]. In fact, several studies have indicated that the trirutile-type antimonates ZnSb2O6 and CoSb2O6 are good candidates for application as gas sensors due to their interesting detection properties [15].
As for the nickel antimoniate, NiSb2O6, this is a trirutile-type oxide that conforms to the formula of the antimonates ASb2O6 (where A is replaced, in this case, by the Ni cation) [16]. This family of compounds has interesting optical, catalytic, and electrical properties that allow them to be applied, for example, in lithium-ion batteries and in many other fields [16,17]. These compounds have traditionally been synthesized using the ceramic method, obtaining different microstructures [17]. However, better results have been achieved using chemical methods, obtaining particles of the order of nanometers, which is a crucial factor in the ability to detect different types of gases [17,18]. We conducted extensive literature research, finding out that NiSb2O6 has already been studied as a potential sensor of static (for CO and C3H8) and dynamic (for LPG) atmospheres. However, we found no evidence that NiSb2O6 has yet been dynamically tested in CO and C3H8. Therefore, to contribute to the studies of this material as a gas sensor, NiSb2O6 powders were synthesized to evaluate their ability to detect carbon monoxide and propane gases at different concentrations and temperatures, achieving excellent results. Additionally, based on the ability of the studied oxide to detect high and low concentrations of both toxic gases, an electronic circuit was successfully designed for a high-efficiency and fast response gas sensor in such atmospheres.

2. Materials and Methods

2.1. Synthesis

The synthesis of the NiSb2O6 powders was performed using a microwave-assisted wet chemical method. An outline of the preparation process can be seen in Figure 1. For this, ~1.4533 g of Ni(NO3)2·6H2O (Sigma-Aldrich, 99%), ~2.2803 g of SbCl3 (Aldrich, 99%), 2 mL of ethylenediamine C2H8N2 (Sigma-Aldrich, 99%), and ethyl alcohol (CTR, 99.5%) as a solvent were used. Each salt was dissolved in 5 mL of ethyl alcohol, and the ethylenediamine was dissolved in 10 mL of the solvent. The three solutions were left stirring on a magnetic plate at room temperature for approximately 20 min. Once the salts were completely dissolved, the solutions were mixed with ethylenediamine to form a suspension, which was left under moderate stirring for 24 h at room temperature. After that, the solvent was evaporated by microwave radiation. This was accomplished using a domestic microwave oven (General Electric, model ST0912C01352). Evaporation consisted of applying a power of 70 W at exposure intervals of 60 s. After this, a green paste was obtained, which was dried in a Novatech muffle at 200 °C for 8 h. Then, the powders were calcined at 700, 800, and 900 °C for 5 h in the same muffle. The calcination ramp was 100 °C/h.

2.2. Physical Characterization

The crystal analysis of the NiSb2O6 powders was performed at 700, 800, and 900 °C through X-ray powder diffraction (Panalytical Empyream). Cuα radiation (λ = 1.5406 Å) was used, scanning from 10 to 90° at a rate of 0.02°/second. To determine the forbidden band gap of the calcined oxide powders at 700, 800, and 900 °C, a spectrophotometer (UV–Vis–NIR, model UV-3600 Plus) was used coupled to a deuterium and tungsten-halogen lamp. The absorbance spectra were taken in a range of 200 to 800 nm. The microstructure was analyzed with field-emission scanning electron microscopy (FE-SEM, Tescan MIRA 3 LMU) with an acceleration voltage of 10 kV in a high vacuum.

2.3. Tests in CO

To evaluate the capacity and efficiency of the NiSb2O6 powders calcined at 900 °C for detecting air–CO concentrations, experiments were carried out in static and dynamic atmospheres at different temperatures (100–300 °C). In both cases, pellets were made with 0.4 g of powder using a hydraulic press (Simplex-Italy Equip-25 Ton) compressing at 10 tons for 20 min. The dimensions of the pellets were 12 mm in diameter and 0.5 mm thick. Then, two colloidal silver paint ohmic contacts (Alfa Aesar 99%) were placed on the pellets for having good contact with the system’s electrodes (see Figure 2). Subsequently, the pellets were taken to the measurement chamber.
In static atmospheres, changes in NiSb2O6’s electrical resistance were recorded as a function of gas concentration (1, 50, 100, 200, and 300 ppm of CO) and operating temperature (100, 200, and 300 °C). The sensitivity magnitude in the static tests was estimated using the equation S = (GG − GO)/GO [19], where GG and GO are the electrical conductances in CO and air, respectively. The electrical conductance (1/electrical resistance) was considered according to reference [20]. For the tests in dynamic atmospheres at a constant temperature of 200 °C, an adaptation was made to the measurement system by placing a metal box (volume = 19 cm3) inside of it. The box had two perforations: one for the CO inlet and the other one for the test gas outlet and for introducing the electrodes in contact with the pellets’ surface (see in Figure 2). Subsequently, the gas was extracted using a pump (with a capacity of 10−3 Torr) incorporated in the measurement chamber. The detection tests consisted of supplying two different flows (300 and 400 cm3/min) of extra dry air (21% O2 + 79% N2) as stabilizing gas for the pellets’ surface. The CO flows were 100, 150, and 200 cm3/min. The partial pressure of the CO flows in static and dynamic atmospheres was monitored with a Leybold TM20 electronic detector. Changes in the test gas electrical resistance were recorded with a Keithley 2001 multimeter coupled to a data acquisition and control system using the LabView 8.6 software (National Instruments). Brooks Instruments mass flow regulators with capacities of 2600 cm3/min (GF100CXXC-SH452.6L) and 10 cm3/min (GF100CXXC-SH40010C) were used for the CO flows in the dynamic tests.

2.4. Tests in Propane

For the dynamic tests in air–propane atmospheres, thick films were prepared with the NiSb2O6 powders calcined at 900 °C. For this, a ceramic base was used with a circular cavity in the center and 4 ceramic columns around its perimeter. A hole was made in the middle of each column through which high purity platinum wires (0.006 in diameter) were introduced, functioning as electrodes connected to the detection system. For the preparation of the thick films, approximately 0.4 g of the material was placed in a vial with 3 mL of isopropyl alcohol and dispersed with an ultrasonic generator (Brason 2510 Ultrasonic). The dispersed material was then drip deposited into the circular cavity, forming films with dimensions of ~0.5 mm thick and ~0.3 mm in diameter. Subsequently, the films were subjected to a heat treatment at 300 °C for 4 h in a programmable muffle (Vulcan, model 5‒550).
The films were placed on an aluminum plate inside a quartz tube and in a tubular oven with programmable temperature control (Lindberg/blue). Air and propane flows were controlled by the same mass flow regulators as for CO. The changes in the sensitivity magnitude |Z| were recorded with an Agilent 4263B multimeter controlled through the LabView 8.6 software (National Instruments).

3. Results and Discussion

3.1. XRD Analysis

Figure 3 shows the crystal evolution of the NiSb2O6 powders calcined at 700, 800, and 900 °C. In these diffraction patterns, it is observed that the material showed its characteristic peaks at the three calcination temperatures.
However, portions of secondary phases to the compound’s main phase were identified for each heat treatment. At 700 °C, although the peaks of the material’s crystalline structure were obtained, secondary phases were identified for Sb2O3 (PDF # 71‒0383) and Sb2O5 (PDF # 50‒1376), located at the 2θ points 28.14°, 28.96°, 30.11°, and 33.78° corresponding to the fewer intensity peaks. At 800 °C, crystal planes of NiSb2O6 with more intense peaks of greater purity and crystallinity were observed. These reflections were identified at 2θ = 19.22°, 21.43°, 27.14°, 33.50°, 34.99°, 38.80°, 40.19°, 44.73°, 53.22°, 56.03°, 60.23°, 62.77°, 63.32°, 67.20°, 67.78°, 73.89°, 80.89°, and 86.83° (PDF # 38‒1083). In addition, a small phase associated with NiO (PDF # 44‒1159) was located at 2θ = 37.19°. At 900 °C, planes attributed to NiO were again found, but with less intensity than at 800 °C. As expected, comparing the results of Figure 3 with PDF # 38‒1083, it was found that NiSb2O6 belongs to the trirutile-type family of materials [21], with a tetragonal structure (network parameters: a = 4.641 Å, c = 9.223 Å) and space group P42/mnm [22].
Based on the diffractograms’ peaks, the crystallite size was estimated with Scherrer’s equation [23] as follows:
t = 0.9 λ β c o s θ
where λ is the wavelength of the radiation (in this case, Cu = 1.5406 Å), β is the width of the peak measured at half the maximum intensity, and θ is the Bragg angle. For the calculation, the most intense peak corresponding to the plane (110) located at 2θ = 27.14° was considered. From the diffractograms depicted in Figure 4a, at 700 °C, the crystal size was estimated at ~36.43 ± 1.65 nm; at 800 °C, it was ~38.40 ± 0.96 nm; and at 900 °C, it was ~44.02 ± 0.50 nm. Accordingly, it was verified that when increasing the temperature, the crystallite size increased. This is attributed to the fact that a particles’ agglomeration occurred due to the increasing calcination temperature [24,25] and to the material’s residence time in the muffle at the given temperature. It is worth mentioning that the time for the thermal treatments was relatively short (5 h), thus achieving that meant the crystal size stayed in the order of nanometers [25].
Figure 4b,c show a projection of the octahedral gap formed by the Ni2+ and Sb5+ cations, which are coordinated with the oxygen anions O2−. Additionally, Figure 4c depicts the atomic arrangement of NiSb2O6’s crystal structure seen from the plane of greatest intensity (110). Both structures were obtained with the Diamond software using PDF # 38‒1083 and reference [26]. It is important to mention that the cations’ arrangement forms a triplicate rutile-type structure in the c-axis and, according to the literature, this modification on the c-axis of NiSb2O6 represents a trirutile-type structure [16,17,27].
Comparing our results with those of reference [17], where the same compound was synthesized, we obtained the crystalline phase at a shorter exposure time (5 h) than that of such reference, where at 800 °C, the residence time was of 3 days. Furthermore, in our case, a shorter preparation time was required than those reported in references [16,17], where they used the solid-state reaction method, requiring significantly longer preparation times.

3.2. UV–Vis Analysis

Figure 5a shows the characteristic UV‒Vis absorption spectra of NiSb2O6 calcined at 700, 800, and 900 °C in a wavelength range of 200 to 800 nm (1.55 to 6.2 eV). NiSb2O6’s absorption bands were identified in the range from 200 to 500 nm [16,28], which are characteristic of semiconductor oxides with a trirutile-type structure [29,30].
In the range from 600 to 700 nm, the bands corresponding to the electronic interactions occurring in the oxygen–metal (O‒Ni) bond [31,32,33] were recorded. To estimate the values of the forbidden bandwidth (Eg) of the NiSb2O6 powders, Tauc’s equation was used [34].
( E ) = A ( E E g ) 1 2 E ,
where E is the energy of the incident photon, α is the optical absorption coefficient, Eg is the forbidden bandwidth, and A is the proportionality constant. In our case, a direct transition (n = 1/2) was considered for NiSb2O6 [16,28]. Tauc’s equation was used to elaborate the graph of (αℎυ)2 vs. E, superimposing a straight line on the spectrum’s final part (Figure 5b). With this, the bandgap energy of the calcined material at 700, 800, and 900 °C was estimated at ~3.10 ± 0.02, ~3.14 ± 0.04, and ~3.23 ± 0.03 eV, respectively. According to Figure 5, the material’s crystal size strongly influenced the variation of the forbidden band [25,28]. In our case, as the calcination temperature increased, the forbidden band value had a slight increase. This is attributed to the fact that, when the calcination temperature rose from 700 to 900 °C, the crystal size increased (see Figure 4a), provoking the energy variations of NiSb2O6.
Then, the crystal size, whether it increases or decreases, has a very significant effect on the compound’s bandgap value. In the literature, bandgap values of ~2.7–2.8 eV at a calcination temperature of 700 °C and ~2.9 eV at 800 °C for materials such as our oxide were reported [16]. In reference [28], a bandgap value of ~3.92 eV was estimated for NiSb2O6. In comparison, our results are within the accepted energy range by different studies for the same or similar material to ours (see Figure 5) [16,25,29,35].

3.3. SEM Analysis

Five SEM micrographs of NiSb2O6 powders calcined at 900 °C are shown in Figure 6. These images show that the material’s surface was made up of polyhedron-shaped particles and irregular particles of different sizes (Figure 6a–e). Figure 6a,b show a large agglomeration of these particles, which were distributed over all the analyzed areas. Additionally, it is clear to observe that the polyhedra were composed of smaller, agglomerated particles, which gave rise to the microstructures shown in Figure 6c–e.
The morphologies shown in Figure 6 are attributed to the fact that the microstructural features are strongly related to the synthesis method [24]. Reference [36] reported that the preparation route, the calcination temperature, and the application of microwave radiation favor obtaining specific morphologies, such as octahedral structures, spheres, wires, and hexagonal nanostructures, among others. On the other hand, other studies informed that chemical methods are desirable in the formation of specific microstructures adequate for gas detection because the nucleation and growth processes that take place with them favor the control of the particles’ nanometric size [37,38]. In our case, the use of a wet-chemistry process assisted with microwave radiation in the presence of ethylenediamine favored the growth of morphologies such as those presented in Figure 6a‒e. The results followed the nucleation and growth principles proposed in reference [18].
From Figure 6, the particle size was estimated considering areas of the material’s surface where the particles were clearly identifiable. A size in the range of 40–150 nm was estimated, with an average of ~92.30 nm and a standard deviation of ~± 24.64 nm (see Figure 7a). The size of the octahedral polyhedra was estimated in the range of 80–240 nm, with an average of ~154.78 nm and a standard deviation of ~±32.55 nm (Figure 7b).

3.4. CO Analysis

The gas response was analyzed with the material synthesized at 900 °C, where a minor presence of the secondary phase was obtained. To evaluate NiSb2O6’s detection capacity at different temperatures (100–300 °C) and CO concentrations (1–300 ppm), tests were conducted in static atmospheres using pellets made with powders of the material calcined at 900 °C. The results are shown in Figure 8 as a function of the test gas concentration (Figure 8a) and operating temperature (Figure 8b).
According to the analysis, the pellets did not show changes in electrical resistance at 100 °C regardless of the increase in the CO concentration. In contrast, at 200 °C, a small variation in resistance was recorded, obtaining a sensitivity magnitude of ~3.14 at 300 ppm CO. When increasing the operating temperature to 300 °C, very significant changes in electrical resistance took place, with a maximum sensitivity magnitude of ~72.67 at 300 ppm of CO. During the experiments, it was observed that the increase in the CO concentration caused the increase in the pellets’ response (see Table 1).
This was due to the fact that when the test gas was in contact with the oxygen on the pellets’ surface, it oxidized, causing a transfer of charge carriers on the material’s surface [11,39,40]. In the literature, it is reported that the increase in response in materials such as NiSb2O6 at temperatures above 150 °C is due to the O and O 2 ionic oxygen species [37,39], which are very reactive at relatively high temperatures (in our case, at 300 °C) [11,39]. At temperatures below 200 °C, the oxygen species that appear the most are of the O 2 type [11,39]. Such species are less reactive at temperatures below 100 °C because the thermal energy supplied is not enough to favor chemical reactions between the oxygen species, the CO, and the surface of the pellets, causing low kinetic interaction between the material nanoparticles and the CO molecules [40,41,42]. This resulted in not recording significant increases in the pellets’ response at 100 °C (see Figure 8a,b). It is important to mention that the excellent response and good stability recorded at 300 °C were mainly associated with the microstructural features obtained, which led to an increase in catalytic sites for the gas reaction with the nanoparticles’ surface [43,44,45].
The dependence of gas sensing on the operating temperature is due to several facts. One of them is that the type of oxygen species adsorbed on the oxide depends on temperature [28]. As mentioned, atomic species, which are more reactive than molecular species, predominate at temperatures above 150 °C [13]. Another fact is that the oxidation reaction is an activated process, where its rate increases with temperature. Thermal energy is needed to activate the adsorption of ionized oxygen species and overcome the barriers of the sensing reactions [20]. Additionally, desorption and diffusion processes are also temperature dependent. Thus, the sensitivity can rise appreciably with increasing temperature [24]. Different literature reports have shown that the gas response at low temperatures starts to disappear at around 200 °C and is replaced by the typical sensitivity response at high temperatures [13,20]. The optimum temperature depends on the sensing material and the sensed gas. Therefore, as in previous reports [29,43], a noticeable response variation was observed at 300 °C. Comparing our results with references [29,42,43,46], where similar experiments were carried out in static gas atmospheres, we found that our pellets had a greater response and a better performance in detecting CO at different operating temperatures (100–300 °C).
We also carried out tests in dynamic atmospheres as a function of time, applying direct current (d. c.) signals at 200 °C. For this, extra dry air flows (21% O2 + 79% N2; 400 cm3/min for 7 min at the test’s start and 300 cm3/min for 13 min at the test’s end) were supplied to the measuring chamber to stabilize the pellets’ surface, while the CO flows were 100, 150, and 200 cm3/min. The total time for the dynamic cycles was approximately 20 min. The results are shown in Figure 9.
It is observed in Figure 9a that, as the CO flow increases, pellets’ electrical resistance decreases significantly. The estimated drop in electrical resistance was ~1008.59 MΩ at 100 cm3/min, ~1273.27 MΩ at 150 cm3/min, and ~1474.49 MΩ at 200 cm3/min. The response and recovery times were ~2.24 and ~1.39 min at 100 cm3/min, and ~2.77 and ~2.00 min at 150 cm3/min. At 200 cm3/min, the response time was ~2.10 min. The response and recovery times were calculated according to reference [47], where 90% of the electrical resistance in the test gas and 10% of the resistance in the air are considered. We observed that the response and recovery times in the dynamic tests (see Figure 9) were short. That was because by introducing the CO into the measurement chamber, it reacted immediately with the material’s surface due to the effect of temperature, with the consequent immediate change in electrical resistance (that is, the response time). On the other hand, when the test gas stopped injecting and the air was introduced (during the recovery time), the material tended to return almost immediately to its initial resistance values, which meant that the detection process was reversible.
The changes in electrical resistance shown in Figure 9a are typical for n-type semiconductors [11]; however, it is reported in the literature that NiSb2O6 tends to be a p-type semiconductor when exposed to atmospheres (for example, CO2 and LPG) such as ours (CO) [28]. We also made measurements in large air flows (400 and 300 cm3/min), obtaining a supersaturation of oxygen species ( O ) on the pellets’ surface, originating an oxidation process with the test gas [45]. This caused all the charge carriers (electrons and holes) to interact with each other, allowing competition between them and resulting in a majority flow of electrons on the pellets’ surface [48]. Another factor that influenced the trends shown in Figure 9a was the operating temperature [12,40]. According to the literature, the oxygen species on the pellets’ surface are adsorbed by the effect of temperature (in our case at 200 °C) [11], releasing electrons that are then captured to form ionized species ( O 2 , O , or   O 2 ) [49]. This means that, when the pellets’ surface comes into contact with the CO molecules at 200 °C, the latter react with the ionized species, causing an electron transfer during the adsorption and desorption of the gas molecules on the material’s surface [11] and, with it, the decrease in electrical resistance depicted in Figure 9a. Shifts in percent sensitivity as a function of time are shown in Figure 9b. The response (sensitivity) was determined considering the variations in electrical resistance for 100, 150, and 200 cm3/min of CO at 200 °C when the surface of the pellets was stabilized by means of extra-dry air flows, as in the case shown in Figure 9a. To obtain the graphs of Figure 9b, the following formula was considered [50]: S(%) = (GG − GO)/GO × 100, where GG and GO are the electrical conductances (1/electrical resistance) in CO and air, respectively. From the results, for the flow of 100 cm3/min, the estimated sensitivity was ~57.10% (first curve). For the 150 cm3/min flow, the sensitivity increased to ~84.77% (second curve). Finally, for the flow of 200 cm3/min, the percentage obtained was ~113.75%. As expected, increasing the CO flow increased the pellets’ percent sensitivity. With this, it can be corroborated that the increase in time-dependent sensitivity in a material such as NiSb2O6 is closely related to the variation in the test gas concentration [29]. Indeed, it is reported in the literature that as the concentration of the test gases increases, the sensitivity of the material increases [32,34]. This is due to the fact that the higher the gas concentration value is, the more easily it reacts with the oxygen species ( O ) on the pellets’ surface due to the temperature effect (in our case, at 200 °C), causing an increase in the gas oxidation and, thus, an electron release, causing NiSb2O6’s sensitivity to rise [20,34]. Some studies report that the adsorption of oxygen species on a semiconductor’s surface due to the operating temperature effect is what causes the high sensitivity of materials such as ours [28,29,35,48,49,51]. In particular, our results (displayed in Figure 9) show better thermal stability, high sensitivity, reproducibility, and efficiency in the detection of different CO concentrations compared to those reported in references [14,28,29,31,34,35,42,46,49], where tests in static and dynamic CO, CO2, and LPG atmospheres were performed.
A mechanism that explains the interaction between a material such as NiSb2O6 and a gas such as CO is based on the drop of the depletion layer due to oxygen adsorption [51,52]. When a gas such as CO is adsorbed on the material’s surface, it reacts with the chemisorbed oxygen species [52], producing CO2 molecules and a release of electrons that move to the conduction band [20]. This produces changes in the material’s conductance (or electrical resistance) [53] and, therefore, variations in the electrical response in CO atmospheres (see Figure 9a,b). Therefore, since the reaction mechanism between a gas such as CO and the surface of a semiconductor such as NiSb2O6 depends largely on the operating temperature [20,52], a reaction that may be occurring on the pellets’ surface at 200 °C is as follows [20,41,49,52,54]:
      O 2 + 2 C O   2 C O 2 + e
where O 2 is the oxygen ions on the material’s surface and e is the electrons in the bulk material [20,49,54]. Furthermore, the response depicted in Figure 9 depended on the partial pressure of the oxygen in the extra dry air [44,49,54], as well as on the porosity, morphology, and particle size [55]. If the average particle size decreases, NiSb2O6’s surface area increases [4] and, consequently, the gas–solid interaction is favored [55]. In our case, we obtained a material with an average particle size of ~95.63 nm, which meant better oxygen adsorption and desorption on the material’s surface and, therefore, a high NiSb2O6 response, as shown in Figure 9a,b.

3.5. Propane Analysis

Figure 10 shows the results of the dynamic response at 0.1 and 1 kHz of NiSb2O6 films exposed to propane at 360 °C. In Figure 10a,b, the results are shown for films stabilized through a flow of extra dry air (1500 cm3/min) for 5 min. Then, without removing the air, 600 ppm of propane was injected into the measuring chamber for 5 min, recording the decrease in sensitivity magnitude |Z| by varying the frequency from 0.1 to 1 kHz. The values of |Z| at 0.1 kHz were in the range of 6.6–7.05 MΩ, with response and recovery times of 42.96 and 35.56 s, respectively. At 1 kHz, the |Z| values were 1.74–1.75 MΩ, with response and recovery times of 11.91 and 5.88 s, respectively. The response and recovery times were calculated for both frequencies considering 90% of the response in the test gas and 10% of the value when exposed only to air [47]. Although NiSb2O6 is a p-type semiconductor [28], the trend in Figure 10a,b is attributed to the fact that the decrease in |Z| is caused by the continuous flow of oxygen-rich air (21% O2), which causes oxidation ( O ) [48,54,56] and an oxygen saturation ( O and O 2 ) [29,35,43,54,57] that reacts with the propane on the films’ surface, causing the excellent response shown in Figure 10a–e.
We observed that the good dynamic response of the NiSb2O6 films in propane atmospheres was caused by the rapid adsorption of the gas that reacted with the film’s surface, causing an electron exchange between the oxygen species adsorbed on the surface and the surface itself [48]. It has been reported in the literature that the electrons on the material’s surface are the most probable cause of the electrical response variation and, therefore, of the improvement in the sensor response [11,48,50,58]. The variations in the impedance magnitude (|Z|) and in the response and recovery times were due to the chemical reaction between the surface of the thick films and the propane gas. The reaction occurred because there was a high kinetic activity of the gas molecules due to the operating temperature, causing an immediate response (response time) to 600 ppm of the gas. As in the case of CO, when the propane stopped injecting and the air was introduced into the system, the impedance values tended to immediately return to their stable values (recovery time). It was corroborated that at 0.1 and 1 kHz, the response and recovery times decreased significantly compared to other semiconductors tested as gas sensors [14,28,34,58].
The reaction mechanism that could explain the adsorption and desorption of propane gas molecules on a semiconductor’s surface, such as the one studied here, is not yet fully established. Some authors report that the propane molecules on the film’s surface react with adsorbed oxygenated species, producing CO2, water vapor, and the release of electrons on the surface [59]. Other authors believe that a reaction that could be occurring between a given semiconductor and propane gas could be the following: C3H8 + 10 O → 3CO2 + 4H2O + e [35,60]. This means that when the propane gas is adsorbed, it dissociates on the material’s surface because of the temperature (in our case at 360 °C), causing changes in the sensor’s electrical response (see Figure 10a–e).
At this point, it is convenient to address several sensing mechanisms of metal oxide semiconductors in more detail. These mechanisms can be classified into the following two categories: microscopic and macroscopic [61]. The first one comprises the mechanisms of Fermi level control, grain boundary barriers control, and electron-depletion and hole-accumulation layers. Additionally, the changes in electrical properties are accompanied by changes in physical properties such as energy bands and work functions.
The macroscopic category includes mechanisms of adsorption-desorption patterns, bulk resistance control, and gas diffusion control. In the case of bulk resistance control, the change in the electrical properties of the sensors can be caused by phase transformations, which occur in certain types of materials. The gas diffusion control mechanism dictates that the most important factor in the diffusion process is related to the microstructure of the material, which, in turn, affects the surface chemical reactions and the sensing properties [61,62].
The adsorption/desorption model is the current conventional mechanism, which states that the electron capture by adsorbed species, and their subsequent desorption, is responsible for changes in the sensor’s electrical properties. This model includes oxygen adsorption, chemical adsorption/desorption, and physical adsorption/desorption. The theory of electron depletion and hole accumulation layers is an extension of the oxygen adsorption model. As mentioned in the discussion section, the material’s sensitivity is primarily related to the oxygen chemisorption, where the formed ionic or molecular species are temperature dependent. Chemical adsorption/desorption also extends to the material’s exposure to different gases [11,20,61,62].
Various theories have been developed on the adsorption of gases. For example, Freundlich and Kuster’s mathematical modeling of an isotherm for gas adsorbates, Langmuir’s isotherm theory of adsorption of the unimolecular layer, the BET theory of multilayer adsorption, and dynamic adsorption and desorption processes based on Langmuir’s theory [63]. Additionally, the adsorption of toxic small molecules has been theoretically investigated through adsorption configurations, adsorption energy, charge transfers, and electronic properties. Eight possible orientations of CO adsorption are documented for pentagraphene-based materials, where chemical adsorption is favored more than physical adsorption [64]. Likewise, in CO sensing using GaN, four stable gas adsorption configurations have been reported [65]. Theoretical studies on propane adsorption have also been conducted, where it did not have a significant site preference on the studied surfaces [66].
For the design of our 600-ppm propane detector, changes in | Z | at 1 kHz and 360 °C were considered. The calculations shown in Figure 10c–e were conducted with the following formula:
Z ( t ) = | Z ( t ) | e θ ( t ) = r ( t ) + i X c = 1 2 π f C
where i is an imaginary operator, r ( t ) = | Z ( t ) | cos θ ( t ) is the resistance (Figure 10c), X c = | Z ( t ) | sen θ ( t ) is the reactance (Figure 10d), and C ( t ) = 1 2 π f | Z ( t ) | sen θ ( t ) is the capacitance in faradays (Figure 10e).
Considering the values in Figure 10c–e, we developed our electrical circuit using the following equation:
z s ( t ) = z c + z d ( t ) = z c + r d ( t ) r d ( t ) + X d ( t )
where z s ( t ) is the sensor’s impedance, z c is the constant impedance, and z d ( t ) is the sensor’s dynamic impedance in the test gas. Therefore, if we consider an operation point at [ t ¯ ,   z ( t ¯ ) ] and [ t ¯ ,   θ ( t ¯ ) ] [67], the impedance would be as follows:
        z s ( s ) = z c + z d ( s ) = z c + r d s r d s + X d s .
Substituting the identity X d s = 1 s C d s , the impedance takes the following form:
z s ( s ) = z c + z d ( s ) = z c + r d s C d s s r d s C d s s + 1 .
If t ¯ = 300   s , we obtain z c = 1754362 e 74.3106 , r d s = 18167   Ω , C d s = 9.37   pF , and the impedance as follows:
z s ( s ) = 1754362 e 74.3106 + 0.0001704 s 0.0001704 s + 1 .
From Equation (8) and using the impedance z d ( s ) , we obtained the Bode diagram shown in Figure 11a.
Comparing this diagram with NiSb2O6’s electrical response (Figure 10a,b), the results are consistent because | Z | decreases when the frequency increases from 0.1 to 1 kHz.
The detector’s electronic diagram in propane atmospheres is shown in Figure 11b. NiSb2O6 pellets working as detectors were placed in one of the Wheatstone bridge’s arms, whose output voltages were V a = z 1 z s + z 1 V a c and V b = z 2 z c a + z 2 V a c , where z c a is the variable impedance employed to calibrate the bridge, z 1 and z 2   are impedances, and V a c is the supply voltage. Voltages V a and V b are the input signals of an adder‒subtractor circuit, whose output voltage is V a b = V a V b = z 1 z s + z 1 V a c z 2 z c a + z 2 V a c . This output voltage is the input signal of the comparator circuit, whose inverter’s input and supply V c c were connected to a common for avoiding negative voltages.
The detector’s operating principle is shown in Figure 11c, which is based on the following two basic stages: calibration and detection. In the calibration stage, the sensor is exposed to the air-atmosphere, where the Wheatstone bridge was calibrated through the variable impedance z c until the condition z 1 = z 2 = z c = z s was met, yielding V a b = 0   V and no alarm signal: V A l a r m 0   V . In the test gas detection stage, the sensor showed electrical variations in propane atmospheres, whereby the Wheatstone bridge was decalibrated, causing V a b 0 , V a > V b , and, as a result, the comparator circuit to be positively saturated with V A l a r m V c c , which was the detector’s alarm signal in the test gas.
In general, it is well known that the performance of semiconductors such as ours (NiSb2O6) depends, to a great extent, on the microstructure and the crystal size. The decrease in crystal size in semiconductors increases the specific surface area [10,20,24]. Consequently, there is an increase in the number of potentially active catalytic sites that favor the high adsorption of gases on the semiconductor’s surface. That is because the space charge zone, or depletion layer, which is a function of charge carriers, has a thickness LS of less than 100 nm [61,62]. According to some references, the charged outer layer depends on the partial pressure and the concentration of the test gas [34,55]. However, the conductance (or electrical resistance) is closely related to the crystal size D [34,68], which means that when D is less than 2L (in our case 44.02 ± 0.50 nm), the response of semiconductor oxides is significantly increased [62]. Then, if the crystal (or particle) size is D < 2L, the crystals will participate in the electronic transport during sensing [34,54,68,69,70], provoking variations in the electrical resistance and, therefore, increases in the sensor’s response, high electrical sensitivity, thermal stability, and high efficiency in the gas detection [68,70]. Thus, it is evident that obtaining nanometric particle sizes during the synthesis process will positively impact the detection properties, as reported in this work (see Figure 8, Figure 9 and Figure 10).

4. Conclusions

The synthesis of NiSb2O6 powders was conducted successfully by means of a wet chemistry process assisted with microwave radiation. This method allowed us to obtain different morphologies for use as a gas sensor (in this work: CO and propane). Using UV–Vis, forbidden bandgap values of ~3.06, ~3.14, and ~3.25 eV were estimated for the powders of NiSb2O6 calcined at 700, 800, and 900 °C, respectively. It was observed that the bandgap value was closely related to the crystallite size (~32, ~40, and ~44 nm at 700, 800, and 900 °C, respectively) and the calcination temperature in the synthesis process. By means of scanning electron microscopy (SEM), microstructures in the form of polyhedra and irregular particles of different sizes, distributed over the entire surface of the material, were observed. This was largely due to the chelating agent (ethylenediamine), the heat treatment, and the synthesis method. In static and dynamic tests using pellets made with powders of the oxide, a good response was obtained at different concentrations and flows of CO and operating temperatures, obtaining a maximum sensitivity of ~72.4 at 300 ppm of CO and 300 °C. In dynamic tests conducted using thick films prepared with the material’s powders, an excellent response was obtained at 360 °C in 600 ppm of propane and applying an alternating current (at 0.1 and 1 kHz). Good stability, reproducibility, performance, and short response and recovery times were found. The excellent material’s dynamic response in propane allowed us to design an efficient prototype for detecting atmospheres of such gas at different working frequencies. The synthesis method used and the high material’s crystallinity contributed greatly to obtaining excellent results in the CO and propane detection tests; therefore, we can affirm that NiSb2O6 calcined at 900 °C is a strong candidate for being applied as a gas detector.

Author Contributions

Formal analysis, J.T.G.-B. and O.B.-A.; investigation, J.A.R.-O. and V.M.S.-G.; methodology, J.A.R.-O., J.T.G.-B., A.G.-B., V.M.R.-B., L.G.-O., V.M.S.-G. and M.J.-R.; project administration, H.G.-B.; resources, A.G.-B. and O.B.-A.; supervision, H.G.-B.; validation, A.G.-B.; visualization, V.M.R.-B. and L.G.-O.; writing—original draft, J.A.R.-O., J.T.G.-B., A.G.-B., V.M.R.-B., L.G.-O., O.B.-A., V.M.S.-G., M.J.-R. 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 was funded by the National Council of Science and Technology of Mexico (CONACYT), grant number 888575.

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 authors upon request.

Acknowledgments

Jorge Alberto Ramírez Ortega thanks CONACyT for a scholarship. Likewise, we thank M. de la Luz Olvera Amador, Juan Pablo Morán Lázaro, Juan Reyes Gómez, Emilio Huízar Padilla, and Miguel Ángel Luna Arias for their technical assistance. This work was conducted following the research guidelines for “Nanostructured Semiconductor Oxides” of the CUCEI’s UDG-CA-895 “Nanostructured Semiconductors” academic group, University of Guadalajara.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bergmann, S.; Li, B.; Pilot, E.; Chen, R.; Wang, B.; Yang, J. Effect modification of the short-term effects of air pollution on morbidity by season: A systematic review and meta-analysis. Sci. Total Environ. 2020, 716, 136985. [Google Scholar] [CrossRef] [PubMed]
  2. Manisalidis, I.; Stavropoulou, E.; Stavropoulos, A.; Bezirtzoglou, E. Environmental and Health Impacts of Air Pollution: A Review. Front. Public Health 2020, 8, 14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Kheirbek, I.; Haney, J.; Douglas, S.; Ito, K.; Matte, T. The contribution of motor vehicle emissions to ambient fine particulate matter public health impacts in New York City: A health burden assessment. Environ. Health 2016, 15, 1–14. [Google Scholar] [CrossRef] [Green Version]
  4. Majhi, S.M.; Mirzaei, A.; Kim, H.W.; Kim, S.S.; Kim, T.W. Recent advances in energy-saving chemiresistive gas sensors: A review. Nano Energy 2021, 79, 105369. [Google Scholar] [CrossRef]
  5. Nikolic, M.V.; Milovanovic, V.; Vasiljevic, Z.Z.; Stamenkovic, Z. Semiconductor gas sensors: Materials, technology, design, and application. Sensors 2020, 20, 6694. [Google Scholar] [CrossRef]
  6. Nayyar, A.; Puri, V.; Le, D.-N. A Comprehensive Review of Semiconductor-Type Gas Sensors for Environmental Monitoring. Rev. Comput. Eng. Res. 2016, 3, 55–64. [Google Scholar] [CrossRef] [Green Version]
  7. Morrison, S.R. Selectivity in semiconductor gas sensors. Sens. Actuators 1987, 12, 425–440. [Google Scholar] [CrossRef]
  8. Korotcenkov, G. Handbook of Gas Sensor Materials; Burnaby, DC, Canada, 2014; Volume 2, ISBN 9781461473886. [Google Scholar]
  9. Liu, X.; Cheng, S.; Liu, H.; Hu, S.; Zhang, D.; Ning, H. A survey on gas sensing technology. Sensors 2012, 12, 9635–9665. [Google Scholar] [CrossRef] [Green Version]
  10. Gautam, Y.K.; Sharma, K.; Tyagi, S.; Ambedkar, A.K.; Chaudhary, M.; Pal Singh, B. Nanostructured metal oxide semiconductor-based sensors for greenhouse gas detection: Progress and challenges. R. Soc. Open Sci. 2021, 8. [Google Scholar] [CrossRef]
  11. Shankar, P.; Rayappan, J. Gas sensing mechanism of metal oxides: The role of ambient atmosphere, type of semiconductor and gases-A review. Sci. Lett. J. 2015, 4, 126. [Google Scholar]
  12. Amiri, V.; Roshan, H.; Mirzaei, A.; Neri, G.; Ayesh, A.I. Nanostructured metal oxide-based acetone gas sensors: A review. Sensors 2020, 20, 3096. [Google Scholar] [CrossRef] [PubMed]
  13. Saruhan, B.; Lontio Fomekong, R.; Nahirniak, S. Review: Influences of Semiconductor Metal Oxide Properties on Gas Sensing Characteristics. Front. Sensors 2021, 2, 1–24. [Google Scholar] [CrossRef]
  14. Delgado, E.; Michel, C.R. CO2 and O2 sensing behavior of nanostructured barium-doped SmCoO3. Mater. Lett. 2006, 60, 1613–1616. [Google Scholar] [CrossRef]
  15. Balasubramaniam, M.; Balakumar, S. A review on multifunctional attributes of zinc antimonate nanostructures towards energy and environmental applications. Chem. Pap. 2020, 74, 55–75. [Google Scholar] [CrossRef]
  16. Alireza, H. Effects of reaction temperature and raw material type on optical properties and crystal phase growth of solid state synthesized NiSb2O6 nanomaterials. Adv. Mater. Process. 2017, 5, 56–65. [Google Scholar]
  17. Larcher, D.; Prakash, A.S.; Laffont, L.; Womes, M.; Jumas, J.C.; Olivier-Fourcade, J.; Hedge, M.S.; Tarascon, J.-M. Reactivity of Antimony Oxides and MSb2O6 (M = Cu, Ni, Co), Trirutile-type Phases with Metallic Lithium. J. Electrochem. Soc. 2006, 153, A1778. [Google Scholar] [CrossRef]
  18. Matijević, E. Uniform Inorganic Colloid Dispersions. Achievements and Challenges. Langmuir 1994, 10, 8–16. [Google Scholar] [CrossRef]
  19. Sberveglieri, G.; Comini, E.; Faglia, G.; Atashbar, M.Z.; Wlodarski, W. Titanium dioxide thin films prepared for alcohol microsensor applications. Sens. Actuators B Chem. 2000, 66, 139–141. [Google Scholar] [CrossRef]
  20. Mahajan, S.; Jagtap, S. Metal-oxide semiconductors for carbon monoxide (CO) gas sensing: A review. Appl. Mater. Today 2020, 18, 100483. [Google Scholar] [CrossRef]
  21. Ehrenberg, H.; Wltschek, G.; Rodriguez-Carvajal, J.; Vogt, T. Magnetic structures of the tri-rutiles NiTa2O6 and NiSb2O6. J. Magn. Magn. Mater. 1998, 184, 111–115. [Google Scholar] [CrossRef]
  22. Han, J.; Xu, M.; Jia, M.; Liu, T. Evaluation of reduced graphene oxide-supported NiSb2O6 nanocomposites for reversible lithium storage. Ceram. Int. 2016, 42, 14782–14787. [Google Scholar] [CrossRef]
  23. Cullity, B.D.; Stock, S.R. Elements of X-ray Diffraction; Prentice Hall: Upper Saddle River, NJ, USA, 2001. [Google Scholar]
  24. Comini, E.; Baratto, C.; Concina, I.; Faglia, G.; Falasconi, M.; Ferroni, M.; Galstyan, V.; Gobbi, E.; Ponzoni, A.; Vomiero, A.; et al. Metal oxide nanoscience and nanotechnology for chemical sensors. Sens. Actuators B Chem. 2013, 179, 3–20. [Google Scholar] [CrossRef]
  25. Wang, S.-F.; Sun, G.Z.; Fang, L.M.; Lei, L.; Xiang, X.; Zu, X.T. A comparative study of ZnAl2O4 nanoparticles synthesized from different aluminum salts for use as fluorescence materials. Sci. Rep. 2015, 5, 1–12. [Google Scholar] [CrossRef]
  26. Available online: https://materials.springer.com/isp/crystallographic/docs/sd_1815953 (accessed on 22 August 2021).
  27. Matsubara, N.; Damay, F.; Vertruyen, B.; Barrier, N.; Lebedev, O.I.; Boullay, P.; Elkaïm, E.; Manuel, P.; Khalyavin, D.D.; Martin, C. Mn2TeO6: A Distorted Inverse Trirutile Structure. Inorg. Chem. 2017, 56, 9742–9753. [Google Scholar] [CrossRef]
  28. Singh, A.; Singh, A.; Singh, S.; Tandon, P. Nickel antimony oxide (NiSb2O6): A fascinating nanostructured material for gas sensing application. Chem. Phys. Lett. 2016, 646, 41–46. [Google Scholar] [CrossRef]
  29. Guillén-Bonilla, A.; Rodríguez-Betancourtt, V.M.; Guillén-Bonilla, J.T.; Sánchez-Martínez, A.; Gildo-Ortiz, L.; Santoyo-Salazar, J.; Morán-Lázaro, J.P.; Guillén-Bonilla, H.; Blanco-Alonso, O. A novel CO and C3H8 sensor made of CuSb2O6 nanoparticles. Ceram. Int. 2017, 43, 13635–13644. [Google Scholar] [CrossRef]
  30. 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] [Green Version]
  31. Balamurugan, C.; Maheswari, A.R.; Lee, D.W. Structural, optical, and selective ethanol sensing properties of p-type semiconducting CoNb2O6 nanopowder. Sens. Actuators B Chem. 2014, 205, 289–297. [Google Scholar] [CrossRef]
  32. Barreca, D.; Massignan, C.; Daolio, S.; Fabrizio, M.; Piccirillo, C.; Armelao, L.; Tondello, E. Composition and microstructure of cobalt oxide thin films obtained from a novel cobalt(II) precursor by chemical vapor deposition. Chem. Mater. 2001, 13, 588–593. [Google Scholar] [CrossRef]
  33. Yang, H.; Bai, X.; Hao, P.; Tian, J.; Bo, Y.; Wang, X.; Liu, H. A simple gas sensor based on zinc ferrite hollow spheres: Highly sensitivity, excellent selectivity and long-term stability. Sens. Actuators B Chem. 2019, 280, 34–40. [Google Scholar] [CrossRef]
  34. Guillén-Bonilla, A.; Blanco-Alonso, O.; Guillén-Bonilla, J.T.; de la Luz Olvera-Amador, M.; Rodríguez-Betancourtt, V.M.; Sánchez-Martínez, A.; Morán-Lázaro, J.P.; Martínez-García, M.; Guillén-Bonilla, H. Synthesis and characterization of cobalt antimonate nanostructures and their study as potential CO and CO2 sensor at low temperatures. J. Mater. Sci. Mater. Electron. 2018, 29, 15632–15642. [Google Scholar] [CrossRef]
  35. Rodríguez-Betancourtt, V.M.; Bonilla, H.G.; Flores Martínez, M.; Guillén Bonilla, A.; Moran Lazaro, J.P.; Bonilla, J.T.G.; González, M.A.; De La Luz Olvera Amador, M. Gas Sensing Properties of NiSb2O6 Micro-and Nanoparticles in Propane and Carbon Monoxide Atmospheres. J. Nanomater. 2017, 2017, 8792567. [Google Scholar] [CrossRef] [Green Version]
  36. Polshettiwar, V.; Baruwati, B.; Varma, R.S. Self-assembly of metal oxides into three-dimensional nanostructures: Synthesis and application in catalysis. ACS Nano 2009, 3, 728–736. [Google Scholar] [CrossRef] [PubMed]
  37. Tricoli, A.; Righettoni, M.; Teleki, A. Semiconductor gas sensors: Dry synthesis and application. Angew. Chem.-Int. Ed. 2010, 49, 7632–7659. [Google Scholar] [CrossRef] [PubMed]
  38. Polte, J. Fundamental growth principles of colloidal metal nanoparticles—a new perspective. CrystEngComm 2015, 17, 6809–6830. [Google Scholar] [CrossRef] [Green Version]
  39. Mirzaei, A.; Lee, J.H.; Majhi, S.M.; Weber, M.; Bechelany, M.; Kim, H.W.; Kim, S.S. Resistive gas sensors based on metal-oxide nanowires. J. Appl. Phys. 2019, 126. [Google Scholar] [CrossRef] [Green Version]
  40. Oleksenko, L.P.; Maksymovych, N.P. Sensors for CO Based on Semiconductor Nanomaterials Pd/SnO2. Theor. Exp. Chem. 2019, 55, 201–206. [Google Scholar] [CrossRef]
  41. Habib, I.Y.; Tajuddin, A.A.; Noor, H.A.; Lim, C.M.; Mahadi, A.H.; Kumara, N.T.R.N. Enhanced Carbon monoxide-sensing properties of Chromium-doped ZnO nanostructures. Sci. Rep. 2019, 9, 1–12. [Google Scholar] [CrossRef] [Green Version]
  42. Gildo-Ortiz, L.; Rodríguez-Betancourtt, V.M.; Blanco-Alonso, O.; Guillén-Bonilla, A.; Guillén-Bonilla, J.T.; Guillén-Cervantes, A.; Santoyo-Salazar, J.; Guillén-Bonilla, H. A simple route for the preparation of nanostructured GdCoO3 via the solution method, as well as its characterization and its response to certain gases. Results Phys. 2019, 12, 475–483. [Google Scholar] [CrossRef]
  43. Guillén, H.; Rodríguez Betancourtt, V.M.; Guillen, J.T.; Gildo, L.; Guillen, A.; Casallas, Y.; Blanco, O.; Reyes, J. Sensitivity tests of pellets made from manganese antimonate nanoparticles in carbon monoxide and propane atmospheres. Sensors 2018, 18, 2299. [Google Scholar] [CrossRef] [Green Version]
  44. Alrammouz, R.; Podlecki, J.; Abboud, P.; Sorli, B.; Habchi, R. A review on flexible gas sensors: From materials to devices. Sens. Actuators A Phys. 2018, 284, 209–231. [Google Scholar] [CrossRef]
  45. 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]
  46. Morán Lázaro, J.P.; Guillen López, E.S.; López Urias, F.; Muñoz Sandoval, E.; Blanco Alonso, O.; Guillén Bonilla, H.; Guillén Bonilla, A.; Rodríguez Betancourtt, V.M.; Sanchez Tizapa, M.; Olvera Amador, M. de la L. Synthesis of ZnMn₂O₄ Nanoparticles by a Microwave-Assisted Colloidal Method and their Evaluation as a Gas Sensor of Propane and Carbon Monoxide. Sensors 2018, 18, 701. [Google Scholar] [CrossRef] [Green Version]
  47. Arshak, K.; Moore, E.; Lyons, G.M.; Harris, J.; Clifford, S. A Review of Gas Sensors Employed in Electronic Nose Applications; Limerick, Ireland, 2004; Volume 24, ISBN 0260228001. [Google Scholar]
  48. Mcaleer, J.F.; Moseley, P.T.; Norris, J.O.W.; Williams, D.E. Tin Dioxide Gas Sensors. J. Chem. Soc. Faraday Trans. I 1987, 83, 1323–1346. [Google Scholar] [CrossRef]
  49. Nandy, T.; Coutu, R.A.; Ababei, C. Carbon monoxide sensing technologies for next-generation cyber-physical systems. Sensors 2018, 18, 3443. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Carbajal-Franco, G.; Tiburcio-Silver, A.; Domínguez, J.M.; Sánchez-Juárez, A. Thin film tin oxide-based propane gas sensors. Thin Solid Films 2000, 373, 141–144. [Google Scholar] [CrossRef]
  51. Kumar, A.; Kumar, R.; Singh, R.; Prasad, B.; Kumar, D.; Kumar, M. Effect of Surface State Density on Oxygen Chemisorption, Grain Potential and Carrier Concentration for Different Grain Sizes of Nanocrystallite Metal Oxide Semiconductors: A Numerical Modelling Approach. Arab. J. Sci. Eng. 2021, 46, 617–630. [Google Scholar] [CrossRef]
  52. Hjiri, M.; Bahanan, F.; Aida, M.S.; El Mir, L.; Neri, G. High Performance CO Gas Sensor Based on ZnO Nanoparticles. J. Inorg. Organomet. Polym. Mater. 2020, 30, 4063–4071. [Google Scholar] [CrossRef]
  53. Gao, X.; Zhang, T. An overview: Facet-dependent metal oxide semiconductor gas sensors. Sens. Actuators B Chem. 2018, 277, 604–633. [Google Scholar] [CrossRef]
  54. Dey, A. Semiconductor metal oxide gas sensors: A review. Mater. Sci. Eng. B 2018, 229, 206–217. [Google Scholar] [CrossRef]
  55. Ha, N.H.; Thinh, D.D.; Huong, N.T.; Phuong, N.H.; Thach, P.D.; Hong, H.S. Fast response of carbon monoxide gas sensors using a highly porous network of ZnO nanoparticles decorated on 3D reduced graphene oxide. Appl. Surf. Sci. 2018, 434, 1048–1054. [Google Scholar] [CrossRef]
  56. 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. Actuators B Chem. 2011, 160, 580–591. [Google Scholar] [CrossRef]
  57. 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]
  58. 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]
  59. Gildo-Ortiz, L.; Guillén-Bonilla, H.; Reyes-Gómez, J.; Rodríguez-Betancourtt, V.M.; Olvera-Amador, M.D.L.L.; Eguiá-Eguiá, S.I.; Guillén-Bonilla, A.; Santoyo-Salazar, J. Facile Synthesis, Microstructure, and Gas Sensing Properties of NdCoO3 Nanoparticles. J. Nanomater. 2017, 2017, 8174987. [Google Scholar] [CrossRef] [Green Version]
  60. Jayaraman, V.K.; Maldonado Álvarez, A.; Amador, M.D.l.L.O. A simple and cost-effective zinc oxide thin film sensor for propane gas detection. Mater. Lett. 2015, 157, 169–171. [Google Scholar] [CrossRef]
  61. Haocheng, J.; Zeng, W.; Li, Y. Gas sensing mechanisms of metal oxide semiconductors: A focus review. Nanoscale 2019, 11, 22664–22684. [Google Scholar] [CrossRef]
  62. Wang, C.; Yin, L.; Zhang, L.; Xiang, D.; Gao, R. Metal oxide gas sensors: Sensitivity and influence factors. Sensors 2010, 10, 2088–2106. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Du, H.; Xie, G.; Su, Y.; Tai, H.; Du, X.; Yu, H.; Zhang, Q. A new model and its application for the dynamic response of RGO resistive gas sensor. Sensors 2019, 19, 889. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Thanh-Tien, N.; Phuong-Than, L.V.; Yen-Mi, T. An ab initio study of small gas molecule adsorption on the edge of N-doped sawtooth penta-graphene nanoribbons. Pap. Phys. 2021, 13, 130003. [Google Scholar] [CrossRef]
  65. Roohi, H.; Askari-Ardehjani, N. Exploring the adsorption of CO toxic gas on pristine and M-doped (M = Ti, Cr, Fe, Ni and Zn) GaN nanosheets. New J. Chem. 2019, 43, 15280–15292. [Google Scholar] [CrossRef]
  66. Nykanen, L.; Honkala, K. Density functional theory study on propane and propene adsorption on Pt(111) and PtSn alloy surfaces. J. Phys. Chem. C 2011, 115, 9578–9586. [Google Scholar] [CrossRef]
  67. Guillén-Bonilla, A.; Rodríguez-Betancourtt, V.M.; Guillén-Bonilla, H.; Gildo-Ortiz, L.; Blanco-Alonso, O.; Franco-Rodríguez, N.E.; Reyes-Gómez, J.; Casillas-Zamora, A.; Guillen-Bonilla, J.T. A new CO2 detection system based on the trirutile-type CoSb2O6 oxide. J. Mater. Sci. Mater. Electron. 2018, 29, 15741–15753. [Google Scholar] [CrossRef]
  68. Bochenkov, V.E.; Sergeev, G.B. Preparation and chemiresistive properties of nanostructured materials. Adv. Colloid Interface Sci. 2005, 116, 245–254. [Google Scholar] [CrossRef] [PubMed]
  69. Yamazoe, N. Toward innovations of gas sensor technology. Sens. Actuators B Chem. 2005, 108, 2–14. [Google Scholar] [CrossRef]
  70. Neri, G. First fifty years of chemoresistive gas sensors. Chemosensors 2015, 3, 1–20. [Google Scholar] [CrossRef]
Figure 1. Synthesis process of NiSb2O6 powders in air.
Figure 1. Synthesis process of NiSb2O6 powders in air.
Applsci 11 09536 g001
Figure 2. Schematic of the system employed in the static and dynamic tests in CO.
Figure 2. Schematic of the system employed in the static and dynamic tests in CO.
Applsci 11 09536 g002
Figure 3. X-ray diffraction patterns of NiSb2O6 calcined at 700, 800, and 900 °C in air.
Figure 3. X-ray diffraction patterns of NiSb2O6 calcined at 700, 800, and 900 °C in air.
Applsci 11 09536 g003
Figure 4. (a) Crystallite size as a function of the calcination temperature, (b) projection of an octahedral hole, (c) atomic ordering of the trirutile-type NiSb2O6’s structure.
Figure 4. (a) Crystallite size as a function of the calcination temperature, (b) projection of an octahedral hole, (c) atomic ordering of the trirutile-type NiSb2O6’s structure.
Applsci 11 09536 g004
Figure 5. (a) UV–Vis spectra of NiSb2O6, (b) determination of the forbidden band at the different calcination temperatures.
Figure 5. (a) UV–Vis spectra of NiSb2O6, (b) determination of the forbidden band at the different calcination temperatures.
Applsci 11 09536 g005
Figure 6. SEM images of NiSb2O6 powders calcined at 900 °C at the following magnifications: (a) 7.06 kx, (b) 13.7 kx, (c) 33.0 kx, (d) 39.9 kx, (e) 43.3 kx.
Figure 6. SEM images of NiSb2O6 powders calcined at 900 °C at the following magnifications: (a) 7.06 kx, (b) 13.7 kx, (c) 33.0 kx, (d) 39.9 kx, (e) 43.3 kx.
Applsci 11 09536 g006
Figure 7. Morphologies’ distribution of the NiSb2O6 powders calcined at 900 °C: (a) particle size, (b) size of the polyhedron-type morphologies.
Figure 7. Morphologies’ distribution of the NiSb2O6 powders calcined at 900 °C: (a) particle size, (b) size of the polyhedron-type morphologies.
Applsci 11 09536 g007
Figure 8. Response of NiSb2O6 nanoparticles obtained at 900 °C in CO atmospheres as a function of (a) CO concentration, (b) operating temperature.
Figure 8. Response of NiSb2O6 nanoparticles obtained at 900 °C in CO atmospheres as a function of (a) CO concentration, (b) operating temperature.
Applsci 11 09536 g008
Figure 9. Sensing tests at 200 °C and different CO and air flows as a function of time: (a) variation of electrical resistance, (b) variation of sensitivity percentage.
Figure 9. Sensing tests at 200 °C and different CO and air flows as a function of time: (a) variation of electrical resistance, (b) variation of sensitivity percentage.
Applsci 11 09536 g009
Figure 10. (a,b) Dynamic response at 0.1 and 1 kHz of NiSb2O6 in 600 ppm of C3H8 at 360 °C, (ce) resistance, reactance, and capacitance as a function of time at 1 kHz in 600 ppm of C3H8 at 360 °C.
Figure 10. (a,b) Dynamic response at 0.1 and 1 kHz of NiSb2O6 in 600 ppm of C3H8 at 360 °C, (ce) resistance, reactance, and capacitance as a function of time at 1 kHz in 600 ppm of C3H8 at 360 °C.
Applsci 11 09536 g010
Figure 11. (a) Bode diagram of NiSb2O6 pellets’ behavior, (b) electronic diagram of the C3H8-detector, (c) detector’s principle of operation.
Figure 11. (a) Bode diagram of NiSb2O6 pellets’ behavior, (b) electronic diagram of the C3H8-detector, (c) detector’s principle of operation.
Applsci 11 09536 g011
Table 1. Response of NiSb2O6 pellets at different concentrations and operating temperatures.
Table 1. Response of NiSb2O6 pellets at different concentrations and operating temperatures.
Temperature (°C)CO
(ppm)
Response
(R)
Temperature (°C)CO
(ppm)
Response
(R)
1000.87 10035.52
2002000.6730020055.50
3003.14 30072.67
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ramírez-Ortega, J.A.; Guillén-Bonilla, J.T.; Guillén-Bonilla, A.; Rodríguez-Betancourtt, V.M.; Gildo-Ortiz, L.; Blanco-Alonso, O.; Soto-García, V.M.; Jiménez-Rodríguez, M.; Guillén-Bonilla, H. Preparation of Powders Containing Sb, Ni, and O for the Design of a Novel CO and C3H8 Sensor. Appl. Sci. 2021, 11, 9536. https://doi.org/10.3390/app11209536

AMA Style

Ramírez-Ortega JA, Guillén-Bonilla JT, Guillén-Bonilla A, Rodríguez-Betancourtt VM, Gildo-Ortiz L, Blanco-Alonso O, Soto-García VM, Jiménez-Rodríguez M, Guillén-Bonilla H. Preparation of Powders Containing Sb, Ni, and O for the Design of a Novel CO and C3H8 Sensor. Applied Sciences. 2021; 11(20):9536. https://doi.org/10.3390/app11209536

Chicago/Turabian Style

Ramírez-Ortega, Jorge Alberto, José Trinidad Guillén-Bonilla, Alex Guillén-Bonilla, Verónica María Rodríguez-Betancourtt, Lorenzo Gildo-Ortiz, Oscar Blanco-Alonso, Víctor Manuel Soto-García, Maricela Jiménez-Rodríguez, and Héctor Guillén-Bonilla. 2021. "Preparation of Powders Containing Sb, Ni, and O for the Design of a Novel CO and C3H8 Sensor" Applied Sciences 11, no. 20: 9536. https://doi.org/10.3390/app11209536

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