Solution-processed NO2 gas sensor based on poly(3-hexylthiophene)-doped PbS quantum dots operable at room temperature

Characteristics of the synthesized PbS QDsAbsorbance, X-ray diffraction (XRD), and transmission electron microscopy (TEM) analyses of the synthesized QDs were conducted to evaluate their characteristics and successful synthesis. The obtained results are presented in Fig. 4. The absorbance spectra of the synthesized QDs presented in Fig. 4a reveals peaks at 905, 1210, and 1392 nm, indicating a high absorbance in the infrared region. The wavelength band of QDs changes depending on the particle size owing to quantum confinement. Based on quantum confinement, when the size of a particle is smaller than the Bohr radius, the instability increases with decreasing particle size, whereas the bandgap of the particle increases. Based on Eq. (1), the bandgap and wavelength band are inversely proportional. Because the band gap and particle size are inversely proportional, the wavelength band and size of the QDs are directly proportional.Fig. 4Measured results: (a) Absorbance properties; (b) XRD patterns; (c), (d), and (e) TEM images of the synthesized QDs.$$Eg \left(eV\right)=\frac{1240}{\lambda }$$
(1)
Furthermore, to determine the crystal structure and particle size of PbS QDs, XRD (D8-DISCOVER, Bruker AXS) analysis was performed, and the results are displayed in Fig. 4b. For XRD analysis, the PbS QD sample solutions (10 mg/mL) were spin-coated on a 10 × 10–mm2 glass substrate at 1000 rpm for 30 s, followed by annealing at 110 °C for 30 min. The lattice peaks in the patterns of all the synthesized PbS QDs were in conformance with those reported for PbS QDs in literature, and the peak intensity of (220) reflection increased with increasing particle size26. In addition, the peaks of P3HT were confirmed at approximately 23.6° and 37° only in P3HT-doped QDs (JCPDS File no-4S-2040)27. The crystallite size of PbS QDs (1210 nm) was calculated from the XRD results using the Scherrer equation28.$$ D_{hkl} \left( {{\text{nm}}} \right) = ({\text{K}}*\lambda )/\beta \cos \theta . $$
(2)
Here, K is the Scherrer constant, λ is the wavelength of X-ray, b is the maximum half width in radians, and q is the half-value between the incident angle and scattered X-ray wavelength vector. Gaussian fitting was performed to measure the theta value and full width at half maximum (FWHM), and the peaks with a precise fit were selected. K is 0.94 for a spherical crystallite possessing cubic symmetry with q = 15.1, FWHM = 1.74, and λ = 0.15418 nm, which is the wavelength of Cu K-alpha X-rays. The size of PbS QDs calculated using the Scherrer equation was 4.94 nm. Additionally, the TEM images presented in Fig. 4c–e reveal that the particle sizes of the synthesized PbS QDs with the absorption wavelengths of 905, 1210, and 1392 nm are 2.9, 4.8, and 5.2 nm, respectively. The lattice space and selected area electron diffraction (SAED) of PbS QDs were analyzed through TEM image, and the lattice space and SAED of (111), (220), and (200) were confirmed. These results confirm the successful synthesis of QDs. In addition, the lattice space and SAED of (020) were confirmed only in P3HT-doped PbS QDs, which confirmed that they were P3HT doped.
Performance of the NO
2
gas sensor
To determine the optimal P3HT doping ratio, PbS QDs with an absorbance wavelength of 905 nm were used to evaluate the reactivity of various P3HT concentrations. Figure 5a–d illustrate the real-time responsivity of PbS QDs with the absorbance wavelength of 905 nm based on the P3HT doping ratio under the conditions of 100–20 ppm NO2 gas. Responsivity was calculated using Eq. (3).Fig. 5(a–d) Current, (e) response characteristics based on the P3HT ratio at a NO2 gas concentration of 100–20-ppm and (f) responses at 100-ppm NO2 gas.$$\text{Response}=\frac{{I}_{1} – {I}_{0}}{ {I}_{0}}=\frac{\Delta I}{ {I}_{0}}$$
(3)
Here, I0 is the initial current of the sensor measured in air and I1 is the current when the sensor is exposed to the target gas.The highest and lowest response values were obtained when the ratio of PbS QDs to P3HT was 1:1 and 1:2, respectively. Figure 5e depicts the reactivity as a function of the NO2 gas concentration for different P3HT doping ratios, and Fig. 5f presents the reactivity according on the P3HT doping concentration under the condition of 100 ppm NO2 gas. The results confirm that the response at the ratio of 1:1 is approximately 40 times higher than that at the ratio of 1:2, and therefore, the former is adopted as the optimal P3HT doping ratio. When the ratio is 1:1, a thin film is stably formed; PbS QDs are uniformly deposited; sufficient reaction occurs upon gas contact. However, when the ratio is 1:2, P3HT is highly dominant. Consequently, the thin film is produced in a gel form, and the number of PbS QDs in contact with NO2 gas decreases. Therefore, the QD thin films with absorbance at 905, 1210, and 1392 nm were fabricated at an optimized ratio of 1:1 for the subsequent experiments.Next, the reactivity was measured as a function of the NO2 gas concentration in real time for the QDs with various wavelengths, and the results are presented in Fig. 6a–d. The gas concentration was varied from 20 to 1 ppm in 5-ppm decrements, and both the reaction and recovery times were maintained at 10 min. The measurements at 20-ppm NO2 revealed that the device comprised 905-nm PbS QDs exhibited the highest response of 2.06, which is ~ 2.1 times higher than that obtained for 1210-nm PbS QDs possessing a responsivity of 0.98. The core size of PbS QDs with a wavelength of 905 nm is 2.9 nm, which is smaller than those of the QDs with the wavelengths of 1210 (core size = 4.8 nm) and 1392 nm (core size = 5.2 nm). Additionally, a repeatability test was conducted with the response and recovery times of 200 and 800 s, respectively, at the NO2 gas concentration of 20 ppm to confirm the reliability of the fabricated sensor (Fig. 6e). The error in the response measurements repeated for five times was within 1%. The size of the PbS QDs with a wavelength of 905 nm is 3.9 nm, and therefore, a uniform thin film with a large surface area can be obtained compared with that formed using large PbS QDs. The atomic force microscopy (AFM) results presented in Fig. 7 confirm that the thin film formed from 905 nm PbS QDs exhibits the smallest height deviation among the tested PbS QD thin films. Therefore, the area of reaction with gas is the largest for the 905-nm PbS QDs, and the quality of the thin film is high, resulting in stable electrical characteristics. Consequently, a higher reactivity is expected when the size of the synthesized QDs is smaller. However, the minimum limit of the wavelength range of PbS QDs is 905 nm, which renders the fabrication of small QDs impossible.Fig. 6(a–d) Current characteristics based on the size of PbS QDs at a NO2 gas concentration of 20–1 ppm, and (e) repeatability graph of the 2.9-nm PbS QD–based sensor at 25 °C.Fig. 7Surface AFM measurement results for the fabricated sensors based on the size of PbS QDs: (a) 2.9, (b) 4.8, and (c) 5.2 nm.To evaluate the low-concentration NO2 gas sensing characteristics, the fabricated devices were evaluated at NO2 gas concentrations of 1–400 ppb with a resolution of 200 ppb. Figure 8 reveals that the device fabricated using 905-nm QDs exhibits the highest reactivity. At a concentration of 400 ppb NO2, the response values of the device were 0.21, 0.19, and 0.13, respectively. Therefore, the NO2 gas detection limit of the fabricated device was confirmed as 400 ppb. Furthermore, the repeatability of low-concentration NO2 gas sensing was tested. From Fig. 8e at a concentration of 1 ppm, the error in the measurement of responsivity is within 1%, which is the same as that achieved at high concentrations. Metal oxide-based NO2 gas sensors have high sensitivity, but they operate at high temperatures and require high-temperature calcination processes or complicated semiconductor processes. The proposed sensor has the advantages of being easy to mass-produce through a solution process, not requiring a high-temperature calcination process or semiconductor process and being able to detect with good sensitivity at room temperature. As can be seen in Fig. 9, when heated to 100 °C, it has a four times higher response, which is higher than that of other studies. To compare the response with the reference sensors presented in the Table 1, the sensor response at 10 ppm in each paper was compared. Compared to not only the PbS QDs-based sensor but also the metal oxide-based sensor, the proposed sensor was confirmed to have the good response. In addition, the widest range of measured concentration characteristics was confirmed, and the LOD was also confirmed to be excellent at 70 ppb10,29,30,31,32,33,34,35,36. There are previous studies that show high sensitivity in a relatively small NO2 gas detection range of ppb level, and they are mainly studied for application in the atmospheric environment or respiratory/bio fields36. However, since the emission sources such as power plants, factories, and automobile exhaust gases emit more than several tens of ppm of NO2 gas due to fossil fuel combustion, such prior research has limitations in industrial use due to its small detection range. In addition, according to the regulations of the Occupational Safety and Health Administration (OSHA), if the workers at the industrial sites are exposed to NO2 gas above 5 ppm, it can be dangerous to the workers and requires warning and evacuation. Therefore, the sensor proposed in this study can not only secure the safety of the workers by detecting NO2 concentration in real time, but also improve air pollution problems through a direct monitoring system at NO2 emission sources such as power plants, factories, and automobile exhaust.Fig. 8(a–d) Current characteristics depending on the size of PbS QDs at the NO2 gas concentrations of 1–0.2 ppm, and (e) repeatability characteristics of the 2.9-nm PbS QD–based sensor at 25 °C.Fig. 9Response characteristics of the fabricated sensor according to temperature conditions (20–1 ppm of NO2 gas).Table 1 Recent reported researches of the NO2 gas sensor.Furthermore, to confirm the operation characteristics at high temperatures, the reactivity of the sensor toward 20–1-ppm NO2 gas was tested at 25 °C (room temperature), 50 °C, 100 °C, and 150 °C.As demonstrated in Fig. 9, there is negligible difference in the responsivity values obtained at 50 °C and room temperature which is considered erroneous. The response at 100 °C is approximately 4 times higher than that obtained at room temperature, and the responsivity of NO2 gas and PbS QDs increases with increasing temperature. This is because H2O evaporates at 100 °C, the number of sites that can react on the sensor surface increases, and physical oxygen adsorption occurs more easily on the surface, and the responsivity increases accordingly.Before NO2 exposure, oxygen molecules in the air attach to the surface of PbS QDs and capture electrons to form O2− below 100 °C, or O− between 100 and 150 °C creating a surface state similar to that of an acceptor33,37. The as-formed surface state causes the excitation of the electrons in the valence band and induces the formation of numerous holes in the p-type PbS QDs, creating an electron depletion layer near the surface.$$ {\text{O}}_{{{2}({\text{gas}})}} \to {\text{O}}_{{{2}({\text{ads}})}} $$
(4)
$$ {\text{O}}_{{{2}({\text{ads}})}} + {\text{ e}}^{ – } \to {\text{O}}_{{2({\text{ads}})}}^{ – } $$
(5)
$$ {\text{O}}_{{2({\text{ads}})}}^{ – } + {\text{ e}}^{ – } \to {\text{2O}}_{{({\text{ads}})}}^{ – } $$
(6)
When exposed to NO2 gas, the oxidizing gas acts as an electron acceptor, and a reaction occurs on the PbS surface according to the following equations.$$ {\text{NO}}_{{{2}({\text{gas}})}} + {\text{ e}}^{ – } \to {\text{NO}}_{{2({\text{ads}})}}^{ – } $$
(7)
$$ {\text{NO}}_{{{2}({\text{gas}})}} + {\text{ O}}_{{2({\text{ads}})}}^{ – } + {\text{ 2e}}^{ – } \to {\text{NO}}_{{2({\text{ads}})}}^{ – } + {\text{ 2O}}_{{({\text{ads}})}}^{ – } $$
(8)
$$ {\text{NO}}_{{2({\text{ads}})}}^{ – } + {\text{ 2O}}_{{({\text{ads}})}}^{ – } {\text{ + e}}^{ – } \to {\text{NO}}_{{{2}({\text{gas}})}} + {\text{ 2O}}_{{2({\text{ads}})}}^{ – } $$
(9)
NO2 molecules are directly adsorbed and acquire electrons from the PbS QD surface. Consequently, the hole concentration increases, whereas the depth of the electron depletion layer and height of the potential barrier decrease because of the oxidizing nature of NO2 and its substantially higher electron affinity than oxygen. Therefore, the sensor resistance decreases, and the current increases10,29,30,31,32,33,34,35,36,37,38. Additionally, owing to the high hole mobility of P3HT, the sensor response can be enhanced by improving the flow of charges18,19. The principle of NO2 gas detection is illustrated in Fig. 10.Fig. 10Mechanism of NO2 gas sensing by PbS QDs.At 150 °C, the response is reduced compared with that at 100 °C, which is considered to be because of the deterioration of the PbS QDs thin film. Therefore, the fabricated sensor can detect NO2 gas with high sensitivity even at temperatures up to 150 °C. Since most metal oxide-based gas sensors cannot react with oxygen in the air at low temperatures, they cannot form a site for reacting with the target gas in the material. Therefore, physical adsorption, which is a direct charge transfer between the sensing material and the target gas, is dominant. However, according to the reference, PbS QDs form a reaction site and have high reactivity due to relatively high binding energy to oxygen (0.5 eV) and NO2 (0.8 eV) even below 100 °C39. Because most metal oxide-based gas sensors require high temperatures above 300 °C for successful detection, PbS QDs-based gas sensors have the advantage of being able to operate at room temperature.One issue with this sensor has observed that the properties of the sensing material deteriorated at high temperatures. In this manuscript, experimental measurements were not conducted at temperatures above 200 °C because the PbS QDs thin film was observed to be damaged by high temperature. This phenomenon is a major obstacle to commercialization. One possible way to resolve this issue is to form a polymer film on the surface of the sensing material which served as protective layer from the heat40,41. It was observed that the porosity of film can allow the gas to react with the sensing material, and greatly buffers the damage caused by high temperatures. The new findings will be reported shortly.Additionally, to confirm the selectivity of the NO2 gas sensor fabricated using P3HT-doped PbS QDs, its reactivity to other gases was measured. The experiment was performed using 100 ppm of CO2, H2S, CO, CH4 and SO2 gases for 10 min, and the measured results were compared with those obtained with 100 ppm of NO2 gas. As demonstrated in Fig. 11, although certain reactivity toward other gases is observed, it is negligible at a level of up to 7% compared with that toward NO2 gas10,28,31. In addition, to confirm the long-term stability of the fabricated sensor, the responsivity to 100 ppm NO2 gas was measured at 30 cycles every 5 days for a month. As can be seen in Fig. 12, it was confirmed that stable characteristics were maintained within the 3% error range caused by the measurement system and the measurement environment. Furthermore, the response of the fabricated PbS QD–based NO2 gas sensor to humidity was measured at 25 °C. The relative humidity of the environment was set to 30%, 60%, and 90% with a stabilization time of 20 min, and thereafter, the measurement was performed using 20 ppm NO2. Humidity was controlled through the humidity generator of the MFC, and the humidity in the chamber was measured using a humidity sensor (RH-BTA, Vernier Software & technology). As the humidity increased, the resistance increased, whereas the initial current gradually decreased. Additionally, noise was generated owing to the presence of moisture. As presented in Fig. 13, the response values are 2.13, 1.89, and 1.79 at 30%, 60%, and 90% humidity, respectively, and the number of binding sites of NO2 decreased because of the presence of moisture. However, this difference can be mitigated through packaging, and the fabricated sensor can be stably operated irrespective of the changes in season and weather.Fig. 11Selectivity characteristics of the fabricated sensor according to the gas type.Fig. 12Long-term reliability test of the fabricated sensor.Fig. 13Changes in (a–c) current and (d) response of the fabricated NO2 gas sensor depending on humidity variations (20-ppm NO2; 25 °C).

Hot Topics

Related Articles