Next Article in Journal
Design and Characterization of ITO-Covered Resonant Nanopillars for Dual Optical and Electrochemical Sensing
Next Article in Special Issue
In Situ Growth of Dopamine on QCM for Humidity Detection
Previous Article in Journal
A New Chemosensor Based on a Luminescent Complex for the Investigation of Some Organophosphorus Pesticides in Environmental Samples
Previous Article in Special Issue
Influence of Positive Ion (Al3+, Sn4+, and Sb5+) Doping on the Basic Resistance and Sensing Performances of ZnO Nanoparticles Based Gas Sensors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Au/La Co-Modified In2O3 Nanospheres for Highly Sensitive Ethanol Gas Detection

1
School of Electrical and Computer Engineering, Jilin Jianzhu University, Changchun 130118, China
2
State Key Laboratory of Integrated Optoelectronics, College of Electronic Science and Engineering, Jilin University, Changchun 130012, China
*
Author to whom correspondence should be addressed.
Chemosensors 2022, 10(10), 392; https://doi.org/10.3390/chemosensors10100392
Submission received: 27 August 2022 / Revised: 21 September 2022 / Accepted: 21 September 2022 / Published: 24 September 2022

Abstract

:
In this paper, the gas-sensitive properties of co-doping the rare earth element La and noble metal Au in In2O3 nanospheres were investigated for ethanol detection. Through XRD and SEM characterization, the grain size of La-In2O3 and Au/La-In2O3 nanoparticles was smaller than that of pure In2O3. As expected, the smaller grain size sample has shown a higher response for ethanol vapor. Compared with the pure In2O3 nanoparticles, the 2 mol%Au/2 mol%La-In2O3 sample has shown better ethanol-sensing properties, including higher sensitivity (S = 381) and lower operating temperature (210 °C) for 100 ppm ethanol vapor. In addition, the Au/La-In2O3 sensor presented a fast response time (1 s). The enhancement mechanism of the ethanol response was discussed for Au/La-In2O3 nanoparticles. The obtained experimental results would provide a new road for designing higher response sensors.

1. Introduction

Ethanol (C2H5OH), a common volatile compound gas, is widely applied to industrial production, the food industries and the biomedical field. However, exposure to excess ethanol gas can bring about headaches, liver and kidney damage, central nervous system harm and breathing difficulties [1,2]. Moreover, many traffic accidents are induced by ethanol abuse at present, which has caused around 1.35 million people deaths per year globally. Hence, the quantitative investigation of ethanol has become a key role for protecting human health, ensuring food safety and avoiding drunk driving via breathing test [3,4].
Recently, a variety of ethanol sensors have been designed with different sensing principles [5,6,7], such as those based on a chemiresistive model [8], optical fiber [9] and electrical properties [10]. Among them, chemiresistive model sensors present lower power, fast response, compact size and high sensitivity, resulting in the best option for ethanol detection. The characteristics of the chemiresistive model sensor indicate that the resistance of the sensor changes with the target gas concentration variation because of the physical/chemical adsorption of the target gas. Different metal oxide semiconductor materials have been applied to the chemiresistive model ethanol sensor, such as ZnO, SnO2, TiO2, Fe2O3, In2O3 and so on [11,12,13,14,15,16]. Among these host materials, In2O3 is a wide bandgap (3.55–3.75 eV) n-type semiconductor, which is widely applied in gas sensors because of its simple preparation, good electrical conductivity, considerable chemical stability and environmentally friendly characters. However, the ethanol sensors designed with pure In2O3 as the base sensing materials have some disadvantages, such as lower sensing response, longer response/recovery time, poor selectivity and high operating temperature. Generally, the doping noble metal in In2O3 host induces a remarkable improvement of the gas sensing properties [17,18,19,20,21,22]. In addition, many papers have also confirmed that the lattice defect induced by the impurity doping was an effective road to increase the gas-sensing properties of n-type semiconductor materials. At present, the trivalent rare earth (RE) ions have been applied in many fields due to their optical, magnetic and electrochemical performances based on 4f-4f electronic transitions. Particularly, the prominent catalytic property and large oxygen ion mobility for RE materials contribute to their chemical sensing applications. Many RE ions doping gas-sensing materials have been reported for improving the gas-sensing properties of metal oxide semiconductors, such as Ce/SnO2, Eu/SnO2, Sm/SnO2 and Y/SnO2 [23,24,25,26]. Anand et al. prepared Dy3+-doped In2O3 gas sensors by the co-precipitation method with good sensitivity to 50 ppm ethanol at 300 °C and good response values to 50 ppm acetone at 350 °C [27]. Duan et al. have synthesized Ho-doped In2O3 gas-sensitive materials with a response value of up to 60 ppm for 100 ppm ethanol at 240 °C, which is three times higher than that of pure In2O3. In addition, the sensitivity of Ho-doped In2O3 to 200 ppb ethanol reached 2[S = Rg/Ra] [28]. Qin et al. prepared Er-doped In2O3 gas sensors by the carbon thermal reduction method and the Er doping significantly reduced the operating temperature of the sensors. Moreover, the response of the Er-doped In2O3 gas sensor for 100 ppm ethanol at 220 °C is 4.8[S = Rg/Ra], which is twice as high as that of the pure In2O3 gas sensor [29]. Therefore, designing an ethanol-sensing material through the codoping of noble metal and RE ions would become one of the feasible routes to enhancing the ethanol-sensing property.
In this paper, the La3+ ions and noble metal Au co-doping of In2O3 nanomaterials were prepared through the hydrothermal method. The resistive-type sensor based on La/Au-In2O3 nanomaterials was designed to detect ethanol. The ethanol-sensing performances (sensitivity, selectivity and response/recovery time) of the fabricated sensor were investigated. It has been found that the remarkable enhancement of the response for ethanol was obtained, which is attributed to La/Au codoping in In2O3. The possible sensing mechanism was demonstrated based on the change of semiconductor conductance.

2. Experiment

2.1. Synthesis of La/Au-In2O3 Nanomaterials

The pure In2O3, 2 mol%La-In2O3 and x mol%Au/2 mol%La-In2O3 (x = 1, 2, 3) nanomaterials were prepared through a simple hydrothermal method. The main raw materials of indium nitrate (In(NO3)3·4.5H2O), lanthanum nitrate (La(NO3)3·6H2O) and chloroauric acid (HAuCl4) were purchased from Sinopharm Chemical Reagent Co. All raw materials in the experiments were of analytical grade and did not require further purification. The synthesis process of 2 mol% La:In2O3 sample was as follows. According to the calculated amount, a certain amount of In(NO3)3·4.5H2O, La(NO3)3·6H2O and sodium citrate were dissolved in deionized water. First, the solution(A) of La(NO3)3 was added to the solution(B) of In(NO3)3 with a parting funnel. Then 10 mL of sodium citrate solution(C) was added to the above-mixed solution(A + B) with a separatory funnel and stirring continued for 30 min with a magnetic stirrer. The above-mixed solution was transferred to a PTFE-lined stainless steel reactor (50 mL) and kept at 200 °C for 4 h. The obtained white precipitate was washed repeatedly with deionized water and ethanol. Then, it was dried at 70 °C for 10 h. Finally, the samples were calcined at 500 °C in oxygen for 3 h. The preparation of La/Au-In2O3 sample was compounded with La-In2O3, HAuCl4 and L-lysine. First, HAuCl4 solution and L-lysine solution were added to La-In2O3 suspension sequentially with a separatory funnel and stirred for 10 min, then treated with ultrasound for 20 min. The precipitate was centrifuged and collected, followed by washing several times with ethanol and deionized water, then dried in a blast dryer at 60 °C for 12 h. Finally, the prepared material was calcined at 300 °C for 30 min, after which the La/Au-In2O3 nanomaterials were obtained.

2.2. Characterization

The crystalline phase of the prepared sample was characterized by X-ray diffraction (XRD, Rigaku Ultima IV) under an operating voltage and current of 40 kV and 40 mA. The Cu-Ka radiation was used with the wavelength λ = 1.54056 Å. The morphology of the samples was characterized by scanning electron microscopy (SEM, Quanta 450).
The gas sensor was prepared through sensing materials and ceramic tubes with gold electrodes. First, the appropriate amount of sample was ground into a homogeneous paste. Then the material is uniformly coated on ceramic tubes. Finally, the sensor was aged for 5 days at 200 °C in air using a chromium–nickel coil as a heating resistance wire. In addition, the gas-sensitive performance of the sensor was tested using the CGS-8 (Beijing Elite Tech Co., Ltd., Beijing, China) intelligent gas sensing analysis system. The response of the gas sensor was usually defined as Ra/Rg, where Ra was the resistance of the sensor in air and Rg was the resistance of the sensor in the gas to be measured [30,31].

3. Results and Discussion

3.1. Structural and Morphology

The crystalline structures of pure In2O3, La-In2O3, and Au/La-In2O3 (Au = 1 mol%, 2 mol%, 3 mol%) samples were analyzed by XRD, and the results are shown in Figure 1. It can be seen that the main diffraction peaks of the five samples match well with hexagonal phase In2O3(JCPD#73-1809). However, three weak diffraction peaks at (2θ = 35.35°, 50.90°, 60.60°) are also presented, which belong to cubic phase In2O3. This phenomenon may be the result of the combined effect of the surfactant sodium citrate, hydrothermal temperature and time, which is consistent with the results observed by Feng Chen et al. [32,33]. In addition, no diffraction peaks of La2O3 and Au are detected because of their low doping amount. To further characterize the features of La/Au co-doped In2O3, the lattice parameters of samples are calculated. The Scherrer formula can be used to estimate the grain size [34]:
D = K λ / B cos θ
where K is the Scherrer constant (K = 0.943 for cubic particles); B is the half height and width of the diffraction peak; (λ = 1.54056 nm) is the x-ray wavelength; and θ is the Bragg diffraction angle. The grain size of pure, 2 mol%La, 1 mol%Au/2 mol%La, 2 mol%Au/2 mol%La and 3 mol%Au/2 mol%La doping In2O3 samples are listed in Table 1. It can be seen that the grain size of samples decreases significantly after doping with La and Au.
The SEM images of pure In2O3, 2 mol%La-In2O3 and 2 mol%Au/2 mol%La-In2O3 are shown in Figure 2. As seen in Figure 2a,b, the particles of pure In2O3, 2 mol%La-In2O3 present porous aggregate state. The size of particle should belong to nanoscale. While the morphology of 2 mol%Au/2 mol%La-In2O3 shows the uniformly sized dispersed spheres about 3 nm. The formation of spheres may be due to the addition of L-lysine as a dispersant and sonication.

3.2. Ethanol-Sensitive Properties

To investigate the ethanol-sensing performances of Au/La-In2O3 nanomaterials, the ethanol (100 ppm) sensing responses of pure In2O3, 2%La-In2O3, 1%Au/2%La-In2O3, 2%Au/2%La-In2O3 and 3%Au/2%La-In2O3 were tested at different temperatures from 190 °C to 260 °C in Figure 3. The sensitivities of 2%La-In2O3, 1%Au/2%La-In2O3, 2%Au/2%La-In2O3 and 3%Au/2%La-In2O3 sensors first increase rapidly, then decrease and reach maximum value at 210–230 °C, which reveal a “pyramid” shape. This can be explained as follows: at low temperature the adsorbed ethanol molecules are not activated enough to overcome the activation energy barrier to react with the adsorbed oxygen species, while at high temperatures the gas adsorption is too difficult to be adequately compensated for the increased surface reactivity. That is, at lower operating temperatures, the activation energy of ethanol vapor molecules is weak, and the ethanol molecules are probably not absorbed by sensing material, which results in the lower response. With the operating temperature increasing, the activation energy of ethanol molecules is enhanced. The reaction of ethanol molecules and adsorbed oxygen is increased, which leads to the gas response improvement. However, when the working temperature is over a certain value, the escaping rate of ethanol molecules from the sensing materials is enhanced, which leads to the weak reaction between ethanol molecules and oxygen molecules. For this reason, the response value begins to decrease. As shown in Figure 3, the optimal operating temperature of the 2%Au/2%La-In2O3 sensor is 210 °C with a response of 381. In addition, 2%La-In2O3 sensors give at 230 °C a response of 43. However, the optimal operating temperature of pure In2O3 is over 260 °C. In addition, the 2% Au/2%La-In2O3 sensor shows the higher sensitivity (381) at the lowest optimal operating temperature (210 °C) to 100 ppm ethanol. Loading Au and La onto In2O3 helps to lower the optimum working temperature, because the Au can decrease the activation energy barrier of the surface reaction due to its excellent catalytic property. It is noted that the decoration of Au and La can significantly enhance the response. Due to RE ions’ catalytic properties, the codoping La can increase the reaction rate. Amazingly, the Au doping can significantly increase the sensitivity, proving that noble metals play a catalytic role. The response of the La/Au-In2O3 sensing materials at first increases with increasing Au concentration, and reaches the maximum value for 2%Au/2%La-In2O3, then decreases. The 2%Au/2%La-In2O3 sensor has achieved a response of 381, about 14 times more than that of the pure In2O3 sensor and about 8.9 times that of the 2%La-In2O3 sensors. As mentioned above, the catalytic performance of Au will further increase the response of the La/In2O3. However, the overloading Au doping value will result in the decrease of the contact surface between the gas and the In2O3 host. This means that the response will decrease. This suggests that co-doping of RE ions and noble metals is an effective method for enhancing the semiconductor materials gas-sensing response.
The sensing response of 2%Au/2%La-In2O3 nanomaterials at 210 °C was measured. Figure 4a shows the four consecutive repeatability experiment results of 2%Au/2%La-In2O3 sensor toward 50 ppm. The results show that the sensor has a fast response time and good repeatability. The response–recovery times of 2%Au/2%La-In2O3 sensor are investigated toward 50 ppm in Figure 4b. The response–recovery times of the sensor reached 1/394 s. This indicates that the sensor has a super-fast response for the application.
In Figure 5a the response-recovery curves of the sensor for different ethanol concentrations are investigated from 1 to 100 ppm. The results show that the sensor has a fast and increasing response amplitude at increasing concentrations of ethanol vapor, and then returns to the initial state after re-exposure to air, with a lower detection limit of 1 ppm (response value of 1.48). The linear fitting relationship between the response values of the Au/La-In2O3 sensor and different ethanol concentration is shown in Figure 5b. The response of the sensor increases linearly with increasing ethanol concentration in the range of 20–80 ppm. The relation is Y = 2.89X − 26.89, where the correlation coefficient is R2 = 0.9785. It indicates that the R2 is about 1 with better linear fitting. Therefore, the Au/La-In2O3 sensor can be applied to the quantitative detection of ethanol from 20 to 80 ppm.
The sensing response of the 2%Au/2%La-In2O3, 2%La-In2O3 and pure In2O3 at optimal working temperature for different gases at 100 ppm, as shown in Figure 6. The target gases include ethanol, trimethylamine, methanol, SO2, xylene and NH3. The results show that the In2O3 sensor co-doped with Au and La has high selectivity for ethanol.

3.3. Mechanistic Analysis

As reported, the possible gas-sensing mechanism of the metal oxide semiconductor is discussed with the “gas adsorption and gas desorption reaction”. Once the sensing materials are exposed to—oxygen in the air, the physical and chemical adsorption process of the oxygen will occur on the surface of the sensor, which generates the anionic oxygen ( O 2 , O, O2−). The surface of In2O3 materials will lose electrons, which reduces the resistance of the sensor. When the sensor is exposed to ethanol, the reaction of the ethanol and adsorbed oxygen anions will occur at the surface of the sensor. The electrons are released to the sensing materials and the result is a decreased resistance for the gas-sensing application. The process can be described by following equation.
O2(gas)→O2(ads)
O2(ads) + e→O2(ads) (100 °C < T)
O2(ads)+ e→2O(ads) (100 °C < T < 300 °C)
O(ads) + e→O2−(ads) (300 °C > T)
C2H5OH(gas) + 3O2(ads)→2CO2 + 3H2O + 6e
C2H5OH(gas) + 6O(ads)→2CO2 + 3H2O + 6e
Through the process, the electrons would be released back to the In2O3. So, the resistance of the sensing materials begins to decrease. In this paper, the ethanol gas response of 2%Au/2%La-In2O3 shows the best sensing performance, while that of 2%La-In2O3 also has a better sensing property than pure In2O3. The possible reasons for the remarkable enhancement for ethanol detection could be described as follows:
(1) The response of gas sensors is greatly influenced by the grain size based on the previous reports. The smaller grain size of the sensing material will result in larger specific surface area, which generates higher gas sensitivity [35,36]. For 2%La-In2O3 materials, the average grain size has been calculated to be 12.17 nm, which is smaller than that of pure In2O3. Therefore, the response of the 2%La-In2O3 sensor is larger than that of the pure In2O3 sensor. In addition, the doping La3+ ions maybe replace the site of the In3+ ions (Figure 7). The XRD results indicate that the doping of La3+ leads to a structural disorder of In2O3, and even more lattice defects. It is well known that intrinsic defects or the nonstoichiometry mainly determine the sensing materials. So, the sensor can absorb more oxygen on the surface, and the 2%La-In2O3 sensor has much more chemisorbed oxygen than the pure In2O3 sensor. As proposed, the sensing response is expected to occur mainly via electron transfer and/or variation of chemisorbed oxygen species on the sensor surface, so the sensor performances can be improved with the increase of chemisorbed oxygen. Hence, the La-In2O3 sensor contains more chemisorbed oxygen to react with the ethanol and brings about the higher response to it, which agrees with the testing results of the ethanol-sensing properties [37].
(2) The Au codoping in La-In2O3 material for ethanol response improvement can be explained from two aspects. First, the catalytic effect of Au significantly enhances the values of the absorbed oxygen, which results in the decrease of the optimal working temperature [38]. Second, the Shottky junction between Au and In2O3 will be formed on the In2O3 nanomaterials. As shown in Figure 8, it is known that the work function of In2O3 (4.8 eV) is lower than that of Au (5.1 eV), and electrons flow from the conduction band of In2O3 to the surface of Au, which makes the energy band of In2O3 more curved, and the interface of Au-doped In2O3 has an increased width of the electron depletion layer and increased resistance compared to the pure In2O3 interface [39,40]. When the sensor is put into ethanol vapors, the ethanol molecules react with a large number of oxygen negative ions and release electrons into the conduction band of In2O3, the width of the electron depletion layer decreases and the resistance decreases. This results in a sensor with good gas-sensitive performance.

4. Conclusions

The Au/La-In2O3 nanoparticles were prepared by a simple hydrothermal method and ultrasonic treatment. The structure and morphology of all the Au/La-In2O3 samples were characterized. The morphology of pure In2O3 and La-In2O3 showed porous aggregate state nanoparticles, while that of Au/La-In2O3 showed uniformly sized dispersed spheres of about 3 nm. The gas-sensing properties were investigated for ethanol vapor. The excellent gas-sensitive performance of Au/La-In2O3 sample was obtained for ethanol. Compared to the pure In2O3 materials, the 2%Au/2%La-In2O3 sensor achieved a highest response of 381, about 14 times that of the pure In2O3 sensor and about 8.9 times that of the 2%La-In2O3 sensors for 100 ppm ethanol. In addition, a fast response time (1 s), better selectivity and lower operating temperature were obtained for ethanol vapor. The enhancement mechanism of the ethanol response was discussed for Au/La-In2O3 nanoparticles. The above experimental results would provide a new road for designing higher-response sensors.

Author Contributions

Conceptualization, T.Z.; methodology, Y.Z.; Investigation, P.L.; data curation, B.L.; writing—review and editing, H.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Written informed consent has been obtained from the patient(s) to publish this paper.

Data Availability Statement

Not applicable.

Acknowledgments

This work was supported by Natural Science Foundation of China (Grant No.: 61705077); Project of Jilin Provincial Science and Technology Department (20200403072SF); Project of Jilin Province Development and Reform Commission (2019C048-4, 2020C021-5).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Vishwakarma, A.K.; Yadava, L. Structural and sensing properties of ethanol gas using Pd-doped SnO2 thick film gas sensor. Environ. Sci. Pollut. Res. 2021, 28, 3920–3927. [Google Scholar] [CrossRef] [PubMed]
  2. Qin, W.; Yuan, Z.; Gao, H.; Meng, F. Ethanol sensors based on porous In2O3 nanosheet-assembled micro-flowers. Sensors 2020, 20, 3353. [Google Scholar] [CrossRef] [PubMed]
  3. Kuchi, P.S.; Roshan, H.; Sheikhi, M.H. A novel room temperature ethanol sensor based on PbS:SnS2 nanocomposite with enhanced ethanol sensing properties. J. Alloys Compd. 2020, 816, 152666. [Google Scholar] [CrossRef]
  4. Xu, K.; Zhao, W.; Yu, X.; Duan, S.; Zeng, W. Enhanced ethanol sensing performance using Co3O4-ZnSnO3 arrays prepared on alumina substrates. Phys. E. Low-Dimens. Syst. Nanostruct. 2020, 117, 113825. [Google Scholar] [CrossRef]
  5. Kim, J.H.; Wu, P.; Kim, H.W.; Kim, S.S. Highly selective sensing of CO, C6H6 and C7H8 gases by catalytic functionalization with metal nanoparticles. ACS Appl. Mater. Interfaces 2016, 8, 7173–7183. [Google Scholar] [CrossRef] [PubMed]
  6. Li, G.; Su, Y.; Chen, X.; Chen, L.; Li, Y.; Guo, Z. Enhanced chemiresistive sensing performance of well-defined porous CuO-doped ZnO nanobelts toward VOCs. Nanoscale Adv. 2019, 1, 3900–3908. [Google Scholar] [CrossRef] [PubMed]
  7. Zhang, X.; Song, D.; Liu, Q.; Chen, R.; Liu, J.; Zhang, H.; Yu, J.; Liu, P.; Wang, J. Designed synthesis of Co-doped spongy-like In2O3 for high sensitive detection of acetone gas. Cryst. Eng Commun. 2019, 21, 1876–1885. [Google Scholar] [CrossRef]
  8. Anand, K.; Kaur, J.; Singh, R.C.; Thangaraj, R. Structural, optical and gas sensing properties of pure and Mn-doped In2O3 nanoparticles. Ceram. Int. 2016, 3, 223. [Google Scholar] [CrossRef]
  9. Hu, J.; Chen, Y.; Ma, Z.; Zeng, L.; Zhou, D.; Peng, Z.; Sun, W.; Liu, Y. Temperature-compensated optical fiber sensor for volatile organic compound gas detection based on cholesteric liquid crystal. Opt. Lett. 2021, 46, 3324–3327. [Google Scholar] [CrossRef] [PubMed]
  10. Ahmed, M.A.; Monir, H.; Al-Keisy, A. Evaluation of the graphene as gas sensor for NH3 via electrical properties. Surf. Rev. Lett. 2020, 27, 1950215. [Google Scholar] [CrossRef]
  11. Wang, L.; Ma, S.; Li, J.; Wu, A.; Luo, D.; Yang, T.; Cao, P.; Ma, N.; Cai, Y. Mo-doped SnO2 nanotubes sensor with abundant oxygen vacancies for ethanol detection. Sens. Actuators B Chem. 2021, 347, 130642. [Google Scholar] [CrossRef]
  12. Chang, X.; Xu, S.; Liu, S.; Wang, N.; Sun, S.; Zhu, X.; Li, J.; Ola, O.; Zhu, Y. Highly sensitive acetone sensor based on WO3 nanosheets derived from WS2 nanoparticles with inorganic fullerene-like structures. Sens. Actuators B Chem. 2021, 343, 130135. [Google Scholar] [CrossRef]
  13. Ivanovskaya, M.; Kotsikau, D.; Faglia, G.; Nelli, P. Influence of chemical composition and structural factors of Fe2O3/In2O3 sensors on their selectivity and sensitivity to ethanol. Sens. Actuators B Chem. 2003, 96, 498–503. [Google Scholar] [CrossRef]
  14. Park, S.; Ko, H.; An, S.; Lee, W.I.; Lee, S.; Lee, C. Synthesis and ethanol sensing properties of CuO nanorods coated with In2O3. Ceram. Int. 2013, 39, 5255–5262. [Google Scholar] [CrossRef]
  15. Saito, N.; Haneda, H.; Watanabe, K.; Shimanoe, K.; Sakaguchi, I. Highly sensitive isoprene gas sensor using Au-loaded pyramid-shaped ZnO particles. Sens. Actuators B Chem. 2021, 326, 128999. [Google Scholar] [CrossRef]
  16. Akamatsu, T.; Itoh, T.; Tsuruta, A.; Masuda, Y. CH3SH and H2S sensing properties of V2O5/WO3/TiO2 gas sensor. Chemosensors 2021, 9, 113. [Google Scholar] [CrossRef]
  17. Liu, J.; Zhang, L.; Cheng, B.; Fan, J.; Yu, J. A high-response formaldehyde sensor based on fibrous Ag-ZnO/In2O3 with multi-level heterojunctions. J. Hazard. Mater. 2021, 413, 125352. [Google Scholar] [CrossRef]
  18. Hu, J.; Liang, Y.; Sun, Y.; Zhao, Z.; Zhang, M.; Li, P.; Zhang, W.; Chen, Y.; Zhuiykov, S. Highly sensitive NO2 detection on ppb level by devices based on Pd-loaded In2O3 hierarchical microstructures. Sens. Actuators B Chem. 2017, 252, 116–126. [Google Scholar] [CrossRef]
  19. Han, B.; Wang, H.; Yang, W.; Wang, J.; Wei, X. Hierarchical Pt-decorated In2O3 microspheres with highly enhanced isoprene sensing properties. Ceram. Int. 2021, 47, 9477–9485. [Google Scholar] [CrossRef]
  20. Xu, X.; Fan, H.; Liu, Y.; Wang, L.; Zhang, T. Au-loaded In2O3 nanofibers-based ethanol micro gas sensor with low power consumption. Sens. Actuators B Chem. 2011, 160, 713–719. [Google Scholar] [CrossRef]
  21. Li, S.; Diao, Y.; Yang, Z.; He, J.; Wang, J.; Liu, C.; Liu, F.; Lu, H.; Yan, X.; Sun, P.; et al. Enhanced room temperature gas sensor based on Au-loaded mesoporous In2O3 nanospheres@polyaniline core-shell nanohybrid assembled on flexible PET substrate for NH3 detection. Sens. Actuators B Chem. 2018, 276, 526–533. [Google Scholar] [CrossRef]
  22. Bai, J.; Luo, Y.; An, B.; Wang, Q.; Cheng, X.; Li, J.; Pan, X.; Zhou, J.; Wang, Y.; Xie, E. Ni/Au bimetal decorated In2O3 nanotubes for ultra-sensitive ethanol detection. Sens. Actuators B Chem. 2020, 311, 127938. [Google Scholar] [CrossRef]
  23. Jiang, Z.; Zhao, R.; Sun, B.; Nie, G.; Ji, H.; Lie, J.; Wang, C. Highly sensitive acetone sensor based on Eu-doped SnO2 electrospun nanofibers. Ceram. Int. 2016, 42, 15881–15888. [Google Scholar] [CrossRef]
  24. Habibzadeh, S.; Khodadadi, A.A.; Mortazavi, Y. CO and ethanol dual selective sensor of Sm2O3-doped SnO2 nanoparticles synthesized by microwave-induced combustion. Sens. Actuators B Chem. 2010, 144, 131–138. [Google Scholar] [CrossRef]
  25. Qin, W.F.; Xu, L.; Song, J.; Xing, R.Q.; Song, H.W. Highly enhanced gas sensing properties of porous SnO2-CeO2 composite nanofibers prepared by electrospinning. Sens. Actuators B Chem. 2013, 185, 231–237. [Google Scholar] [CrossRef]
  26. Chao, L.; Lei, B.; Fang, S.; Xu, J.; Chen, R. Synthesis and characterization of Y-doped SnO2 as sensor materials. J. Porous Mater. 2007, 14, 481–484. [Google Scholar]
  27. Anand, K.; Kaur, J.; Singh, R.C.; Thangaraj, R. Temperature dependent selectivity towards ethanol and acetone of Dy3+-doped In2O3 nanoparticles. Chem. Phys. Lett. 2017, 670, 37–45. [Google Scholar] [CrossRef]
  28. Duan, H.; Wang, Y.; Li, S.; Li, H.; Liu, L.; Du, L.; Cheng, Y. Controllable synthesis of Ho-doped In2O3 porous nanotubes by electrospinning and their application as an ethanol gas sensor. J. Mater. Sci. 2018, 53, 3267–3269. [Google Scholar] [CrossRef]
  29. Qin, Z.; Liu, Y.; Chen, W.; Ai, P.; Wu, Y.; Li, S.; Yu, D. Highly sensitive alcohol sensor based on a single Er-doped In2O3 nanoribbon. Chem. Phys. Lett. 2016, 646, 12–17. [Google Scholar] [CrossRef]
  30. Wang, H.; Chen, M.; Rong, Q.; Zhang, Y.; Hu, J.; Zhang, D.; Zhou, S.; Zhao, X.; Zhang, J.; Zhu, Z.; et al. Ultrasensitive xylene gas sensor based on flower-like SnO2/Co3O4 nanorods composites prepared by facile two-step synthesis method. Nanotechnology 2020, 31, 255501. [Google Scholar] [CrossRef] [PubMed]
  31. Wang, Y.; Lin, Y.; Jiang, D.; Li, F.; Li, C.; Zhu, L.; Wen, S.; Ruan, S. Special nanostructure control of ethanol sensing characteristics based on Au@In2O3 sensor with good selectivity and rapid response. RSC Adv. 2015, 5, 9884–9890. [Google Scholar] [CrossRef]
  32. Chen, F.; Yang, M.; Wang, X.; Song, Y.; Guo, L.; Xie, N.; Kou, X.; Xu, X.; Sun, Y.; Lu, G. Template-free synthesis of cubic-rhombohedral-In2O3 flower for ppb level acetone detection. Sens. Actuators B Chem. 2019, 290, 459–466. [Google Scholar] [CrossRef]
  33. Song, L.; Dou, K.; Wang, R.; Leng, P.; Luo, L.; Xi, Y.; Kaun, C.; Han, N.; Wang, F.; Chen, Y. Sr-doped cubic In2O3/rhombohedral In2O3 homojunction nanowires for highly sensitive and selective breath ethanol sensing: Experiment and DFT simulation studies. ACS Appl. Mater. Interfaces 2020, 12, 1270–1279. [Google Scholar] [CrossRef]
  34. Xiao, H.; Xue, C.; Song, P.; Li, J.; Wang, Q. Preparation of porous LaFeO3 microspheres and their gas-sensing property. Appl. Surf. Sci. 2015, 337, 65–71. [Google Scholar] [CrossRef]
  35. Qiao, X.; Xun, Y.; Yang, K.; Ma, J.; Li, C.; Wang, H.; Jia, L. Mo doped BiVO4 gas sensor with high sensitivity and selectivity towards H2S. Chem. Eng. J. 2020, 395, 125144. [Google Scholar] [CrossRef]
  36. Zhao, C.; Gong, H.; Niu, G.; Wang, F. Electrospun Ca-doped In2O3 nanotubes for ethanol detection with enhanced sensitivity and selectivity. Sens. Actuators B Chem. 2019, 299, 126946. [Google Scholar] [CrossRef]
  37. Kapse, V.D.; Ghosh, S.A.; Chaudhari, G.N.; Raghuwanshi, F.C.; Gulwade, D.D. H2S sensing properties of La-doped nanocrystalline In2O3. Vacuum 2008, 83, 346–352. [Google Scholar] [CrossRef]
  38. Xia, H.; Wang, Y.; Kong, F.; Wang, S.; Zhu, B.; Guo, X.; Zhang, J.; Wang, Y.; Wu, S. Au-doped WO3-based sensor for NO2 detection at low operating temperature. Sens. Actuators B Chem. 2008, 134, 133–139. [Google Scholar] [CrossRef]
  39. Sun, Y.; Zhao, Z.; Zhou, R.; Li, P.; Zhang, W.; Suematsu, K.; Hu, J. Synthesis of In2O3 nanocubes, nanocube clusters, and nanocubes-embedded Au nanoparticles for conductometric CO sensors. Sens. Actuators B Chem. 2021, 345, 130433. [Google Scholar] [CrossRef]
  40. Hwang, J.; Jung, H.; Shin, H.; Kim, D.; Kim, D.S.; Ju, B.; Chun, M. The effect of noble metals on co gas sensing properties of In2O3 nanoparticles. Appl. Sci. 2021, 11, 4903. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of the pure In2O3, 2 mol%La-In2O3 and x mol%Au/La-In2O3 (x = 1, 2, 3) samples.
Figure 1. XRD patterns of the pure In2O3, 2 mol%La-In2O3 and x mol%Au/La-In2O3 (x = 1, 2, 3) samples.
Chemosensors 10 00392 g001
Figure 2. The SEM images of pure In2O3 (a), 2 mol%La-In2O3 (b) and 2 mol%Au/2 mol%La-In2O3 (c).
Figure 2. The SEM images of pure In2O3 (a), 2 mol%La-In2O3 (b) and 2 mol%Au/2 mol%La-In2O3 (c).
Chemosensors 10 00392 g002
Figure 3. Response of In2O3 doped with different Au and La content to 100 ppm ethanol at 190–260 °C.
Figure 3. Response of In2O3 doped with different Au and La content to 100 ppm ethanol at 190–260 °C.
Chemosensors 10 00392 g003
Figure 4. Response of 2%Au/2%La-In2O3 sensor to the same concentration of ethanol (a) and different concentrations of ethanol (b) at optimum operating temperature.
Figure 4. Response of 2%Au/2%La-In2O3 sensor to the same concentration of ethanol (a) and different concentrations of ethanol (b) at optimum operating temperature.
Chemosensors 10 00392 g004
Figure 5. The gas sensing performance: (a) The response to different ethanol concentrations. (b) The linear fitting relationship between response and ethanol concentration.
Figure 5. The gas sensing performance: (a) The response to different ethanol concentrations. (b) The linear fitting relationship between response and ethanol concentration.
Chemosensors 10 00392 g005
Figure 6. Selectivity of 2%Au/2%La, 2%La, pure In2O3-based three-type sensor to 100 ppm ethanol, trimethylamine, methanol, sulfur dioxide, xylene, ammonia.
Figure 6. Selectivity of 2%Au/2%La, 2%La, pure In2O3-based three-type sensor to 100 ppm ethanol, trimethylamine, methanol, sulfur dioxide, xylene, ammonia.
Chemosensors 10 00392 g006
Figure 7. Single cell structure of In2O3 and La3+ replaces In3+.
Figure 7. Single cell structure of In2O3 and La3+ replaces In3+.
Chemosensors 10 00392 g007
Figure 8. Energy level diagrams of (a) pure In2O3 and (b) Au/La-In2O3 in air.
Figure 8. Energy level diagrams of (a) pure In2O3 and (b) Au/La-In2O3 in air.
Chemosensors 10 00392 g008
Table 1. Lattice parameters and crystallite size of pure In2O3, 2 mol% La-In2O3 and x mol% Au/La-In2O3 (x = 1, 2, 3).
Table 1. Lattice parameters and crystallite size of pure In2O3, 2 mol% La-In2O3 and x mol% Au/La-In2O3 (x = 1, 2, 3).
SampleLattice Parameters (Å)Crystallite Size (nm)
pure In2O3a = 5.490, b = 5.490, c = 14.50818.29
2%La:In2O3a = 5.482, b = 5.482, c = 14.51512.17
1%Au/2%La:In2O3a = 5.490, b = 5.490, c = 14.52013.18
2%Au/2%La:In2O3a = 5.488, b = 5.488, c = 14.54412.59
3%Au/2%La:In2O3a = 5.489, b = 5.489, c = 14.49113.11
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, H.; Li, P.; Liu, B.; Zhang, T.; Zhang, Y. Au/La Co-Modified In2O3 Nanospheres for Highly Sensitive Ethanol Gas Detection. Chemosensors 2022, 10, 392. https://doi.org/10.3390/chemosensors10100392

AMA Style

Liu H, Li P, Liu B, Zhang T, Zhang Y. Au/La Co-Modified In2O3 Nanospheres for Highly Sensitive Ethanol Gas Detection. Chemosensors. 2022; 10(10):392. https://doi.org/10.3390/chemosensors10100392

Chicago/Turabian Style

Liu, Hang, Peihua Li, Bing Liu, Tong Zhang, and Yuhong Zhang. 2022. "Au/La Co-Modified In2O3 Nanospheres for Highly Sensitive Ethanol Gas Detection" Chemosensors 10, no. 10: 392. https://doi.org/10.3390/chemosensors10100392

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