Next Article in Journal
Batch-Injection Amperometric Determination of Glucose Using a NiFe2O4/Carbon Nanotube Composite Enzymeless Sensor
Previous Article in Journal
CoNiTe2 Nanomaterials as an Efficient Non-Enzymatic Electrochemical Sensing Platform for Detecting Dopamine
Previous Article in Special Issue
Enhancement of H2 Gas Sensing Using Pd Decoration on ZnO Nanoparticles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

SO2 Detection over a Wide Range of Concentrations: An Exploration on MOX-Based Gas Sensors

by
Arianna Rossi
1,*,
Elena Spagnoli
1,
Alan Visonà
1,
Danial Ahmed
1,
Marco Marzocchi
2,
Vincenzo Guidi
1 and
Barbara Fabbri
1,*
1
Department of Physics and Earth Sciences, University of Ferrara, Via Saragat 1/C, 44122 Ferrara, Italy
2
Sacmi Imola S.C., Olfactory Systems, Via Selice Provinciale, 17/a, 40026 Imola, Italy
*
Authors to whom correspondence should be addressed.
Chemosensors 2024, 12(6), 111; https://doi.org/10.3390/chemosensors12060111
Submission received: 10 May 2024 / Revised: 7 June 2024 / Accepted: 10 June 2024 / Published: 14 June 2024
(This article belongs to the Special Issue Gas Sensors and Electronic Noses for the Real Condition Sensing)

Abstract

:
Noxious gases such as sulfur-containing compounds can inflict several different adverse effects on human health even when present at extremely low concentrations. The accurate detection of these gases at sub-parts per million levels is imperative, particularly in fields where maintaining optimal air quality is crucial. In this study, we harnessed the capabilities of nanostructured metal-oxide semiconducting materials to detect sulfur dioxide, since they have been extensively explored starting from the last decades for their effectiveness in monitoring toxic gases. We systematically characterized the sensing performance of seven chemoresistive devices. As a result, the SnO2:Au sensor demonstrated to be the most promising candidate for sulfur dioxide detection, owing to its highly sensitivity (0.5–10 ppm), humidity-independent behavior (30 RH% onwards), and selectivity vs. different gases at an operating temperature of 400 °C. This comprehensive investigation facilitates a detailed performance comparison to other devices explored for the SO2 sensing, supporting advancements in gas detection technology for enhanced workplace and environmental safety.

1. Introduction

Sulfur oxides (SOx) are one of the major contributors to atmospheric pollution, including acid rain and smog formation [1]. Indeed, the oxidation of SO2 leads to the formation of sulfuric acid and fine particulate matter (PM2.5, particulate matter less than or equal to 2.5 μm in diameter). This can have ecological impacts on soil, forests, and freshwater. Moreover, this category of pollutants is strongly harmful to human beings because inhaled SO2 can easily be hydrated in the respiratory tract, producing sulfurous acid that subsequently dissociates to form its derivatives, bisulfite and sulfite anions [2], which can cause damage and increase chronic obstructive pulmonary diseases, e.g., pneumonia and acute bronchitis, and daily mortality [3]. The impact on human health motivated the low Threshold Limit Values, i.e., the Short-Term Exposure Limit (TLV-STEL) and the Permission Exposure Limit Time-Weighted Average (PEL-TWA) for SO2 are 0.25 parts per million (ppm) and 5 ppm, respectively [4,5]. Therefore, many countries have established or are enacting stricter regulations to reduce the SO2 emissions into the atmosphere [1].
In nature, the release of SOx is caused by the combustion of carbonaceous fuels, which occurs in forest fires and volcanoes [6]. In the last case, SO2 fluxes can be substantial in terms of atmospheric source strengths, for instance, the mean SO2 flux from Mt. Etna in Sicily is equivalent to the total of anthropogenic SO2 emissions from France, making it arguably the world’s largest sustained point source emitter of SO2. In addition to these inevitable natural contributions, there are also many anthropogenic sources of SOx releases such as the combustion of fossil fuels in power plants [7], for which the permissible concentration levels are 0.03 ppm (1 h) and 0.14 ppm (24 h) [8] in the case of electricity generation and industry boilers, respectively. Further contributions to SOx emission came from oil refineries, domestic boilers, and motor vehicles, e.g., the European permissible limit of SO2 produced from the automobiles exhaust ranges between 20 and 120 ppm/km [9]. Moreover, the extensive use of food additives leads to residual SO2, generally in the range of 50–250 ppm [10], while the use of pesticides such as ammonium sulfate and potassium sulfate leads to the inevitable decomposition of SO2 gas. Exceeding 2 ppm of SO2 can have detrimental effects on crops, as it can potentially destroy chloroplasts essential for photosynthesis [11].
To minimize the impact on the environment, it has become a primary goal to reduce anthropogenic SOx, particularly by desulfurization of flue gas during combustion, as well as to monitor SOx concentrations in situ at the sources and in the ambient atmosphere. Moreover, air quality monitoring of SO2 is critical for safety in closed spaces and for leak identification.
Hence, an effective monitoring system is crucial to maintaining SO2 levels within safe thresholds. To analyze and quantify SO2 and other sulfur compounds, the conventionally used laboratory technique is gas chromatography coupled with mass spectrometry (GC-MS). GC-MS is acknowledged for its precision and reliability in identifying gaseous compounds, establishing itself as the industry standard. Despite its accuracy, GC-MS presents practical challenges such as the need for bulky and expensive equipment, specialized personnel, and complex pre-treatment processes, resulting in being unsuitable for continuous real-time detections. This underscores the impracticality of such techniques for widespread use in on-site gas monitoring. Therefore, there is a clear need for alternative methods that offer cost-effectiveness, swift processing, rapid response, and reduced reliance on solvents and samples for routine detections.
To date, gas sensors have emerged as both a cost-effective alternative and a complementary system to laboratory-based analyses, providing a rapid and promising solution for SO2 detection. Notably, electronic nose (e-nose) systems have gained attention for odor assessments, leveraging their versatility and resemblance to the human olfactory system. The implementation of e-noses involves training procedures, with qualified samples to build a database of reference, as well as sensor array optimization, which requires the selection of appropriate devices among the different classes of gas sensors. Recent developments involved the use of electrochemical and non-dispersive infrared (NDIR) sensors for SO2 detection. However, some limitations by NDIR devices, e.g., short device lifetime, low selectivity, and spectral interference, have prevented their widespread usage.
Metal-oxide semiconductor (MOX)-based chemoresistive gas sensors are a compelling alternative, showcasing not only high sensitivity but also cost-effectiveness, scalable manufacturing, durability, and seamless integration into Internet of Things (IoT) systems [12]. Previous experimental and theoretical works proved that MOX-based nanostructured materials, such as SnO2, WO3, and ZnO are potential candidates to detect SO2 [7,13,14,15,16,17,18]. Nonetheless, the investigation of MOX for SO2 detection has been hampered by evidence of irreversible interactions between this analyte and the active groups over the sensing layers [19], resulting in sensor instability and a short lifetime. For instance, Berger et al. [20] demonstrated irreversible sulfate formation on the surface of SnO2, which was identified as the cause of the irreversibility of the device response after initial SO2 detection.
Hence, to enhance their sensitivity and selectivity, various approaches, including the do** and/or decoration effect and heterojunction formation, have been investigated for tuning the material properties. In these perspectives, SnO2-based sensors, containing small amounts of metals such as Ni and Cu, have already been characterized and tested to achieve selective and sensitive detection of SO2 gas [21,22].
Consequently, to further investigate the potentialities of chemoresistive gas sensors in monitoring SO2, this study assesses the sensing capability of seven distinct MOX materials. Three of these are pure MOXs, namely, SnO2, WO3 and ZnO, to identify the most promising among them within our measurement system. The best candidate, in terms of sensitivity and selectivity, has been functionalized with noble metals, such as Au, Pt, Pd, and Ag.

2. Materials and Methods

2.1. Chemoresistive Gas Sensors

ZnO, SnO2, WO3, SnO2:Au, SnO2:Pt, SnO2:Pd, and SnO2:Ag were synthesized in powder form by the sol-gel method [23,24,25]. The powders were mixed with organic excipients, namely, α-terpineol, ethyl cellulose, and silica, to form homogeneous pastes that facilitate the screen-printing technique. The resulting pastes were sonicated for 2 h and then screen-printed onto alumina substrates (defined area of 2.54 × 2.54 mm2) [26,27]. The semiconducting MOX thick films (around 20–30 μm of thickness) were sintered in air for 2 h at 650 °C to remove the organic precursors and improve mechanical stability and film adhesion. The substrates were equipped with interdigitated gold contacts on the front side, to provide the bias voltage for measuring the conductance of the film, and a platinum heater was integrated on the back side to thermally activate the sensing material at its optimal working temperature. Finally, these substrates were bonded to a standard TO-39 support through thermo-compression of 60 μm of diameter gold wires (Figure S1, Supporting Information). The integration of electrodes and heaters in the alumina substrate allows for precise control and measurement of the sensing material’s response.

2.2. Lab-Test Setup

The MOX-based sensors were electrically characterized in a sealed test chamber of volume ≈ 622 cm3 in which temperature and relative humidity (RH%) were measured by a commercially available Honeywell (Charlotte, NC, USA) HIH-4000 sensor (accuracy ± 3.5 RH%). Firstly, the devices were exposed to synthetic dry air (20% O2 and 80% N2) for almost one day in order to obtain the baseline stabilization. Secondly, the mixture of gases from certified cylinders were injected inside the chamber through mass-flow controllers, achieving a total flow of 500 standard cubic centimeters per minute (sccm). The filling time of the test chamber was about 1 min and 15 s, as it depends on the size, the geometry of the chamber, and the velocity of the gas flow. The humidity conditions were obtained by fluxing a fraction of the total synthetic air into a gas bubbler filled with deionized water. Power suppliers from Aim TTi (Glebe Rd, Huntingdon, Cambridgeshire, UK) were used to furnish the requisite electric current to the sensor heater, while a multimeter, the K2000 from Keithley (Cleveland, OH, USA), was utilized to measure the electrical conductance of the sensing film. A constant bias of 5 V was maintained between the two interdigitated gold electrodes. The circuitry employed for data acquisition was structured around an operational amplifier (OA), which serves as a pivotal component in processing the signals. Thanks to the configuration in which the input voltage (Vin) and output voltage (Vout) values corresponded to the ends of the sensor resistor (Rs) and the applied load resistor (Rf), respectively, the gain was determined by the relationship V o u t V i n   = R f R S . This fundamental expression elucidates the relationship between the Vin and Vout, facilitating accurate interpretation of the sensor’s response. According to A. Rossi et al. [26], the expression for the sensor conductance GS is given as follows:
G S   = 1 R S = V o u t R f   V i n

2.3. Sensing Measurements

The sensors performances were proven by making a comprehensive electrical characterization.
First, the best operating temperature of each MOX film was investigated considering their response to 1 ppm of SO2 at increased working temperatures, ranging from 200 to 450 °C.
Then, in order to study the sensitivity, the films were exposed to several concentrations of SO2 ranging from 0.5 to 10 ppm under dry conditions. Specifically, to verify possible surface poisoning effects, once achieved the plateau of the signal after the gas injection, the sensors were again exposed to a fully dry air flux to assess the complete recovery of the baseline The sensor response was defined as the following equations for reducing and oxidizing gases, respectively:
R = G g a s   G a i r G a i r   for   reducing   gas
R = G a i r   G g a s G g a s   for   oxidizing   gas
where Ggas and Gair are the conductance values in gas and in air, respectively. The error on the response value was calculated applying error propagation, considering the load resistance tolerance and the instrument uncertainty on the output voltage acquisition. The result was an error of 5% [28].
Due to instrumental limitations, the sensors’ response at the TLV-STEL concentration, 0.25 ppm, was not experimentally evaluated. Otherwise, the theoretical Limit of the Detection (LOD) of the best device was determined as 3(RMS noise/Rair), where RMS noise is the root mean square deviation and Rair is the resistance in dry air [29].
Humidity is a very common interference due to its ability to interact with the film, affecting its conductance and generating competitive −OH groups that limit the adsorption sites on the reactive surface. To explore its effect on the sensor performance, 5 ppm of SO2 was injected into the gas chamber in presence of different percentages of humidity (2–56 RH%).
The stability and repeatability of the best SO2 devices at its operating temperature was assessed under 6 ppm of the target gas in dry and wet (30 RH%) conditions. The response (τres) and the recovery (τrec) times of the sensors were calculated as the times needed to attain 90% of the steady-state response and required to switch back to 90% of the baseline value, respectively.
To verify the influence of possible interfering compounds, they were also exposed to different volatile analytes, such as benzene, ethanol, nitrogen dioxide (NO2), dimethyl disulfide (DMDS), carbon monoxide (CO), and carbon dioxide (CO2). The gases and their concentrations were chosen in order to test the surface reactivity of the semiconductors vs. different chemical functional groups. In particular, the concentrations of the potential inteferents were chosen based on TLV-TWA and the American Conference of Governmental Industrial Hygienists (ACGIH) TLV values [30,31,32,33,34,35].
Moreover, to evaluate the ability of the sensors to detect SO2 in presence of different analytes, we also performed cross-selectivity measurements. Therefore, the sensors were exposed to 5 ppm of SO2 in an atmosphere of 1 ppm of ethanol in dry and wet (30 RH%) conditions.

3. Results

3.1. Operational Temperature

The sensors’ optimal working temperature was determined by measuring conductance variations before and after introducing 1 ppm of SO2 in dry conditions across a range of working temperatures from 200–450 °C (Figure 1). The response of ZnO, SnO2:Pd, SnO2:Au, SnO2, SnO2:Pt, and SnO2:Ag sensors increased with the temperature rise, reaching the highest value at 400 °C. High operating temperatures are commonly used to lower response and recovery times and to optimize the MOX-based sensors catalytic activity towards SO2, facilitating surface redox reactions. Nevertheless, beyond the optimal temperature, gas desorption rates rise significantly, leading to further lower sensor response [24].
In contrast to the other sensors, WO3 exhibited two peak responses at 200 °C and 350 °C, suggesting distinct optimal conditions for detecting the analyte, likely due to different active sites on the surface for SO2 chemisorption, namely, O2 at 200 °C and O at 350 °C, as explained in our previous work [26] and also deepened in the Discussion section of the present study. However, since the response value at 200 °C and 350 °C is almost the same, the lowest working temperature (200 °C) was selected due to alignment with requirements for low-power consumption in portable and IoT applications.
The optimal temperature for each sensor is summarized in Table 1 and was selected for subsequent comprehensive electrical characterization.
Notably, Figure 1 also shows that the SnO2-based sensor had greater response values than the other pure MOXs, namely, ZnO and WO3. That had been the underlying motivation for decorating SnO2 with noble metals rather than ZnO or WO3.

3.2. Sensitivity

To investigate the sensitivity, we exposed the sensors towards several concentrations of the analyte, in particular, SnO2:Au, SnO2:Pd and ZnO from 0.5 to 10 ppm, instead of SnO2, SnO2:Pt and SnO2:Ag from 1 to 10 ppm. Only WO3 was tested from 0.5 to 7 ppm since the response tended to remain constant and low. Figure 2 highlights that SnO2:Au, SnO2:Pd and SnO2 sensors at each tested concentration exhibited higher responses than the others and they did not saturate at the relative highest concentration tested. On the other hand, WO3, ZnO, SnO2:Ag, and SnO2:Pt sensors showed a constant response for concentrations above 2 ppm, which can be attributed to the saturation of active sites on the materials and sulfur poisoning of the surface [36]. Therefore, SnO2:Au, SnO2:Pd and SnO2 sensors exhibited a remarkable capability to detect SO2 across a wide range of concentrations, which make them appropriate for diverse industrial, environmental, and safety monitoring applications.
The calibration curves of the sensors were fitted with asymptotic function ( y = a b c x ), except for the SnO2:Ag device, for which it was not possible to identify a suitable curve fit. All the metrics parameters of the function are listed in Table S1 (Supporting Information).
In particular, it turned out that the SnO2:Au sensor showed better responses than the other devices for SO2 detection with a theoretical LOD of ∼ 0.48 ppm. In Figure S2 (Supporting Information), we reported the SnO2:Au sensor response at three SO2 concentrations close to its LOD. One can observe that below the LOD, the sensor signal (at 0.2 and 0.3 ppm) was noisy and not stable.

3.3. Humidity Effect

The impact of humidity on the sensing performance was assessed by fluxing 5 ppm of SO2 while exposing the devices to escalating levels of water vapor (2–56 RH%). As depicted in Figure 3, the responses exhibited a significant drop at 17 RH% due to the competitive reaction for active sites between water molecules and SO2 [27]. However, for SnO2-based sensors and ZnO, the response to SO2 was clearly discernible, with SnO2:Au exhibiting the strongest signal. Only the WO3 was not able to detect SO2 in humid conditions for values higher than 17 RH%. Starting from 45 RH%, all sensors maintained an almost constant response. This behavior is explainable with the Grotthuss mechanism, in which the formation of continuous water layers over the sensing film hinders the reaction between the analytes and the functional material active sites [37].
To the best of our knowledge, the only work that deeply discussed the humidity effect in SO2 detection was developed by Gaiardo et al., who proposed a novel approach using nanostructured silicon carbide (SiC) as a sensing film [28]. It resulted in a highly selective SO2 sensor even if it operated at a high temperature (600 °C). In this work, they tested five different relative humidities vs. 10 ppm of SO2. Starting from 2 to 60 RH%, the response remained rather stable around 0.25, reaching a peak of 0.27 at 13 RH%. Moreover, in a consecutive study, M. Della Ciana et al. supported the previous work by Gaiardo et al. applying operando Diffuse Reflectance Infrared Fourier Transform (DRIFT) spectroscopy. With this advanced technique, they demonstrated that the enhancement of SO2 response under wet conditions was attributed to the formation of SOx species, which increased the number of charge carriers. However, despite these advancements, especially in SO2-sensing mechanisms in wet conditions, SiC sensors still exhibited a low response (0.20 at 8 RH%) even to higher concentrations of SO2, such as 25 ppm [37]. On the contrary, the response at 8 RH% of SnO2:Au sensor to a SO2 concentration five times less (5 ppm) is equal to 3.7.

3.4. Repeatability

The stability of the dynamic response vs. 6 ppm of SO2 during four consecutive cycles was investigated in dry and wet conditions for the best candidate, SnO2:Au (Figure 4). The film exhibited a clear reversible interaction and good repeatability for detection over time. A doubling increase in the conductance baseline of the sensitive film was observed when exposed to a humid atmosphere due to water’s homolytic dissociation. This chemical process involved water interacting with metal sites and lattice oxygen, leading to the formation of terminal and anchored hydroxyl groups (−OH) on the film’s surface. These −OH groups acted as surface donors, releasing electrons into the conduction band. This alteration in the surface chemistry affected the film’s catalytic properties regarding SO2 detection by replacing the oxygen ion active sites, typically present in dry conditions, with the newly formed hydroxyl groups.
However, the variation of the conductance values over the four test cycles under SO2 exposure was similar in dry and wet conditions. Instead, the values of the conductance baseline mainly increased during the test in dry air, while they recovered almost the initial levels during the measurement in humid air. This confirmed that SnO2:Au sensor can be reliable in real-like conditions limiting the effect of humidity on the signal drift.
The response and recovery times of SnO2:Au sensors are 3 and 17 min, respectively, in dry conditions, whereas in wet conditions, the response and recovery times are 5 min and 19 min, respectively. These values are in line with those exhibited by common MOX-based gas sensors [24,26].
In addition, Figure 4 shows a similar response shape of the sensor in dry and humid conditions, i.e., a rapid increase when the target gas was introduced. This sharp peak can be ascribed to the dynamic equilibrium between gas diffusion into deeper layers of the sensing material and the kinetic processes of gas adsorption/desorption at its surface [38]. At the initial stage, SO2 diffuses along the sensing film, starting to influence the dynamic equilibrium of redox reactions occurring at the surface of SnO2:Au, decreasing, for example, the number of ionized oxygens, resulting in an increase in conductance. Nevertheless, at equilibrium, several factors may influence the conductance in presence of SO2, slightly lowering its initial value. For instance, the competitive diffusion of new analyte and non-reactive reaction products inside and outside the sensing film should be considered, as well as the kinetic regeneration of active sites.
In order to investigate the long-term stability of the SnO2:Au sensor, we verify its response value by comparing three different measurements with 3 ppm of SO2 over a period of 5 months (Figure S3, Supporting Information).
A deepened investigation on SO2 chemoresistive gas sensors was conducted in order to compare the performance of SnO2:Au studied in this work with devices already discussed in the literature; see Table 2. Even if it operates at a high temperature, the SnO2:Au sensor can be obtained with an easier and cheaper synthesis process, such as sol-gel synthesis, enabling a potential large-scale production than the semiconducting composites explored by other authors mentioned. Moreover, the complete inorganic nature of our sensing material guarantees a strong stability with respect to the organic-based semiconductors. The SnO2:Au sensor can detect SO2 over a wide range of concentrations (0.5–10 ppm) and, according to the PEL-TWA (5 ppm), it showed a response at least three times higher than the other devices. Under wet conditions, among the sensing material listed in Table 2, only the Ni/SnO2 sensor exhibited a comparable response to SnO2:Au ones, even if the latter demonstrated a negligible influence by humidity over 30 RH%, which makes it suitable for real applications.
The SnO2:Au sensor showed outstanding sensitivity even at a very low concentration. Indeed, the theoretical limit of detection calculated is equal to 0.48 ppm, a value remarkably comparable to that obtained in the Ag NWs@SnO2/CuO sensor, which reported a LOD of 0.25 ppm.
Concerning the response and recovery times, it was not possible to compare the performance since these values are strongly correlated with geometry and arrangement of the experimental setup, e.g., test chamber volume and gas flow rate.
Therefore, the analysis carried out in this work revealed superior performance of the Au-decorated SnO2 for the sulfur dioxide detection with respect to the current MOX-based sensors investigated from the literature.

3.5. Selectivity

Moreover, in the practical implementation, one should consider the co-presence of other gases whose physicochemical properties may affect the sensor performance. Therefore, the selectivity of the devices was investigated by exposing the sensors to various interferents, such as benzene (0.5 ppm), ethanol (5 ppm), NO2 (3 ppm), DMDS (0.25 ppm), CO2 (5000 ppm) and CO (25 ppm) under dry air. The dynamic signal of output voltage of the sensors for all the measurements are shown in Figures S4–S10 (Supporting Information), together with the load resistances in Table S2. The bar graph in Figure 5 illustrates the response values determined by Equations (2) and (3), to better evaluate the sensor’s selectivity.
As expected, almost all sensors exhibited a high response to ethanol [26,48,49,50], in particular, SnO2:Pd, which therefore cannot be defined as selective for SO2 detection. Among the decorated sensor-based on SnO2, the SnO2:Au is the one that exhibited the lowest response value to ethanol. In contrast, all the sensors exhibited negligible responses to CO, CO2 and benzene. Only SnO2-based sensors showed superior response to sulfur-containing gases such as DMDS and SO2, especially SnO2:Au, which proved to be particularly selective vs. SO2. Then, WO3 and ZnO, two of the most common pristine nanostructured MOX materials, were considered unsuitable for SO2 detection due to their low sensitivity (Figure 2) and stronger selectivity for interfering gases such as ethanol and NO2 than for the analyte of interest (Figure 5).

3.6. Cross-Selectivity

The cross-selectivity of the sensors towards 5 ppm of SO2 in the presence of 1 ppm (baseline) of ethanol was performed in dry and wet conditions (30 RH%), as shown in Figure 6. The presence of ethanol as an interferent during the exposure to SO2 had a notable impact on the response of SnO2:Au sensors. Comparing Figure 6 to Figure 2, one can observe that the response to 5 ppm of SO2 decreased from 10.4 to ~2.5 in dry conditions and ~0.5 in wet conditions. This effect can be attributed to competitive reactions between the reactive gases. Indeed, both ethanol and H2O can react with the material active sites, making them ineffective for detecting SO2. Furthermore, the products of such reactions, for example chemisorbed alkyl chains, might accumulate over the surface, hysterically hindering nearby active sites. However, SnO2:Au demonstrated the effective ability to discriminate against SO2 among ethanol interfering gas.

4. Discussion

Gas-sensing mechanisms of metal oxides, such as SnO2, ZnO, and WO3 presented in this study, involve complex surface reactions, which are influenced by various intrinsic and extrinsic factors. The intrinsic ones are inherent factors such as (i) activation of chemisorbed active sites from oxidative and reductive reactions, (ii) the creation of acidic or basic centers, and (iii) hydrate–hydroxyl layers, that, synergically combined, play a crucial role [51]. Indeed, (iv) the formation of extrinsic active centers on the surface are due to the modifications by catalytic clusters of noble metals, e.g., Au, Pd, Pt, and Ag.

4.1. Intrinsic Properties

4.1.1. Oxygen Chemisorption

The most common sensing mechanism explaining the response of MOX-based sensors to reducing gases involves oxygen chemisorption and subsequent redox reactions of the analytes with the oxygen-based surface ions. In principle, the oxygen molecules, which are chemisorbed on the semiconductors, ionize into O2, O, and O2− anion species by accepting the electrons from MOX materials at different temperatures and decreasing the overall conductance of the sensing film. The oxygen adsorption has been investigated by several works in literature [52], as it follows:
( O 2 ) g a s   ( O 2 ) a d s o r b e d  
( O 2 ) a d s o r b e d   + e O 2
O 2 + e 2 O  
O + e   O 2
Oxygen anions O2, O, and O2− are stable below 100 °C, 100–300 °C, and above 300 °C, respectively [52]. Based on this consideration, since in the current work, the operating temperature for most of the sensors was 400 °C, it is likely that they exhibited the formation of O2− ions (Figure 7a), except for WO3, which operates at a low temperature of 200 °C, where the formation of chemisorbed O2 and O ions was more probable. Differently from the other sensors, WO3 displayed two maxima responses at 200 °C and 350 °C, meaning that there are two distinct conditions for optimal detection of SO2, probably characterized by different active sites, namely, O2 and O ions in the first case and O and O2− in the second case. In the temperature range between these two optimal values (200 °C and 350 °C), an unfavorable competition between different reaction mechanisms of the analyte absorption and desorption would lead to a decrease in WO3-sensing capabilities [26].
When the sensors were exposed to SO2 gas molecules in dry conditions, they reacted with the chemisorbed oxygen species on the surface and released the electrons to the conduction band of the MOX material.
In particular, in the case of the sensor that exhibited the best performance in the present work, i.e., SnO2:Au, the overall surface reaction mechanism was given by the catalytic oxidation of SO2 to SO3 at high temperature (400 °C) [28] (see Figure 7b):
S O 2   + O 2   S O 3   + 2 e  
From the above reaction, it is clear that when the target gas reacts with the oxygen species adsorbed on the SnO2:Au surface, the number of free electrons increases, resulting in an increase in conductivity of the material.

4.1.2. Acid and Basic Centers

Beyond the ionosorption and interaction of oxygen species, surface acidity and basicity play a pivotal role in gas-sensing mechanisms. Indeed, functional material active sites can be basic in nature, such as surface oxygen anions and hydroxyl species, or acidic, like coordinately unsaturated cations and −OH Brønsted acid states [53]. The concentration of different active sites influences the adsorption and reaction of target gases on the sensor’s surface, thereby affecting its sensitivity and selectivity, as deeply discussed by Marikutsa et al. in [53]. The concentration of Brønsted and Lewis acid sites followed the order ZnO < SnO2 < WO3, consistent with the increase in metal–oxygen bond energy, electronegativity, and charge/radius ratio of cations. This sequence also reflects the decreasing optical basicity of oxides [53]. The basic nature of ZnO resulted in selectivity for VOCs like ethanol, but poor detection capabilities for sulfur-containing compounds. On the other hand, WO3, the most acidic in nature, displayed a high response for oxidizing gas, such as NO2, and very low signals for the other reducing gases. Between the three MOXs, SnO2 was the most suitable for the development of chemoresistive SO2 sensors. Berger et al. demonstrated that an increase in surface acidity for SnO2-based material (probably related to sulfate formation) after SO2 exposure may explain the increase in the sensor’s sensitivity [20]. Furthermore, enhanced sensitivity of SnO2-based material was ascribed to the presence of noble metals, such as Au and Pd, which elevate the inherent acidity of pristine SnO2. Therefore, the intrinsic acid centers of SnO2, in combination with the extrinsic active centers, i.e., noble metals, effectively affect the sensing performance towards SO2.

4.1.3. Hydrated Layers

Concerning the gas-sensing mechanism in wet conditions, the presence of H2O molecules affects the reaction scenario (Figure 7c). Indeed, H2O molecules (Lewis base) can be adsorbed as are on the surface or interact with cations (Lewis acid) by a strong binding resulting in water molecules dissociation and then producing the −OH groups. In this second case, strong Lewis acidic cations, i.e., W6+, Sn4+, Zn2+, can bind more than one −OH group, then acquire an acidic proton and behave as Brønsted acid site. This leads to an increase in the semiconductors conductance due to the redox reaction, which provides electrons to the sensing materials.
In our case, before the interaction with the sensing films, water molecules can react with SO2 in the atmosphere by producing H2SO3 and further by-products (see Figure 7d). Subsequently, SO2 and the other gaseous compounds, resulting from the interaction between SO2 and H2O, can interact with oxygen and hydroxyl ions adsorbed on the surface of the materials. The competitive reaction between −OH groups and pre-adsorbed oxygen species contribute to reducing the sensitivity of the materials and, as a result, negatively affect the performance of the sensors with respect to the dry ambient.
Furthermore, the consecutive creation of continuous layers of water molecules would hinder the functional material’s active sites for reaction with the analyte, thus reducing the sensitivity of the MOX films [37].

4.2. Extrinsic Properties

The presence of noble metals nanoparticles, in particular for SnO2:Au sensors, exhibited a good catalytic dissociation performance for oxygen, which promoted the formation of chemically adsorbed oxygen on the material surface [54]. Thus, the reaction rate between chemically adsorbed oxygen and SO2 increased, and the space charge layer narrowed accordingly. This further improved the response of the sensor to SO2.
A further crucial aspect concerns the fact that the SnO2:Au, thanks to the low chemical affinity between Au nanoparticles and sulfur atoms, is capable of preventing irreversible poisoning of the sensing film surface due to the formation of strong metal–sulfide bonds [36].
On the other hand, among SnO2-based gas sensors tested, SnO2:Ag demonstrated a constant response with increasing SO2 concentrations, which is probably due to the production of sulfur stable compounds with Ag metals such as Ag2S consequently poisoning the reactive surface from 2 ppm.

5. Conclusions

The experimentation performed in the present work could guide future developments on modified MOX-based sensors and open new frontiers for olfactory systems dedicated to SO2 monitoring in a wide range of concentrations and in real-world applications. ZnO and WO3 were proven to be selective to different gases and therefore not appropriate for SO2 detection. In contrast, SnO2 was more suitable for the detection of sulfur-containing gases. Therefore, SnO2 was decorated with different noble metals. Addition of Ag or Pt seems to poison the surface, leading to worse sensing performances than pure SnO2. On the other hand, Pd and Au have been demonstrated to be effective in improving the performance of pure SnO2. In particular, SnO2:Au highlighted good sensitivity, cross-selectivity, and humidity-independent behavior for concentrations above 17 RH%. The enhanced reactivity could be explained by the decoration of the SnO2 surface with Au nanoparticles that positively affected the concentration of active sites and acid-base properties of the surface. The effect of catalytic additives on the intrinsic and extrinsic active centers of SnO2 is an important factor that determines the proper conditions for an effective gas-solid interaction. In fact, modification with Au has different effects on the concentration and reactivity of the active sites of pristine SnO2. As a result, modification with Au may cause an increase in the concentration of chemisorbed oxygen species and acid sites, which are the centers of interaction with SO2. Therefore, for the future works, additional studies should be performed under controlled conditions in the presence of the target molecules with the use of in situ or operando techniques, such as operando DRIFT spectroscopy, or chemisorption analysis to explore the active sites of the material in depth.

Supplementary Materials

The following supporting information can be downloaded at: https://mdpi.longhoe.net/article/10.3390/chemosensors12060111/s1, Figure S1: Schematic representation of sensor fabrication; Table S1: The parameters of the calibration fit shown in Figure 2 for ZnO, WO3, SnO2:Pd, SnO2:Au, SnO2, and SnO2:Pt sensors; Table S2: Load resistances of sensors during selectivity measurements; Figure S2: Response to three different concentrations of SO2 around the theoretical LOD of SnO2:Au sensor; Figure S3: Response to 3 ppm of SO2 obtained for SnO2:Au sensor tested with three different measurements over a period of five months; Figure S4: Vout response vs. 0.25 ppm of DMDS; Figure S5: Vout response vs. 5 ppm of ethanol. In the case of WO3, the injection of the gas was different due to a different day of measurement; Figure S6: Vout response vs. 5000 ppm of CO2. In the case of WO3, the injection of the gas was different due to a different day of measurement; Figure S7: Vout response vs. 3 ppm of NO2. In the case of WO3, the injection of the gas was different due to a different day of measurement; Figure S8: Vout response vs. 0.5 ppm of benzene. In the case of WO3, the injection of the gas was different due to a different day of measurement; Figure S9: Vout response vs. 25 ppm of CO. In the case of WO3, the injection of the gas was different due to a different day of measurement; Figure S10: Vout response vs. 5 ppm of SO2. In the case of WO3, the injection of the gas was different due to a different day of measurement.

Author Contributions

Conceptualization, A.R. and B.F.; validation, A.R.; formal analysis, A.R.; investigation, A.R. and A.V.; data curation, A.R.; writing—original draft preparation, A.R. and B.F.; writing—review and editing, E.S. and D.A.; supervision, V.G. and M.M.; funding acquisition, V.G. and B.F. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by POR FSE 2014/2020 by Regione Emilia-Romagna, by PNRR ex D.M. 117 02/03/2023, and by University of Ferrara (FIRD 2022).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in the study are included in the article/Supplementary Material, further inquiries can be directed to the corresponding authors.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Mathieu, Y.; Tzanis, L.; Soulard, M.; Patarin, J.; Vierling, M.; Molière, M. Adsorption of SOx by Oxide Materials: A Review. Fuel Process. Technol. 2013, 114, 81–100. [Google Scholar] [CrossRef]
  2. Meng, Z. Oxidative Damage of Sulfur Dioxide on Various Organs of Mice: Sulfur Dioxide Is a Systemic Oxidative Damage Agent. Inhal. Toxicol. 2003, 15, 181–195. [Google Scholar] [CrossRef] [PubMed]
  3. Barnett, A.G.; Williams, G.M.; Schwartz, J.; Neller, A.H.; Best, T.L.; Petroeschevsky, A.L.; Simpson, R.W. Air Pollution and Child Respiratory Health: A Case-Crossover Study in Australia and New Zealand. Am. J. Respir. Crit. Care Med. 2005, 171, 1272–1278. [Google Scholar] [CrossRef] [PubMed]
  4. Occupational Safety and Health Administration. Available online: https://www.Osha.Gov/Chemicaldata/549 (accessed on 9 May 2024).
  5. López-Cervantes, V.B.; López-Olvera, A.; Obeso, J.L.; Torres, I.K.; Martínez-Ahumada, E.; Carmona-Monroy, P.; Sánchez-González, E.; Solís-Ibarra, D.; Lima, E.; Jangodaz, E.; et al. Robust Co(II)-Based Metal–Organic Framework for the Efficient Uptake and Selective Detection of SO2. Chem. Mater. 2024, 36, 2735–2742. [Google Scholar] [CrossRef]
  6. Hansell, A.; Oppenheimer, C. Health Hazards from Volcanic Gases: A Systematic Literature Review. Arch. Environ. Health 2004, 59, 628–639. [Google Scholar] [CrossRef] [PubMed]
  7. Khan, M.; Rao, M.; Li, Q. Recent Advances in Electrochemical Sensors for Detecting Toxic Gases: NO2, SO2 and H2S. Sensors 2019, 19, 905. [Google Scholar] [CrossRef] [PubMed]
  8. Morulane, K.L.; Swart, H.C.; Motaung, D.E. A Review on Topical Advancement and Challenges of Indium Oxide Based Gas Sensors: Future Outlooks. J. Environ. Chem. Eng. 2024, 12, 112144. [Google Scholar] [CrossRef]
  9. Dey, S.; Mehta, N.S. Automobile Pollution Control Using Catalysis. Resour. Environ. Sustain. 2020, 2, 100006. [Google Scholar] [CrossRef]
  10. Lan, W.; Rao, Y.; Zhao, X.; Zhao, Y.; Min, X.; Wu, Y.; Jiang, Z.; Li, T.; Li, Y.; Chen, H.; et al. Rapid Visual Detection of Sulfur Dioxide Residues in Food Using Acid-Sensitive CdTe Quantum Dots-Loaded Alginate Hydrogel Beads. Food Chem. 2024, 446, 138791. [Google Scholar] [CrossRef]
  11. Peter Singh, L.; Gill, S.S.; Gill, R.; Tuteja, N. Mechanism of Sulfur Dioxide Toxicity and Tolerance in Crop Plants. In Improving Crop Resistance to Abiotic Stress; Tuteja, N., Gill, S.S., Tiburcio, A.F., Tuteja, R., Eds.; Wiley: New York, NY, USA, 2012; pp. 133–163. ISBN 978-3-527-32840-6. [Google Scholar]
  12. Pecunia, V.; Petti, L.; Andrews, J.; Ollearo, R.; Gelinck, G.H.; Nasrollahi, B.; Jailani, J.M.; Li, N.; Kim, J.H.; Ng, T.N.; et al. Roadmap on Printable Electronic Materials for Next-Generation Sensors. Nano Futures 2024. [Google Scholar] [CrossRef]
  13. Zhang, H.; Yu, L.; Li, R.; Zhao, F.; Guo, L.; Wang, P.; Cao, W.; Ding, Z.; Qiao, Z. Layer-Tunable Synthesis of Tetragonal Pr-Doped SnO2 Nanoplates for Enhanced Sensitive SO2 Sensor. Sens. Actuators B Chem. 2024, 409, 135629. [Google Scholar] [CrossRef]
  14. Shin, D.; Besmann, T.M.; Armstrong, B.L. Phase Stability of Noble Metal Loaded WO3 for SO2 Sensor Applications. Sens. Actuators B Chem. 2013, 176, 75–80. [Google Scholar] [CrossRef]
  15. Cao, L.; Yu, L.; Li, S.; Nan, N.; Zhang, C.; Fan, X. In-Situ Growth of Well-Ordered ZnO Nanowire-Networks with Interconnected Junctions for Enhanced SO2 Gas Sensing Properties. Appl. Surf. Sci. 2024, 646, 158899. [Google Scholar] [CrossRef]
  16. Minh, L.H.; Thuy Thu, P.T.; Thanh, B.Q.; Hanh, N.T.; Thu Hanh, D.T.; Van Toan, N.; Hung, C.M.; Van Duy, N.; Van Tong, P.; Hoa, N.D. Hollow ZnO Nanorices Prepared by a Simple Hydrothermal Method for NO2 and SO2 Gas Sensors. RSC Adv. 2021, 11, 33613–33625. [Google Scholar] [CrossRef]
  17. Franco, M.A.; Conti, P.P.; Andre, R.S.; Correa, D.S. A Review on Chemiresistive ZnO Gas Sensors. Sens. Actuators Rep. 2022, 4, 100100. [Google Scholar] [CrossRef]
  18. Kanan, S.; Obeideen, K.; Moyet, M.; Abed, H.; Khan, D.; Shabnam, A.; El-Sayed, Y.; Arooj, M.; Mohamed, A.A. Recent Advances on Metal Oxide Based Sensors for Environmental Gas Pollutants Detection. Crit. Rev. Anal. Chem. 2024, 1–34. [Google Scholar] [CrossRef]
  19. Das, S.; Girija, K.G.; Debnath, A.K.; Vatsa, R.K. Enhanced NO2 and SO2 Sensor Response under Ambient Conditions by Polyol Synthesized Ni Doped SnO2 Nanoparticles. J. Alloys Compd. 2021, 854, 157276. [Google Scholar] [CrossRef]
  20. Berger, F.; Fromm, M.; Chambaudet, A.; Planade, R. Tin Dioxide-Based Gas Sensors for SO2 Detection: A Chemical Interpretation of the Increase in Sensitivity Obtained after a Primary Detection. Sens. Actuators B Chem. 1997, 45, 175–181. [Google Scholar] [CrossRef]
  21. Zhao, C.; Gong, H.; Niu, G.; Wang, F. Ultrasensitive SO2 Sensor for Sub-Ppm Detection Using Cu-Doped SnO2 Nanosheet Arrays Directly Grown on Chip. Sens. Actuators B Chem. 2020, 324, 128745. [Google Scholar] [CrossRef]
  22. Tyagi, P.; Sharma, A.; Tomar, M.; Gupta, V. Metal Oxide Catalyst Assisted SnO2 Thin Film Based SO2 Gas Sensor. Sens. Actuators B Chem. 2016, 224, 282–289. [Google Scholar] [CrossRef]
  23. Gaiardo, A.; Krik, S.; Valt, M.; Fabbri, B.; Tonezzer, M.; Feng, Z.; Guidi, V.; Bellutti, P. Development of a Sensor Array Based on Pt, Pd, Ag and Au Nanocluster Decorated SnO2 for Precision Agriculture. Meet. Abstr. 2021, MA2021-01, 1550. [Google Scholar] [CrossRef]
  24. Spagnoli, E.; Krik, S.; Fabbri, B.; Valt, M.; Ardit, M.; Gaiardo, A.; Vanzetti, L.; Della Ciana, M.; Cristino, V.; Vola, G.; et al. Development and Characterization of WO3 Nanoflakes for Selective Ethanol Sensing. Sens. Actuators B Chem. 2021, 347, 130593. [Google Scholar] [CrossRef]
  25. Gaiardo, A.; Fabbri, B.; Giberti, A.; Guidi, V.; Bellutti, P.; Malagù, C.; Valt, M.; Pepponi, G.; Gherardi, S.; Zonta, G.; et al. ZnO and Au/ZnO Thin Films: Room-Temperature Chemoresistive Properties for Gas Sensing Applications. Sens. Actuators B Chem. 2016, 237, 1085–1094. [Google Scholar] [CrossRef]
  26. Rossi, A.; Spagnoli, E.; Tralli, F.; Marzocchi, M.; Guidi, V.; Fabbri, B. New Approach for the Detection of Sub-Ppm Limonene: An Investigation through Chemoresistive Metal-Oxide Semiconductors. Sensors 2023, 23, 6291. [Google Scholar] [CrossRef]
  27. Rossi, A.; Fabbri, B.; Spagnoli, E.; Gaiardo, A.; Valt, M.; Ferroni, M.; Ardit, M.; Krik, S.; Pedrielli, A.; Vanzetti, L.; et al. Functionalization of Indium Oxide for Empowered Detection of CO2 over an Extra-Wide Range of Concentrations. ACS Appl. Mater. Interfaces 2023, 15, 33732–33743. [Google Scholar] [CrossRef]
  28. Gaiardo, A.; Fabbri, B.; Giberti, A.; Valt, M.; Gherardi, S.; Guidi, V.; Malagù, C.; Bellutti, P.; Pepponi, G.; Casotti, D.; et al. Tunable Formation of Nanostructured SiC/SiOC Core-Shell for Selective Detection of SO2. Sens. Actuators B Chem. 2020, 305, 127485. [Google Scholar] [CrossRef]
  29. Spagnoli, E.; Gaiardo, A.; Fabbri, B.; Valt, M.; Krik, S.; Ardit, M.; Cruciani, G.; Della Ciana, M.; Vanzetti, L.; Vola, G.; et al. Design of a Metal-Oxide Solid Solution for Sub-Ppm H2 Detection. ACS Sens. 2022, 7, 573–583. [Google Scholar] [CrossRef]
  30. Occupational Safety and Health Administration. Available online: https://www.Osha.Gov/Chemicaldata/491 (accessed on 9 May 2024).
  31. Centers for Disease Control and Prevention, The National Institute for Occupational Safety and Health. Available online: https://www.Cdc.Gov/Niosh/Idlh/10102440.Html (accessed on 9 May 2024).
  32. American Conference of Governmental Industrial Hygienist. Available online: https://www.Acgih.Org/Carbon-Monoxide-3/ (accessed on 9 May 2024).
  33. American Conference of Governmental Industrial Hygienist. Available online: https://www.Acgih.Org/Carbon-Dioxide/ (accessed on 9 May 2024).
  34. Occupational Safety and Health Administration. Available online: https://www.Osha.Gov/Chemicaldata/911 (accessed on 9 May 2024).
  35. Centers for Disease Control and Prevention, The National Institute for Occupational Safety and Health. Available online: https://www.Cdc.Gov/Niosh/Pel88/10102-44.Html (accessed on 9 May 2024).
  36. Rodriguez, J. The Chemical Properties of Bimetallic Surfaces: Importance of Ensemble and Electronic Effects in the Adsorption of Sulfur and SO2. Prog. Surf. Sci. 2006, 81, 141–189. [Google Scholar] [CrossRef]
  37. Ciana, M.D.; Valt, M.; Fabbri, B.; Gaiardo, A.; Spagnoli, E.; Krik, S.; Bernardoni, P.; Gilli, N.; Migliori, A.; Quaranta, A.; et al. SO2 Sensing Mechanism of Nanostructured SiC-SiOxC Core Shell: An Operando DRIFT Investigation. Sens. Actuators B Chem. 2022, 371, 132497. [Google Scholar] [CrossRef]
  38. Fu, X.; Liu, J.; Han, T.; Zhang, X.; Meng, F.; Liu, J. A Three-Dimensional Hierarchical CdO Nanostructure: Preparation and Its Improved Gas-Diffusing Performance in Gas Sensor. Sens. Actuators B Chem. 2013, 184, 260–267. [Google Scholar] [CrossRef]
  39. Khune, A.S.; Padghan, V.; Bongane, R.; Narwade, V.N.; Dole, B.N.; Ingle, N.N.; Tsai, M.-L.; Hianik, T.; Shirsat, M.D. Highly Selective Chemiresistive SO2 Sensor Based on a Reduced Graphene Oxide/Porphyrin (rGO/TAPP) Composite. J. Electron. Mater. 2023, 52, 8108–8123. [Google Scholar] [CrossRef]
  40. Kumar, R.; Kaur, A. Chemiresistive Gas Sensors Based on Thermally Reduced Graphene Oxide for Sensing Sulphur Dioxide at Room Temperature. Diam. Relat. Mater. 2020, 109, 108039. [Google Scholar] [CrossRef]
  41. Peng, K.; Li, Q.; Ma, M.; Li, N.; Sheng, H.; Li, H.; Huang, Y.; Yun, F. Acidic Gas Determination Using Indium Tin Oxide-Based Gas Sensors. Sensors 2024, 24, 1286. [Google Scholar] [CrossRef] [PubMed]
  42. Yang, J.; Wang, T.; Zhu, C.; Yin, X.; Dong, P.; Wu, X. AgNWs@SnO2/CuO Nanocomposites for Ultra-Sensitive SO2 Sensing Based on Surface Acoustic Wave with Frequency-Resistance Dual-Signal Display. Sens. Actuators B Chem. 2023, 375, 132966. [Google Scholar] [CrossRef]
  43. Dhingra, V.; Kumar, S.; Kumar, R.; Garg, A.; Chowdhuri, A. Room Temperature SO2 and H2 Gas Sensing Using Hydrothermally Grown GO–ZnO Nanorod Composite Films. Mater. Res. Express 2020, 7, 065012. [Google Scholar] [CrossRef]
  44. Tyagi, P.; Sharma, A.; Tomar, M.; Gupta, V. A Comparative Study of RGO-SnO2 and MWCNT-SnO2 Nanocomposites Based SO2 Gas Sensors. Sens. Actuators B Chem. 2017, 248, 980–986. [Google Scholar] [CrossRef]
  45. Tyagi, P.; Sharma, A.; Tomar, M.; Gupta, V. SnO2 Thin Film Sensor Having NiO Catalyst for Detection of SO2 Gas with Improved Response Characteristics. Sens. Actuators B Chem. 2017, 248, 998–1005. [Google Scholar] [CrossRef]
  46. Liu, Y.; Xu, X.; Chen, Y.; Zhang, Y.; Gao, X.; Xu, P.; Li, X.; Fang, J.; Wen, W. An Integrated Micro-Chip with Ru/Al2O3/ZnO as Sensing Material for SO2 Detection. Sens. Actuators B Chem. 2018, 262, 26–34. [Google Scholar] [CrossRef]
  47. Fu, Y.; Li, J.; Xu, H. SnO2 Recycled from Tin Slime for Enhanced SO2 Sensing Properties by NiO Surface Decoration. Mater. Sci. Semicond. Process. 2020, 114, 105073. [Google Scholar] [CrossRef]
  48. Ivanov, P.; Llobet, E.; Vilanova, X.; Brezmes, J.; Hubalek, J.; Correig, X. Development of High Sensitivity Ethanol Gas Sensors Based on Pt-Doped SnO2 Surfaces. Sens. Actuators B Chem. 2004, 99, 201–206. [Google Scholar] [CrossRef]
  49. Dhongade, V.; Sharma, L.; Kahandal, A.; Aiyer, R.C.; Rajdeo, K.; Tagad, C. Pd Doped SnO2 Based Room Temperature Operable Resistive Sensor for the Detection of Ethanol Vapors. Mater. Today Proc. 2021, 46, 6437–6444. [Google Scholar] [CrossRef]
  50. Aleksanyan, M.; Sayunts, A.; Shahkhatuni, G.; Simonyan, Z.; Shahnazaryan, G.; Aroutiounian, V. Gas Sensor Based on ZnO Nanostructured Film for the Detection of Ethanol Vapor. Chemosensors 2022, 10, 245. [Google Scholar] [CrossRef]
  51. Gaskov, A.; Rumyantseva, M.; Marikutsa, A. Tin Oxide Materials: Synthesis, Properties, and Applications; Elsevier: Amsterdam, The Netherlands, 2020; Chapter 7; pp. 163–212. [Google Scholar] [CrossRef]
  52. Sahm, T.; Gurlo, A.; Bârsan, N.; Weimar, U. Basics of Oxygen and SnO2 Interaction; Work Function Change and Conductivity Measurements. Sens. Actuators B Chem. 2006, 118, 78–83. [Google Scholar] [CrossRef]
  53. Marikutsa, A.; Rumyantseva, M.; Konstantinova, E.A.; Gaskov, A. The Key Role of Active Sites in the Development of Selective Metal Oxide Sensor Materials. Sensors 2021, 21, 2554. [Google Scholar] [CrossRef]
  54. Wu, Z.; Wang, Y.; Wang, P.; Cheng, X.; Wang, Q.; Yang, Y.; Luo, Y.; An, B.; **e, E. UV Irradiation of Au-Modified SnO2 Nanorods Enabling the Rapid Detection of Low-Concentration NO2 at Room Temperature. IEEE Sens. J. 2023, 23, 1803–1808. [Google Scholar] [CrossRef]
Figure 1. Response vs. working temperature of the sensors to 1 ppm of SO2 in dry conditions. The temperature and RH% of the chamber are 25–30 °C and 2 RH%, respectively, during the measurements. The response was calculated by Equation (2).
Figure 1. Response vs. working temperature of the sensors to 1 ppm of SO2 in dry conditions. The temperature and RH% of the chamber are 25–30 °C and 2 RH%, respectively, during the measurements. The response was calculated by Equation (2).
Chemosensors 12 00111 g001
Figure 2. Response vs. several concentrations of SO2 in dry conditions. The calibration curves of SnO2:Au, SnO2:Pd, WO3, SnO2, SnO2:Pt, and ZnO are fitted with asymptotic function.
Figure 2. Response vs. several concentrations of SO2 in dry conditions. The calibration curves of SnO2:Au, SnO2:Pd, WO3, SnO2, SnO2:Pt, and ZnO are fitted with asymptotic function.
Chemosensors 12 00111 g002
Figure 3. Humidity effect on sensor responses towards 5 ppm of SO2.
Figure 3. Humidity effect on sensor responses towards 5 ppm of SO2.
Chemosensors 12 00111 g003
Figure 4. A total of 6 ppm for four SO2 cycles of SnO2:Au at (a) dry and (b) 30% relative humidity.
Figure 4. A total of 6 ppm for four SO2 cycles of SnO2:Au at (a) dry and (b) 30% relative humidity.
Chemosensors 12 00111 g004
Figure 5. Response of the sensors vs. 0.5 ppm of benzene, 5 ppm of ethanol, 3 ppm of NO2, 0.25 ppm of DMDS, 25 ppm of CO, 5000 ppm of CO2, and 5 ppm of SO2 under dry air condition (≈2 RH%).
Figure 5. Response of the sensors vs. 0.5 ppm of benzene, 5 ppm of ethanol, 3 ppm of NO2, 0.25 ppm of DMDS, 25 ppm of CO, 5000 ppm of CO2, and 5 ppm of SO2 under dry air condition (≈2 RH%).
Chemosensors 12 00111 g005
Figure 6. Dynamic sensor responses of SnO2:Au towards 5 ppm of SO2 in an atmosphere of 1 ppm of ethanol in (a) dry and (b) wet conditions (30 RH%). The duration of the measurements depends on the time required for most of the sensors to achieve the steady state in presence of SO2.
Figure 6. Dynamic sensor responses of SnO2:Au towards 5 ppm of SO2 in an atmosphere of 1 ppm of ethanol in (a) dry and (b) wet conditions (30 RH%). The duration of the measurements depends on the time required for most of the sensors to achieve the steady state in presence of SO2.
Chemosensors 12 00111 g006
Figure 7. Reaction mechanisms on the surface: (a) dry air, (b) under SO2 exposure in dry air, (c) wet air, and (d) under SO2 exposure in humid air.
Figure 7. Reaction mechanisms on the surface: (a) dry air, (b) under SO2 exposure in dry air, (c) wet air, and (d) under SO2 exposure in humid air.
Chemosensors 12 00111 g007
Table 1. Optimal operating temperature of the sensors for SO2 detection, obtained from Figure 1.
Table 1. Optimal operating temperature of the sensors for SO2 detection, obtained from Figure 1.
SensorOptimal Working
Temperature [°C]
WO3200
ZnO400
SnO2:Pd400
SnO2:Au400
SnO2400
SnO2:Pt300
SnO2:Ag400
Table 2. Comparison of gas detection features exhibited by SO2 chemoresistive sensors identified in the literature.
Table 2. Comparison of gas detection features exhibited by SO2 chemoresistive sensors identified in the literature.
Sensing
Materials
Operating Temperature
[°C]
Concentration
[ppm]
Response
Formula
Response in Dry/Wet AirLODReference
rGO/TAPPRT5(Rgas − Rair)/Rair-/0.45
at 30 RH%
5 ppm[39]
Ni/SnO2RT5Rgas/Rair-/2.4
at ∼55–65 RH%
-[19]
rGORT5(Rair − Rgas)/Rair3.21/--[40]
ITO(In2O3/SnO2)2405Rair/Rgas-/1.8
at 35 RH%
-[41]
Ag NWs@SnO2/CuO8050Rair/Rgas2.1/-0.25 ppm[42]
GO/ZnO NRRT25Rgas/Rair2.97/--[43]
RGO/SnO2 nanocomposites6010(Rair/Rg) − 11.2/--[44]
SnO2/NiO180500(Rair/Rg) − 156/50
at dry/70 RH%
-[45]
Ru/Al2O3/ZnO350100(Rair − Rgas)/Rair65/-5 ppm[46]
NiO:SnO2240100Rair/Rgas10/9.1
at dry/70 RH%
-[47]
SnO2:Au4005(Ggas − Gair)/Gair10.4/2
at dry/30 RH%
0.48 ppm *This work
The “-” stands for not mentioned in the respective article. Rair and Rgas are the resistance in dry air and in presence of SO2, respectively. * Theoretical value.
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

Rossi, A.; Spagnoli, E.; Visonà, A.; Ahmed, D.; Marzocchi, M.; Guidi, V.; Fabbri, B. SO2 Detection over a Wide Range of Concentrations: An Exploration on MOX-Based Gas Sensors. Chemosensors 2024, 12, 111. https://doi.org/10.3390/chemosensors12060111

AMA Style

Rossi A, Spagnoli E, Visonà A, Ahmed D, Marzocchi M, Guidi V, Fabbri B. SO2 Detection over a Wide Range of Concentrations: An Exploration on MOX-Based Gas Sensors. Chemosensors. 2024; 12(6):111. https://doi.org/10.3390/chemosensors12060111

Chicago/Turabian Style

Rossi, Arianna, Elena Spagnoli, Alan Visonà, Danial Ahmed, Marco Marzocchi, Vincenzo Guidi, and Barbara Fabbri. 2024. "SO2 Detection over a Wide Range of Concentrations: An Exploration on MOX-Based Gas Sensors" Chemosensors 12, no. 6: 111. https://doi.org/10.3390/chemosensors12060111

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