Next Article in Journal / Special Issue
Elucidation of Response Mechanism of a Potentiometric Sweetness Sensor with a Lipid/Polymer Membrane for Uncharged Sweeteners
Previous Article in Journal
A Long Short-Term Memory Neural Network Based Simultaneous Quantitative Analysis of Multiple Tobacco Chemical Components by Near-Infrared Hyperspectroscopy Images
Previous Article in Special Issue
Odor Recognition of Thermal Decomposition Products of Electric Cables Using Odor Sensing Arrays
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Au Functionalized SnS2 Nanosheets Based Chemiresistive NO2 Sensors

1
School of Microelectronics, Dalian University of Technology, Dalian 116024, China
2
Faculty of Electrical and Electronic Engineering, Dalian University of Technology, Dalian 116024, China
3
Laboratory of Chemistry and Physics of Semiconductor and Sensor Materials, Chemistry Department, Moscow State University, Leninskie Gory 1-3, 199991 Moscow, Russia
*
Authors to whom correspondence should be addressed.
Chemosensors 2022, 10(5), 165; https://doi.org/10.3390/chemosensors10050165
Submission received: 5 April 2022 / Revised: 25 April 2022 / Accepted: 27 April 2022 / Published: 28 April 2022
(This article belongs to the Special Issue Bioinspired Chemical Sensors and Micro-Nano Devices)

Abstract

:
Layered Au/SnS2 nanosheet based chemiresistive-type sensors were successfully prepared by using an in situ chemical reduction method followed by the hydrothermal treatment. SEM and XRD were used to study the microscopic morphology and crystal lattice structure of the synthesis of Au/SnS2 nanomaterials. TEM and XPS characterization were further carried out to verify the formation of the Schottky barrier between SnS2 nanosheets and Au nanoparticles. The as-fabricated Au/SnS2 nanosheet based sensor demonstrated excellent sensing properties to low-concentrations of NO2, and the response of the sensor to 4 ppm NO2 at 120 °C was approximately 3.94, which was 65% higher than that of the pristine SnS2 (2.39)-based sensor. Moreover, compared to that (220 s/520 s) of the pristine SnS2-based sensor, the response/recovery time of the Au/SnS2-based one was significantly improved, reducing to 42 s/127 s, respectively. The sensor presents a favorable long-term stability with a deviation in the response of less than 4% for 40 days, and a brilliant selectivity to several possible interferents such as NH3, acetone, toluene, benzene, methanol, ethanol, and formaldehyde. The Schottky barrier that formed at the interface between the SnS2 nanosheets and Au nanoparticles modulated the conducting channel of the nanocomposites. The “catalysis effect” and “spillover effect” of noble metals jointly improved the sensitivity of the sensor and effectively decreased the response/recovery time.

1. Introduction

As a precursor of photochemical smog and acid rain, nitrogen dioxide gas is a typical poisonous and corrosive atmospheric pollution. It is produced from the combustion of fossil fuel and automobile exhaust [1]. Excessive nitrogen dioxide gas generated in the air would lead to a decrease in atmospheric visibility, acidification of soil and water, and corrosion of buildings [2,3]. It also poses a threat to the health of human beings. Long-term exposure to ppm-level NO2 would be particularly dangerous to human health, especially to the respiratory system, and may cause lung malfunction [4]. Furthermore, the detection of nitrogen oxides (NOx) in the human respiration system is currently applied to the diagnosis of human diseases, which is a non-invasive diagnostic method [5]. For instance, ppb-level NOx in human exhaled gases is an important biomarker for diagnosing inflammatory airway disease [6,7]. In addition, NOx can be used as biological indicators for lung diseases [8,9,10]. Therefore, it is of great significance to develop a simple and reliable NO2 gas sensor.
Metal sulfides have triggered the interests of researchers in detecting NO2 gas at low temperatures or even room temperature due to its unique crystal structure and hysic-chemical properties [11,12,13,14]. Compared with traditional metal oxides, most metal sulfides are stacked by a two-dimensional single-layer sheet structure, which make it a narrower band gap, a larger specific surface area, and more active sites. These advantages are beneficial to the adsorption and desorption of gas molecules on the surface of the nanomaterials. Among these metal sulfide materials, tin disulfide (SnS2) is a typical layered two-dimension structure nanomaterial, which has been studied in the field of gas sensing detection [15,16,17]. Ou et al., synthesized a novel gas sensor based on two-dimensional SnS2 flakes, which has a good sensitivity to NO2 [18]. Liu et al., investigated the sensing behavior of nanosheet-assembled SnS2 nanoflowers prepared by a simple hydrothermal method [19]. Gu et al., explored the gas sensing characteristics of a SnS2 nanosheet under visible light illustrated at room temperature [20]. However, it is still a challenge to improve the response and recovery speeds of SnS2-based sensors.
The modification and doping of noble metal nanoparticles on gas sensing materials have been proven to be one of the most effective methods to prolong the lifetime of semiconductor carriers and improve the adsorption capacity of gas molecules on sensitive materials, which could be attributed to the “electronic sensitization” and “spillover effect” of noble metals [21,22,23]. Noble metal/semiconductor heterostructure is an ideal mode for improving the gas performance of semiconductor materials. For instance, Au nanoparticles exhibit excellent catalytic effect at sizes smaller than 10 nm, and it is also the least active noble metal toward atoms and molecules [24,25]. Functionalization of Au nanoparticles with high catalytic activity can not only promote the adsorption of gas molecules on the surface of the material, but can also accelerate the chemical reaction between the adsorbed oxygen and the target gas.
Hence, Au nanoparticles with excellent catalytic performance were selected to decorate the SnS2 material in order to enhance the response value and improve the response/recovery speeds of the SnS2-based sensor to NO2 at the low temperature region by the synergistic effect of “electron sensitization” and “chemical sensitization”. The Au functionalized SnS2 nanosheets with different Au contents was synthesized by a simple in situ chemical reduction method and the gas sensing performance was also systematically investigated. The research results show that the sensor-based on the Au/SnS2-0.5 sample greatly enhanced the sensing performance to NO2 of the SnS2-based sensor at low temperature region (129 °C), which exhibited a fast response/recovery speed, excellent signal reproducibility, a good long-term stability, and anti-interference ability. The sensing mechanism of the sensor is discussed in detail.

2. Experimental

2.1. Preparation of SnS2 Nanosheet

The SnS2 nanosheets were fabricated though the hydrothermal method. A total of 233 mg SnCl2·2H2O, 499 mg Na2S2O3·5H2O, and 160 mg sulfur powder were added into a 50 mL Teflon-lined autoclave, which had been pre-filled with deionized water to 80% of its capacity. The Teflon-lined autoclave was maintained in a muffle furnace at 200 °C for 12 h, and then naturally cooled to room temperature. The filtered yellow precipitates were washed several times with deionized water, carbon disulfide, and ethanol to remove the residuals of sulfur and organic solvents.

2.2. Decoration of Au Nanoparticles on the Surface of SnS2 Nanosheets

Figure 1 illustrates the scheme for decorating the Au nanoparticles on the surface of SnS2 nanosheets. Au functionalized SnS2 nanosheets were performed by in situ chemical reduction of the HAuCl4 by using C6H5Na3O7 as the reducing agent. A total of 100 mg of as-synthesized SnS2 nanosheets were dispersed into 30 mL of deionized water and ultrasonically sonicated for 30 min to form a uniform dispersion. Subsequently, 0.05 mL of HAuCl4 (0.01 M) was added into the above solution. After 10 min, 0.15 mL of C6H14N2O2 solution (0.01 M) was added into the mixture and stirred with a magnetic stirrer at room temperature for 30 min. After stirring for 30 min, a certain amount of Au ions were doped on the SnS2 nanosheets, and then 0.1 mL of 0.1 M trisodium citrate solution was added and stirred for 30 min, so that the Au ions doped on the SnS2 nanosheets could be chemically reduced to Au nanoparticles. The resultant product was washed several times with deionized water and absolute ethyl alcohol to remove the impurities. The washed precipitate was placed in a drying oven and dried at 60 °C for 12 h, thus obtaining the composite material of Au/SnS2 nanosheets, which was defined as Au/SnS2-0.5. For a comparison, the corresponding amounts of chloroauric acid, L-lysine, and trisodium citrate solution were added according to the proportion to prepare the Au/SnS2-0.1, Au/SnS2-0.3, Au/SnS2-1, and Au/SnS2-1.5 composites.

2.3. Preparation of Sensors

Figure 2 shows the structural diagram of the Au interdigital electrode with heating layer function, which was used as the substrate of the gas-sensing material to transmit the chemiresistive sensing signal in this work. The front side of the interdigital electrode is the gold electrode, which is made by printing gold paste on the surface of the alumina ceramic substrate through the screen printing process and then sintering at high temperature (850 °C). The width of each golden interdigital finger and the distance between adjacent interdigitated fingers were both 200 μm, and the gas sensing material was deposited on the surface of the gold electrode by the dip-coating method. The backside of the interdigital electrode had the heating function, which consisted of a RuO2 resistive layer and Ag/Pb pads. The RuO2 resistance layer will reach the different working temperatures by applying corresponding DC voltages between the ends of the Ag/Pb pads. A sample of 10 mg of powdered Au/SnS2 nanocomposite and a small amount of deionized water were added to a mortar and ground for 5 min to form a uniform slurry. The resulting paste material was then dripped onto the forked finger electrode, which was then dried in an oven at 60 °C for 12 h to obtain a sensor for sensitive performance testing.

2.4. Material Characterization

The crystal structure of the as-prepared Au/SnS2 was recorded using X-ray diffraction (XRD: XRD-6000, Shimsdzu Corp, Tokyo, Japan) at a scanning rate of 0.26°/s from 10° to 75°. The scanning electron microscope (SEM: S-3000 N, Hitachi, Japan) and the transmission electron microscope (TEM: Tecnai G220, Philips, The Netherlands) were used to record the microscopic morphology and crystal lattice sizes of the synthesis of Au/SnS2 nanomaterials. The information of the elemental and chemical states were recorded by a Thermo ESCALAB 250Xi X-ray photoelectron spectrometer with an Al Kα source (hv = 1486.6 eV). The beam spot diameter of the monochrome Al source was 500 μm and the power was 150 W.

2.5. Gas Sensing Measurements

The gas sensing performances of the sensors were measured by using fully automatic computer controlled gas distribution and data acquisition systems under a dynamic gas flow region. Various parameters such as the bias voltage, the testing time, and the concentration of the target gases were controlled by the computer. The Agilent 34465A computer-controlled multimeter recorded the electrical resistance of the sensors. The responses (S) of the sensors were defined as the relative change in the resistance of the sensors in the background and those in the tested gas (Equations (1) and (2)):
S = R g R a   ( for   NO 2 )
S = R a R g   for   other   reducing   gases
where Ra and Rg are the stable resistances of sensor in the air and in the tested gases. The response or recovery times of the sensors were defined as the time taken for the sensor to achieve 90% change in the full magnitude change of the gas response.

3. Results and Discussion

3.1. Microstructure Characterization of As-Synthesized Au/SnS2 Nanosheets

Figure 3 shows the XRD spectra of the Au/SnS2 nanocomposites with different proportions. It can be clearly observed from Figure 3a in the XRD spectrogram that there were obvious diffraction peaks at 15.03 and 28.20° in the spectrograms of Au/SnS2-0.1, Au/SnS2-0.3, Au/SnS2-0.5, Au/SnS2-1, and Au/SnS2-1.5, respectively. The (001), (100), (101), (110), and (111) crystal planes of SnS2 corresponding to the 2T structure were consistent with JCPDF#23-0677. However, the existence of elemental Au was not detected in the composite materials with 0.1%, 0.3%, 0.5%, and 1%, which may be due to the much lower content of Au doped in the composite materials than the lowest detection limit of XRD equipment. However, when the doping ratio of Au nanoparticles in the composite reached 1.5%, a weak diffraction peak of the Au element appeared at 38.18° (corresponding card: JCPDF#04-0784).
It can be observed from the enlarged view of the local diffraction peak on the (001) crystal plane in Figure 3b that the diffraction peaks of the Au/SnS2 composite materials had a slight shift with an angle of 0.17 to 0.22°. It indicates that during the doping process, some Au atoms may enter the SnS2 lattice and be doped into the SnS2 lattice or sulfur vacancy. At the same time, with the increase in Au doping content, the diffraction peak intensity of the Au/SnS2 composites became weak, which may be related to the agglomeration of SnS2 nanosheets caused by the use of a higher concentration of trisodium citrate in the process of reduction of a high proportion of Au, which can also be further confirmed from the later SEM results of the composites. In addition, no other diffraction peaks for any impurities were found in the XRD spectra, which also proved that the prepared Au/SnS2 nanomaterials had high purity.
Figure 4 shows the SEM images of the SnS2, Au/SnS2-0.5, and Au/SnS2-1.5 nanocomposites. It can be observed that the pristine SnS2 had a layered-nanosheet structure with particle sizes ranging from 50 nm to 500 nm, as shown in Figure 4a. In the Au/SnS2 nanocomposite, due to the reduced Au nanoparticle size being smaller than the detection limit of the SEM equipment, no obvious Au nanoparticles were observed, and the specific shape of Au nanoparticles could not be clearly observed. However, with the increase in Au doping concentration, the surface of the SnS2 nanosheets in the composites began to agglomerate, and it was obviously observed that there were agglomerated SnS2 nanosheets in Au/SnS2-1.5, as shown in Figure 4b,c. This phenomenon was perhaps due to the reduction in the high doping amount of Au, and the content of the reducing agent trisodium citrate also increased, which resulted in the agglomeration between small and medium-sized SnS2 nanosheets during the reduction process. Figure 4d shows the EDS characterization results of the Au/SnS2-0.5 nanocomposites. It can be observed that the Au/SnS2-0.5 nanocomposites contained Au, S, and Sn elements with the atomic ratio of Au:S:Sn of 66.99:33.36:0.15, and the mass percentage of Au was approximately 0.47 wt%, which was slightly lower than the nominal value (0.5 wt%). This result indicates that the Au element was indeed introduced into the Au/SnS2 nanocomposites in this experiment and the existing form of the Au element will be studied by further characterization.
To further investigate the crystal phase structure of the Au/SnS2 nanocomposites, TEM images of the SnS2 nanosheets and Au/SnS2-0.5 nanocomposites are shown in Figure 5. Figure 5a,b reveals that the pristine SnS2 nanosheets had a hexagonal flake morphology, and the distance between two adjacent groups of lattice stripes was 0.290 nm, which corresponded to the (002) crystal plane of SnS2, and no lattice stripes of other substances appeared. Figure 5c shows the morphology of the Au/SnS2-0.5 nanocomposites had no obvious change, and it was still a layered structure. It notes that rough nanoparticles appeared on the flat surface of the SnS2 nanosheets in Figure 5d, with an average size of about 5 nm, and the spacing between two adjacent groups of stripes was 0.236 nm, which is consistent with the (111) crystal plane spacing of Au. This demonstrates that the small-sized nanoparticles on the surface of SnS2 were Au nanoparticles, which further confirms that the Au nanoparticles had been successfully decorated on the SnS2 nanosheets by the chemical reduction method.
X-ray photoelectron spectroscopy (XPS) was performed to characterize and analyze the chemical components and elemental states on the surface of pristine SnS2 nanosheets and Au/SnS2 nanocomposites. Figure 6 shows the XPS spectra of the SnS2 nanosheets and Au/SnS2-0.5 composite nanomaterials. The C1s peak at 284 eV belonged to foreign hydrocarbons from the XPS instrument itself, which was used as the calibration peak. Figure 6a shows the full survey spectra of the pristine SnS2 and Au/SnS2-0.5 samples, which all contained characteristic peaks of the Sn, S, and O elements. There was a weak peak at around 86 eV, which was assigned to the Au0 in the Au/SnS2-0.5 nanocomposites. Figure 6b–e displays the high-resolution spectra and fitting curves of O ls, Sn 3d, S 2p, and Au 4f of the SnS2 nanosheets and Au/SnS2-0.5 composites in a small energy range.
As shown in Figure 6b, the peak that appeared at 532.2 eV in the O ls spectrum of the pristine SnS2 nanosheet corresponded to the chemisorbed oxygen ( O 2 ads or O ads ) on the surface of the sample [26,27]. No lattice oxygen (O2−) could be observed, indicating that SnS2 did not contain the impurity phase of metal oxides such as SnO2. Compared with the pristine SnS2, the O 1s peak of the Au/SnS2-0.5 composite had a higher peak intensity of chemisorbed oxygen, indicating that the decoration of noble metals promoted the adsorption of chemisorbed oxygen molecules by the sample. The chemisorbed oxygen on the surface of the Au/SnS2-0.5 sample could participate in subsequent redox reactions with NO2 molecules. Thus, it is beneficial to improve the gas-sensing performance of the Au/SnS2-0.5-based sensor.
The refined spectra of the Sn 3d of pristine SnS2 and Au/SnS2 composites was further analyzed as shown in Figure 6c. It was found that both samples contained Sn 3d peaks, and the energy difference between Sn 3d5/2 and Sn 3d3/2 peaks was approximately 8.4 eV, which indicates that the chemical state of SnS2 did not change during the Au reduction process, and Sn ions were still at the highest oxidation state of +4 [28]. Compared to the pristine SnS2, the fitting peak of Sn 3d in the Au/SnS2-0.5 composite shifted to higher binding energy, and the binding energy increased to approximately 0.2 eV. This shift indicates that Au has a certain influence on the electronic state of SnS2 shell, and the electrons transfer from the conduction band of SnS2 to Au.
Figure 6d shows the S 2p spectrums of two samples. The two peaks centered at 161.85 eV and 163.01 eV corresponded to S 2p3/2 and S 2p1/2, respectively [29,30]. It was also found that the fitting peak of S 2p in the Au/SnS2-0.5 sample shifted to a higher energy level. This was due to the work function of Au (5.10 eV) being higher than that of SnS2, which causes the electrons in the outer shell of SnS2 to be attracted to the core of Au. This kind of charge transfer regulates the electron concentration on the surface of Au/SnS2-0.5, resulting in electron-hole separation. The spectra of Au elements in the two samples are shown in Figure 6e. There was no Au characteristic peak in the SnS2 material, but there were obvious Au double peaks in the Au/SnS2-0.5 sample, which were located at 87.9 eV and 84.2 eV, respectively, corresponding to Au 4f7/2 and Au4f5/2 [31,32]. This indicates that Au in the Au/SnS2-0.5 composite was successfully reduced and SnS2 was decorated by the metal Au.

3.2. Gas-Sensing Property of Au/SnS2 Nanosheets

For the gas sensor, the working temperature is one of the important parameters that affect the performance of the sensor. Figure 7 reveals the response of SnS2, Au/SnS2-0.1, Au/SnS2-0.3, Au/SnS2-0.5, Au/SnS2-1, and Au/SnS2-1.5-based sensors to 4 ppm NO2 at the operating temperatures from 100 °C to 180 °C. The response of pristine SnS2-based sensor first increased and then decreased with the increase in temperature, and the response value reached the maximum at the working temperature of 120 °C, which was approximately 2.39. The response of Au-doped SnS2-based sensors had the same changing trend with the working temperatures. At low content Au, the optimal working temperature of the Au/SnS2-0.1, Au/SnS2-0.3, and Au/SnS2-0.5-based sensors was at 120 °C, which indicates that Au decoration does not reduce the working temperature of the SnS2 nanomaterial, but improved the NO2 sensitivity of the sensor. With the further increase in Au content, the optimal working temperature of Au/SnS2-1 and Au/SnS2-1.5-based sensors was 140 °C, which was higher than that of the pristine SnS2-based one. This is perhaps due to the increase in the proportion of sodium citrate added in reducing the high concentration Au, which causes the small-sized SnS2 nanosheets to agglomerate during the reduction of noble metals, and the materials were denser. This is consistent with the results of the SEM, as shown in Figure 4.
The response of the composite material was higher than that of pristine SnS2 with the increase in Au concentration when less than 1 wt%. The response values of the Au/SnS2-0.1, Au/SnS2-0.3, Au/SnS2-0.5, and Au/SnS2-1-based sensors at 120 °C were 2.54, 3.12, 3.94, and 2.94, respectively. When the doping ratio was greater than 1%, the response of the Au/SnS2-1.5-based sensors instead became lower than that of the pristine SnS2-based one. According to literature [33,34], when the size of Au nanoparticles is less than 5 nm, it shows excellent catalytic activities. The Au nanoparticles became agglomerates to form large-sized particles when Au content was high as shown in SEM results. The agglomerated Au nanoparticles usually have poor catalytic performance, which might induce the poor sensing properties of Au/SnS2-1.5-based sensor.
Figure 8 shows the response and resistance curves of the Au/SnS2 composites and SnS2 materials with different Au amounts to 4 ppm NO2 at their respective optimal working temperatures. Figure 8a shows that when the composite was doped with Au at a low concentration, the resistance of the composite increased with the increase in the doping amount of Au particles. However, further increasing the Au content to more than 1%, the resistance of the composite material decreased. As the Au content increased, the number of Au nanoparticles gathered on the surface of the SnS2 increased, the particle size became larger, and agglomerates between Au nanoparticles occurred, which improved the conductivity of the composite material and reduced the resistance. When the sensor was exposed to 4 ppm NO2, the resistance of the sensor increased rapidly. Then, the resistance of the sensor achieved a stable plateau. The resistance of the sensor decreased and gradually returned to the initial resistance after NO2 was terminated. Both SnS2 and the Au/SnS2 composites showed typical n-type semiconductor characteristics.
Figure 8b shows the response curves of the Au/SnS2 composites with different Au content and the pristine SnS2-based sensors to 4 ppm NO2 at the optimal working temperatures. When the Au concentration increased from 0.1 wt% to 0.5 wt%, the sensing performance of the composite material was better than that of pristine SnS2, and the response value to 4 ppm NO2 increased with the increase in Au content. When the Au concentration was 0.5 wt%, the sensing characteristics of the composite material were the best. When the Au content was higher than 1 wt%, the response value of the composite gradually decreased. At the same time, the response and recovery time of the composite materials to 4 ppm NO2 at the optimal temperature were significantly shorter than that of pristine SnS2. The experimental results showed that 0.5 wt% was the optimum doping concentration of Au particles for the SnS2 nanosheet materials.
Response/recovery time is an important performance parameter of a gas sensor. Figure 9 is a comparison diagram of the response curves of the Au/SnS2-0.5 composite material and pristine SnS2 nanosheet material to 4 ppm NO2. It can be clearly observed from the figure that the response value (3.94) of the Au/SnS2-0.5 composite to 4 ppm NO2 was higher than that of pristine SnS2 nanosheet (2.39). The decoration of Au nanoparticles not only significantly improved the response value of SnS2, but also greatly shortened the response/recovery time of SnS2 nanomaterials to NO2. Compared with pristine SnS2, the response/recovery time of Au/SnS2-0.5 was 42 s/127 s, which was reduced 5.0/4.1 times shorter than that of the pristine SnS2 nanosheet material (220 s/520 s). This demonstrates that Au-decoration could enhance the sensing performance of SnS2 and effectively promote the adsorption and desorption process of NO2 molecules on the surface of SnS2 nanomaterials.
The response curve of the Au/SnS2-0.5-based sensor from 0.25 ppm to 8 ppm NO2 at 120 °C is shown in Figure 10a. The sensor showed a quick response to NO2. The resistance of the sensor gradually returned to the initial when the NO2 was terminated. The Au/SnS2 composite sensor had excellent response and recovery characteristics in the range of 0.25 ppm to 8 ppm NO2. Meanwhile, with the increase in the NO2 concentration, the response of the Au/SnS2-0.5 composite material increased. The response value of the Au/SnS2-0.5 composite material under different concentrations of NO2 was fitted, as shown in Figure 10b. The correlation curve between the response value of the Au/SnS2 composite material and NO2 was approximately linear. Furthermore, the theoretical detection limit (LOD) of the Au/SnS2-0.5-based sensor can be estimated by Equations (3) and (4), according to a signal/noise not less than 3 [35,36].
LOD = 3 × RMS noise / slope
RMS noise = i = 1 n ( R i R ¯ ) 2 / N
where R i is the response measured experimentally before NO2 exposure; R ¯ is the average response value; and N is the number of the data point. The theoretical LOD was thus calculated to be approximately 50 ppb using the slope of the fitted linearity of the response versus NO2 concentration and the root-mean-square (RMSnoise) deviation at the baseline. This shows that the decorated Au nanoparticles can improve the sensing performance of the SnS2-based sensor for the detection of NO2.
Figure 11 shows the response of the Au/SnS2-0.5 composite to different interfering gases of 5 ppm at the optimal working temperature (120 °C), in which the interfering gases include NH3, acetone, toluene, benzene, methanol, ethanol, and formaldehyde. The experimental results showed that the Au/SnS2-0.5 composite exhibited the highest response to NO2 gas. Although the SnS2-based sensor had a certain response to NH3, the response was far less than that of NO2. This is perhaps due to the fact that SnS2 has a tendency to adsorb N with a unique lone-pair polarity in the gas molecules, which makes NO2 gas molecules react with the material surface at even lower temperatures, while other interfering gases need higher energy to undergo adsorption and desorption with the material surface. Therefore, the SnS2 showed extremely low sensitivity to other interfering gases at a lower working temperature, showing excellent selectivity to NO2. This excellent selectivity remained unchanged with the decoration of Au nanoparticles.
Figure 12a reveals the signal reproducibility of the Au/SnS2-0.5-based sensor to 4 ppm NO2 at 120 °C. It shows that the resistance of the Au/SnS2-0.5-based sensor had no obvious change after five consecutive measurements, and the average response value was approximately 4.0, which indicates that the Au/SnS2-0.5 composite had superior signal repeatability. Figure 12b shows the long-term stability of the Au/SnS2-0.5-based sensor for 4 ppm NO2 at 120 °C. After 40 days of continuous gas sensing measurements, the response value of the sensor to NO2 was stable and basically remained unchanged. The deviation of the response of the sensor was calculated to be less than 4%, which indicates that the SnS2 nanosheets decorated with Au nanoparticles have brilliant stability. Table 1 shows the performance comparison between the SnS2 sensor in this work and the previously reported SnS2-based NO2 sensor. It can be seen that the Au/SnS2 sensor in this work showed relatively high response and fast response and recovery rate at the low temperature region. In addition, it should be mentioned that this Au/SnS2 material had a problem of high resistance. In future research work, further attempts will be made to reduce the resistance of Au/SnS2 materials by compositing Au/SnS2 materials with some sensitive materials with ultra-high room temperature conductivity (including carbon nanotubes, graphene, and WTe2, etc.) or using laser irradiation to meet the practical application requirements, which has been proven to be feasible in previous work [20].

3.3. Gas-Sensing Mechanism

As shown in Figure 13, due to the different work functions between the SnS2 semiconductor and Au nanoparticles, the work function of SnS2 was smaller than that of Au, and charge transfer occurred across the Au/SnS2 interface [37,38]. The charge was transferred from the conduction band of SnS2 to Au nanoparticles through the interface when Au nanoparticles were successfully deposited onto the SnS2 nanosheets until the Fermi level of SnS2 and Au reached final equilibrium. The Schottky contact was thus formed at the interface between the SnS2 and Au nanoparticles. Compared to the pristine SnS2, the Schottky contact formed at the interface between SnS2 and Au in the composite material narrowed the electronic conducting channel of the Au/SnS2, and the potential barrier difference of the Au/SnS2 changed more after NO2 adsorption and the redox reaction occurred on the surface of the sensing materials according to the below reactions:
O 2 gas + e O 2 ads
O 2 ads + e 2 O ads
NO 2 gas + e NO 2 ads
NO 2 gas + O 2 ads + 2 e NO 2 ads + 2 O ads
NO 2 gas + O ads + 2 e NO 2 ads + O ads 2
NO 2 ads + 2 O ads + e NO gas + 1 2 O 2 gas + 2 O ads 2
This led to the larger change in the resistance of the Au/SnS2-based sensor.
The small-sized Au nanoparticles aggregated together when more Au above 1 wt% was decorated in the SnS2 matrix, resulting in the decrease in active sites between the SnS2 nanosheets and Au nanoparticles. The reduction in the catalytic efficiency of noble metals led to the deterioration of the gas sensing performance of the sensor. In addition, the inherent characteristics of Au nanoparticles allows it to play a dual role in the gas sensing process of the composite material. During the sensor’s transducing process, Au nanoparticles can not only serve as electron donors to increase the number of carriers in the reaction between the composite material surface and the gas molecules, but they can also serve as electron acceptors to effectively separate the carriers generated in the semiconductor, prolonging the life of electron-hole and shortening the response/recovery time of the sensor.
The noble metal Au not only has the function of “electron sensitization”, but also has the function of “chemical catalysis” [39,40]. Compared to the pristine SnS2 nanosheets, Au nanoparticles decorated on the surface of the Au/SnS2 composites were also equivalent to the active sites of the gas sensing reaction. Given the “spillover effect” of noble metals, gas molecules adsorbed on the surface of the Au nanoparticles will further “spillover” to the surface of the SnS2 nanosheets, which leads to an increase in the number of active sites on the surface of Au/SnS2 composite materials [41,42]. Furthermore, Au nanoparticles can effectively promote the ionization of oxygen molecules and reduce the activation energy of the gas molecules, which makes realization of the dynamic equilibrium of the transducing reaction at a lower working temperature. As a result, the response/recovery time of the Au/SnS2 composites decrease and the performance of the sensor is improved.

4. Conclusions

The Au/SnS2 nanocomposites were successfully prepared by the hydrothermal and in situ reduction methods. The enhancement in the sensing performance of two dimensional structured SnS2 nanosheets by the decoration of Au was explored. It was found that the particle morphology and content of Au had a significant influence on the gas sensing performance of the Au/SnS2 composites. The Au/SnS2-0.5 sample with the optimum gold content could significantly improve the sensitivity, response, and recovery rate of the SnS2-based material to a low concentration NO2 gas at a low temperature region. At 120 °C, the response of the Au/SnS2-0.5 nanocomposite to 4 ppm NO2 gas was 3.94, which was significantly higher than that of the original SnS2 material (2.39). Moreover, the modification of Au nanoparticles could also significantly optimize the response and recovery characteristics of the SnS2-based sensor. The response/recovery time of the Au/SnS2-0.5-based sensor to 4 ppm NO2 gas at 120 °C was shortened to 42 s/127 s. The sensor also presented a favorable long-term stability with a deviation in the response of less than 4% for 40 days, and a brilliant selectivity to several possible interferents such as NH3, acetone, toluene, benzene, methanol, ethanol, and formaldehyde. The enhanced sensing performance could be attributed to the synergistic effects from the “electronic sensitization” and “chemical sensitization” furnished by Au nanoparticles introduced in the SnS2 matrix, which makes it a promising candidate material for preparing high-performance NO2 gas sensors.

Author Contributions

Conceptualization, investigation, data curation, formal analysis, software, visualization, original draft preparation, D.G.; Investigation, data curation, visualization, original draft preparation, W.L.; Investigation, data curation, J.W.; Methodology, conceptualization, formal analysis, J.Y.; Resources, writing—review & editing, J.Z.; Supervision, conceptualization, writing—review & editing, B.H.; Resources, software, M.N.R.; Resources, funding acquisition, X.L. All authors have read and agreed to the published version of the manuscript.

Funding

The authors are thankful to the National Natural Science Foundation of China (Nos. 61971085, 62111530055, 61874018), the National Key R&D Program of China (No. 2021YFB3201302), the Russian Foundation for Basic Research (No. 21-53-53018) and the Fundamental Research Funds for the Central Universities of China (No: DUT19RC(3)054).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Saboor, F.H.; Ueda, T.; Kamada, K.; Hyodo, T.; Mortazavi, Y.; Khodadadi, A.A.; Shimizu, Y. Enhanced NO2 gas sensing performance of bare and Pd-loaded SnO2 thick film sensors under UV-light irradiation at room temperature. Sens. Actuators B Chem. 2016, 223, 429–439. [Google Scholar] [CrossRef]
  2. Dai, Z.; Lee, C.S.; Tian, Y.; Kim, I.D.; Lee, J.H. Highly reversible switching from P- to N-type NO2 sensing in a monolayer Fe2O3 inverse opal film and the associated P-N transition phase diagram. J. Mater. Chem. A 2015, 3, 3372–3381. [Google Scholar] [CrossRef]
  3. Huang, Y.; Jiao, W.; Chu, Z.; Ding, G.; Yan, M.; Zhong, X.; Wang, R. Ultrasensitive room temperature ppb-level NO2 gas sensors based on SnS2/rGO nanohybrids with P-N transition and optoelectronic visible light enhancement performance. J. Mater. Chem. C 2019, 7, 8616–8625. [Google Scholar] [CrossRef]
  4. Li, K.; Luo, Y.; Liu, B.; Gao, L.; Duan, G. High-performance NO2-gas sensing of ultrasmall ZnFe2O4 nanoparticles based on surface charge transfer. J. Mater. Chem. A 2019, 7, 5539–5551. [Google Scholar] [CrossRef]
  5. Liu, W.; Gu, D.; Li, X. Detection of Ppb-level NO2 using mesoporous ZnSe/SnO2 core-shell microspheres based chemical sensors. Sens. Actuators B Chem. 2020, 320, 128365. [Google Scholar] [CrossRef]
  6. Kim, J.S.; Yoon, J.W.; Hong, Y.J.; Kang, Y.C.; Abdel-Hady, F.; Wazzan, A.A.; Lee, J.H. Highly sensitive and selective detection of ppb-level NO2 using multi-shelled WO3 yolk-shell spheres. Sens. Actuators B Chem. 2016, 229, 561–569. [Google Scholar] [CrossRef]
  7. Moon, H.G.; Jung, Y.; Han, S.D.; Shim, Y.S.; Shin, B.; Lee, T.; Kim, J.S.; Lee, S.; Jun, S.C.; Park, H.H.; et al. Chemiresistive Electronic Nose toward Detection of Biomarkers in Exhaled Breath. ACS Appl. Mater. Interfaces 2016, 8, 20969–20976. [Google Scholar] [CrossRef]
  8. Zhao, T.; Fu, X.; Cui, X.; Lian, G.; Liu, Y.; Song, S.; Wang, Q.; Wang, K.; Cui, D. An in-situ surface modification route for realizing the synergetic effect in P3HT-SnO2 composite sensor and strikingly improving its sensing performance. Sens. Actuators B Chem. 2017, 241, 1210–1217. [Google Scholar] [CrossRef]
  9. Koo, W.T.; Choi, S.J.; Kim, N.H.; Jang, J.S.; Kim, I.D. Catalyst-decorated hollow WO3 nanotubes using layer-by-layer self-assembly on polymeric nanofiber templates and their application in exhaled breath sensor. Sens. Actuators B Chem. 2016, 223, 301–310. [Google Scholar] [CrossRef]
  10. Lee, S.W.; Lee, W.; Hong, Y.; Lee, G.; Yoon, D.S. Recent advances in carbon material-based NO2 gas sensors. Sens. Actuators B Chem. 2018, 255, 1788–1804. [Google Scholar] [CrossRef]
  11. Liu, L.; Ikram, M.; Ma, L.; Zhang, X.; Lv, H.; Ullah, M.; Khan, M.; Yu, H.; Shi, K. Edge-exposed MoS2 nanospheres assembled with SnS2 nanosheet to boost NO2 gas sensing at room temperature. J. Hazard. Mater. 2020, 393, 122325. [Google Scholar] [CrossRef] [PubMed]
  12. Kim, Y.H.; Phan, D.T.; Ahn, S.; Nam, K.H.; Park, C.M.; Jeon, K.J. Two-dimensional SnS2 materials as high-performance NO2 sensors with fast response and high sensitivity. Sens. Actuators B Chem. 2018, 255, 616–621. [Google Scholar] [CrossRef]
  13. Hermawan, A.; Septiani, N.L.W.; Taufik, A.; Taufik, A.; Yuliarto, B.; Yin, S. Advanced Strategies to Improve Performances of Molybdenum-Based Gas Sensors. Nano-Micro Lett. 2021, 13, 1–46. [Google Scholar] [CrossRef] [PubMed]
  14. Xu, T.; Liu, Y.; Pei, Y.; Chen, Y.; Jiang, Z.; Shi, Z.; Xu, J.; Wu, D.; Tian, Y.; Li, X. The ultra-high NO2 response of ultra-thin WS2 nanosheets synthesized by hydrothermal and calcination processes. Sens. Actuators B Chem. 2018, 259, 789–796. [Google Scholar] [CrossRef]
  15. Cheng, M.; Wu, Z.; Liu, G.; Zhao, L.; Gao, Y.; Zhang, B.; Liu, F.; Yan, X.; Liang, X.; Sun, P.; et al. Highly sensitive sensors based on quasi-2D rGO/SnS2 hybrid for rapid detection of NO2 gas. Sens. Actuators B Chem. 2019, 291, 216–225. [Google Scholar] [CrossRef]
  16. Shafiei, M.; Bradford, J.; Khan, H.; Piloto, C.; Wlodarski, W.; Li, Y.; Motta, N. Low-operating temperature NO2 gas sensors based on hybrid two-dimensional SnS2-reduced graphene oxide. Appl. Surf. Sci. 2018, 462, 330–336. [Google Scholar] [CrossRef]
  17. Huang, Y.; Jiao, W.; Chu, Z.; Wang, S.; Chen, L.; Nie, X.; Wang, R.; He, X. High Sensitivity, Humidity-Independent, Flexible NO2 and NH3 Gas Sensors Based on SnS2 Hybrid Functional Graphene Ink. ACS Appl. Mater. Interfaces 2020, 12, 997–1004. [Google Scholar] [CrossRef]
  18. Ou, J.Z.; Ge, W.; Carey, B.; Daeneke, T.; Rotbart, A.; Shan, W.; Wang, Y.; Fu, Z.; Chrimes, A.F.; Wlodarski, W.; et al. Physisorption-Based Charge Transfer in Two-Dimensional SnS2 for Selective and Reversible NO2 Gas Sensing. ACS Nano 2015, 9, 10313–10323. [Google Scholar] [CrossRef]
  19. Liu, D.; Tang, Z.; Zhang, Z. Nanoplates-assembled SnS2 nanoflowers for ultrasensitive ppb-level NO2 detection. Sens. Actuators B Chem. 2018, 273, 473–479. [Google Scholar] [CrossRef]
  20. Gu, D.; Wang, X.; Liu, W.; Li, X.; Lin, S.; Wang, J.; Rumyantseva, M.N.; Gaskov, A.M.; Akbar, S.A. Visible-light activated room temperature NO2 sensing of SnS2 nanosheets based chemiresistive sensors. Sens. Actuators B Chem. 2020, 305, 127455. [Google Scholar] [CrossRef]
  21. Kondalkar, V.V.; Duy, L.T.; Seo, H.; Lee, K. Nanohybrids of Pt-Functionalized Al2O3/ZnO Core-Shell Nanorods for High-Performance MEMS-Based Acetylene Gas Sensor. ACS Appl. Mater. Interfaces 2019, 11, 25891–25900. [Google Scholar] [CrossRef] [PubMed]
  22. Sankar, G.R.; Navaneethan, M.; Patil, V.L.; Ponnusamy, S.; Muthamizhchelvan, C.; Kawasaki, S.; Patil, P.S.; Hayakawa, Y. Sensitivity enhancement of ammonia gas sensor based on Ag/ZnO flower and nanoellipsoids at low temperature. Sens. Actuators B Chem. 2018, 255, 672–683. [Google Scholar]
  23. Li, F.; Guo, S.; Shen, J.; Shen, L.; Sun, D.; Wang, B.; Chen, Y.; Ruan, S. Xylene gas sensor based on Au-loaded WO3·H2O nanocubes with enhanced sensing performance. Sens. Actuators B Chem. 2017, 238, 364–373. [Google Scholar] [CrossRef]
  24. Walker, J.M.; Akbar, S.A.; Morris, P.A. Synergistic effects in gas sensing semiconducting oxide nano-heterostructures: A review. Sens. Actuators B Chem. 2019, 286, 624–640. [Google Scholar] [CrossRef]
  25. Zhang, D.; Yang, Z.; Li, P.; Pang, M.; Xue, Q. Flexible self-powered high-performance ammonia sensor based on Au-decorated MoSe2 nanoflowers driven by single layer MoS2-flake piezoelectric nanogenerator. Nano Energy 2019, 65, 103974. [Google Scholar] [CrossRef]
  26. Gu, D.; Li, X.; Zhao, Y.; Wang, J. Enhanced NO2 sensing of SnO2/SnS2 heterojunction based sensor. Sens. Actuators B Chem. 2017, 244, 67–76. [Google Scholar] [CrossRef]
  27. Hao, J.; Zhang, D.; Sun, Q.; Zheng, S.; Sun, J.; Wang, Y. Hierarchical SnS2/SnO2 nanoheterojunctions with increased active-sites and charge transfer for ultrasensitive NO2 detection. Nanoscale 2018, 10, 7210–7217. [Google Scholar] [CrossRef]
  28. Wu, J.; Wu, Z.; Ding, H.; Wei, Y.; Huang, W.; Yang, X.; Li, Z.; Qiu, L.; Wang, X. Flexible, 3D SnS2/Reduced graphene oxide heterostructured NO2 sensor. Sens. Actuators B Chem. 2020, 305, 127445. [Google Scholar] [CrossRef]
  29. Yu, J.; Xu, C.Y.; Ma, F.X.; Hu, S.P.; Zhang, Y.W.; Zhen, L. Monodisperse SnS2 nanosheets for high-performance photocatalytic hydrogen generation. ACS Appl. Mater. Interfaces 2014, 6, 22370–22377. [Google Scholar] [CrossRef]
  30. Zhang, D.; Zong, X.; Wu, Z.; Zhang, Y. Hierarchical Self-Assembled SnS2 Nanoflower/Zn2SnO4 Hollow Sphere Nanohybrid for Humidity-Sensing Applications. ACS Appl. Mater. Interfaces 2018, 10, 32631–32639. [Google Scholar] [CrossRef]
  31. Feng, Z.; Ma, Y.; Natarajan, V.; Zhao, Q.; Ma, X.; Zhan, J. In-situ generation of highly dispersed Au nanoparticles on porous ZnO nanoplates via ion exchange from hydrozincite for VOCs gas sensing. Sens. Actuators B Chem. 2018, 255, 884–890. [Google Scholar] [CrossRef]
  32. Li, G.; Cheng, Z.; Xiang, Q.; Yan, L.; Wang, X.; Xu, J. Bimetal PdAu decorated SnO2 nanosheets based gas sensor with temperature-dependent dual selectivity for detecting formaldehyde and acetone. Sens. Actuators B Chem. 2019, 283, 590–601. [Google Scholar] [CrossRef]
  33. Suchomel, P.; Kvitek, L.; Prucek, R.; Panacek, A.; Halder, A.; Vajda, S.; Zboril, R. Simple size-controlled synthesis of Au nanoparticles and their size-dependent catalytic activity. Sci. Rep. 2018, 8, 4589. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Najafishirtari, S.; Guardia, P.; Scarpellini, A.; Prato, M.; Marras, S.; Manna, L.; Colombo, M. The effect of Au domain size on the CO oxidation catalytic activity of colloidal Au-FeOx dumbbell-like heterodimers. J. Catal. 2016, 338, 115–123. [Google Scholar] [CrossRef]
  35. Wu, J.; Feng, S.; Li, Z.; Tao, K.; Chu, J.; Miao, J.; Norford, L.K. Boosted sensitivity of graphene gas sensor via nanoporous thin film structures. Sens. Actuators B Chem. 2018, 255, 1805–1813. [Google Scholar] [CrossRef]
  36. Li, J.; Lu, Y.; Ye, Q.; Cinke, M.; Han, J.; Meyyappan, M. Carbon nanotube sensors for gas and organic vapor detection. Nano Lett. 2003, 3, 929–933. [Google Scholar] [CrossRef]
  37. Nakate, U.T.; Bulakhe, R.N.; Lokhande, C.D.; Kale, S.N. Au sensitized ZnO nanorods for enhanced liquefied petroleum gas sensing properties. Appl. Surf. Sci. 2016, 371, 224–230. [Google Scholar] [CrossRef]
  38. Wang, X.; Wang, J. Effects of Pt and Au adsorption on the gas sensing performance of SnS2 monolayers: A DFT study. Mater. Sci. Semicond. Process. 2021, 121, 105416. [Google Scholar] [CrossRef]
  39. Wang, F.; Wong, R.J.; Ho, J.H.; Jiang, Y.; Amal, R. Sensitization of Pt/TiO2 Using Plasmonic Au Nanoparticles for Hydrogen Evolution under Visible-Light Irradiation. ACS Appl. Mater. Interfaces 2017, 9, 30575–30582. [Google Scholar] [CrossRef]
  40. Yang, X.; Fu, H.; Tian, Y.; Xie, Q.; Xiong, S.; Han, D.; Zhang, H.; An, X. Au decorated In2O3 hollow nanospheres: A novel sensing material toward amine. Sens. Actuators B Chem. 2019, 296, 126696. [Google Scholar] [CrossRef]
  41. Li, F.; Li, C.; Zhu, L.; Guo, W.; Shen, L.; Wen, S.; Ruan, S. Enhanced toluene sensing performance of gold-functionalized WO3·H2O nanosheets. Sens. Actuators B Chem. 2016, 223, 761–767. [Google Scholar] [CrossRef]
  42. Xu, Y.; Ma, T.; Zheng, L.; Sun, L.; Liu, X.; Zhao, Y.; Zhang, J. Rational design of Au/Co3O4-functionalized W18O49 hollow heterostructures with high sensitivity and ultralow limit for triethylamine detection. Sens. Actuators B Chem. 2019, 284, 202–212. [Google Scholar] [CrossRef]
Figure 1. Schematic view of modifying Au nanoparticles on SnS2 nanosheets.
Figure 1. Schematic view of modifying Au nanoparticles on SnS2 nanosheets.
Chemosensors 10 00165 g001
Figure 2. Structure diagram of the Au interdigital electrodes.
Figure 2. Structure diagram of the Au interdigital electrodes.
Chemosensors 10 00165 g002
Figure 3. (a) XRD patterns of the prepared samples and (b) partial enlarged view of the XRD.
Figure 3. (a) XRD patterns of the prepared samples and (b) partial enlarged view of the XRD.
Chemosensors 10 00165 g003
Figure 4. SEM images of (a) SnS2, (b) Au/SnS2-0.5, (c) Au/SnS2-1.5, and (d) EDS elemental estimation of the Au/SnS2-0.5 sample.
Figure 4. SEM images of (a) SnS2, (b) Au/SnS2-0.5, (c) Au/SnS2-1.5, and (d) EDS elemental estimation of the Au/SnS2-0.5 sample.
Chemosensors 10 00165 g004
Figure 5. (a) TEM image of SnS2, (b) HRTEM image of SnS2, (c) TEM image of Au/SnS2-0.5, and (d) HRTEM image of Au/SnS2-0.5.
Figure 5. (a) TEM image of SnS2, (b) HRTEM image of SnS2, (c) TEM image of Au/SnS2-0.5, and (d) HRTEM image of Au/SnS2-0.5.
Chemosensors 10 00165 g005
Figure 6. XPS spectra of SnS2 and Au/SnS2-0.5: (a) full survey spectra of samples, (b) O 1s region, (c) Sn 3d region, (d) S 2p region, and (e) Au 4f region.
Figure 6. XPS spectra of SnS2 and Au/SnS2-0.5: (a) full survey spectra of samples, (b) O 1s region, (c) Sn 3d region, (d) S 2p region, and (e) Au 4f region.
Chemosensors 10 00165 g006
Figure 7. The response of as-fabricated sensors to 4 ppm NO2 in the temperature range of 100 °C to 180 °C.
Figure 7. The response of as-fabricated sensors to 4 ppm NO2 in the temperature range of 100 °C to 180 °C.
Chemosensors 10 00165 g007
Figure 8. (a) The resistance and (b) the response curves of SnS2, Au/SnS2-0.1, Au/SnS2-0.3, Au/SnS2-0.5, Au/SnS2-1, and Au/SnS2-1.5-based sensors to 4 ppm NO2 at 120 °C.
Figure 8. (a) The resistance and (b) the response curves of SnS2, Au/SnS2-0.1, Au/SnS2-0.3, Au/SnS2-0.5, Au/SnS2-1, and Au/SnS2-1.5-based sensors to 4 ppm NO2 at 120 °C.
Chemosensors 10 00165 g008
Figure 9. The response of SnS2 and Au/SnS2-0.5-based sensors to 4 ppm NO2 at 120 °C, and the inset images are the response/recover time of the samples.
Figure 9. The response of SnS2 and Au/SnS2-0.5-based sensors to 4 ppm NO2 at 120 °C, and the inset images are the response/recover time of the samples.
Chemosensors 10 00165 g009
Figure 10. (a) The resistance curve of the Au/SnS2-0.5-based sensor to 0.25 ppm–8 ppm NO2 at 120 °C. (b) The fitting curve of the response of the sensor with NO2 concentrations.
Figure 10. (a) The resistance curve of the Au/SnS2-0.5-based sensor to 0.25 ppm–8 ppm NO2 at 120 °C. (b) The fitting curve of the response of the sensor with NO2 concentrations.
Chemosensors 10 00165 g010
Figure 11. The selectivity of the Au/SnS2-0.5-based sensor to 4 ppm NO2 and several possible interfering gas species.
Figure 11. The selectivity of the Au/SnS2-0.5-based sensor to 4 ppm NO2 and several possible interfering gas species.
Chemosensors 10 00165 g011
Figure 12. (a) The repeatability of the Au/SnS2-0.5-based sensor to 4 ppm NO2 at 120 °C. (b) The long-term stability of the sensor.
Figure 12. (a) The repeatability of the Au/SnS2-0.5-based sensor to 4 ppm NO2 at 120 °C. (b) The long-term stability of the sensor.
Chemosensors 10 00165 g012
Figure 13. Schematic diagram of the charge transfer between the SnS2 nanosheet and Au nanoparticles.
Figure 13. Schematic diagram of the charge transfer between the SnS2 nanosheet and Au nanoparticles.
Chemosensors 10 00165 g013
Table 1. A comparison of the NO2 sensing performance of SnS2-based sensors.
Table 1. A comparison of the NO2 sensing performance of SnS2-based sensors.
Sensing MaterialTarget GasConcentration (ppm)Temperature (°C)ResponseResponse/Recovery TimeRef.
SnS2 nanoflowersNO20.11205.7850 s/1050 s[19]
SnS2-nanosheetsNO2102502.494 s/40 s[12]
SnS2-nanosheetsNO2101204.7120 s/170 s[18]
SnO2/SnS2NO28805.3159 s/297 s[26]
Au/SnS2-0.5NO241203.9442 s/127 sThis work
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Gu, D.; Liu, W.; Wang, J.; Yu, J.; Zhang, J.; Huang, B.; Rumyantseva, M.N.; Li, X. Au Functionalized SnS2 Nanosheets Based Chemiresistive NO2 Sensors. Chemosensors 2022, 10, 165. https://doi.org/10.3390/chemosensors10050165

AMA Style

Gu D, Liu W, Wang J, Yu J, Zhang J, Huang B, Rumyantseva MN, Li X. Au Functionalized SnS2 Nanosheets Based Chemiresistive NO2 Sensors. Chemosensors. 2022; 10(5):165. https://doi.org/10.3390/chemosensors10050165

Chicago/Turabian Style

Gu, Ding, Wei Liu, Jing Wang, Jun Yu, Jianwei Zhang, Baoyu Huang, Marina N. Rumyantseva, and Xiaogan Li. 2022. "Au Functionalized SnS2 Nanosheets Based Chemiresistive NO2 Sensors" Chemosensors 10, no. 5: 165. https://doi.org/10.3390/chemosensors10050165

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