• Journal of Inorganic Materials
  • Vol. 36, Issue 11, 1223 (2021)
Pengpeng LI, Bing WANG*, and Yingde WANG*
Author Affiliations
  • Science and Technology on Advanced Ceramic Fibers and Composites Laboratory, College of Aerospace Science and Engineering, National University of Defense Technology, Changsha 410073, China
  • show less
    DOI: 10.15541/jim20210142 Cite this Article
    Pengpeng LI, Bing WANG, Yingde WANG. Ultrafast CO Sensor Based on Flame-annealed Porous CeO2 Nanosheets for Environmental Application [J]. Journal of Inorganic Materials, 2021, 36(11): 1223 Copy Citation Text show less

    Abstract

    As a highly toxic gas, CO is one of the culprits of air pollution. Long-term inhalation will also cause great harm to the human body and even death. How to achieve rapid CO monitoring is an important challenge in sensing field. CO as one of the of air pollution, long time inhalation will reduce the oxygen carrying capacity of blood and in severe cases death. Therefore, effective monitoring of CO is necessary. In this study, porous CeO2 nanosheets (CeO2 NSs) were obtained via flame annealing intermediate product CeCO3OH nanosheets synthesized by simple hydrothermal method. Through controlling of the flame time, oxygen vacancies were introduced into the CeO2 NSs. As a result, the CeO2 NSs annealed with 2 min (CeO2-2min NSs) exhibited outstanding reproducibility and selectivity towards CO gas. Particularly, the response/recovery time were extremely fast (2 s/2 s) towards 500 μL/L CO at 450 ℃ as well as finely functional relationship between response and concentration of CO at a wide detection range (10-10000 μL/L). The superior gas sensing performance of CeO2-2min NSs can be ascribed to the high aspect ratios of the porous two-dimensional structure and abundant oxygen vacancies in the crystals. This work may supply a strategy for designing ultrafast gas sensors which can detect target gas at a wide range.

    Carbon monoxide (CO) which is colorless, odorless, but toxic, explosive, mainly produced by automotive emissions, coal combustion and forest fires etc.[1] Actually, any incomplete combustion of carbon could generate CO. However, CO is extremely harmful to human beings causing dizziness, nausea or even death[2]. According to the data from World Health Organization, one can tolerate 9 μL/L CO within 8 h, 26 μL/L within 1 h, and beyond this the gas could threaten human life[3]. As a result, there are various CO gas sensors to monitor CO. However, to the best of our knowledge, the detection ranges of these CO gas sensors are limited. Therefore, developing CO gas sensors with a wide detection range becomes a crucially important task.

    Many materials could be used to detect CO gas[4]. Taking flexibility and versatility in practical applications into account, metal oxides are promising sensing materials for CO gas sensors[5]. Specifically, SnO2, ZnO, Co3O4 and TiO2etc. have been reported as CO gas sensors, which exhibit a detection range from dozens of μL/L to thousands of μL/L[6,7]. Among them, cerium oxide (CeO2), featured of structural stability and high redox capability[8], has been extensively investigated in gas sensing field[9]. Good CO sensing performance was obtained by pristine CeO2 nanospheres with high surface area and the detection range was 10-30 μL/L[10]. SnO2 mixed CeO2 thin films exhibited a response/recovery time of 26 s/30 s towards 500 μL/L CO and the detection concentration was in the range of 5-5000 μL/L[11]. Although various CeO2, such as samples with different morphologies, mixture of other metal oxides or precious metals, has been reported, the detection range of these CO sensors are not wide enough as well as dozens of seconds of response time to meet practical requirements in some cases.

    Two-dimensional nanosheets exhibit superiority in specific surface area, mobility of ions and the number of active sites, being applied in many aspects[12]. In the context of gas sensing, two-dimensional nanosheets also show remarkable performance. For instance, Liu et al.[13] synthesized porous ultrathin ZnO nanosheets with high specific surface areas for acetylacetone gas sensing and presented significant results, demonstrating the important role of oxygen vacancies in gas sensing spontaneously. Self-assembled monolayer CuO nanosheets were used for H2S gas sensing with improved sensitivity[14]. Selective alkene gas sensing was achieved by SnO2 nanosheets with distinct performance, which is affected by the content of oxygen vacancies[15]. However, to the best of our knowledge, there is almost few reports about CeO2 NSs, especially for the influence of oxygen vacancies of CeO2 NSs on the sensing performance.

    In this study, we prepared CeO2 NSs via flame annealing under different time intermediate product CeCO3OH nanosheets synthesized by simple hydrothermal method. The as-prepared samples were characterized by various methods. These flame-annealed CeO2 NSs show ultrafast response/recovery time towards CO, wide-range CO detection as well as good reproducibility and selectivity. Moreover, the mechanism of CO gas sensing is discussed.

    1 Experimental

    1.1 Materials

    Cerium (Ⅲ) chloride heptahydrate (CeCl3·7H2O) and sodium oleate were purchased from Shanghai Macklin Biochemical Co., Ltd. Ammonium hydroxide (NH3·H2O) and cyclohexane were provided by Sinopharm Chemical Reagent Co., Ltd. All of the chemicals and solvents were reagent-grade without further purification. Deionized water was used in all procedure.

    1.2 Synthesis of CeO2 nanosheets

    Intermediate product CeCO3OH was synthesized by a simple hydrothermal method according to our previous work[16] as shown in Fig.1. The specific procedure was listed as follows. CeCl3·7H2O solution was dropped into sodium oleate solution with stirring. After that, NH3·H2O was dropped into it with stirring. Subsequently, the suspension was proceeded with hydrothermal reaction. The precipitant was collected by centrifugation and washed with cyclohexane. Whereafter, the intermediate product was annealed with liquefied butane gas flame for 0.5, 2, 5 min and the resulted CeO2 nanosheets were denoted as CeO2-0.5min NSs, CeO2-2min NSs, and CeO2-5min NSs, respectively (The flame temperature is in the range of 1000-1200 ℃). The samples were collected for further characterization.

    Schematic illustration of preparation of CeO2 nanosheets

    Figure 1.Schematic illustration of preparation of CeO2 nanosheets

    1.3 Characterization

    The crystal structure was confirmed by X-ray Diffraction (XRD, Bruker AXS D8 Advanced Diffractometer, German) with Cu Kα radiation (λ=0.15418 nm). The morphology and structures were characterized via transmission electron microscope (TEM, Tian G2 60-300, FEI, USA). The element compositions and chemical states were obtained by X-ray photoelectron spectroscope (XPS, Thermo Scientific Escalab 250Xi, USA) equipped with a monochromatic Al Kα source. The specific surface area and porosity were evaluated using an automated surface area and pore size analyzer (Quantachrome Autosorb iQ3, USA), calculated by Brunauer-Emmett-Teller (BET) and Barrett-Joyner-Halenda (BJH) methods on the basis of nitrogen adsorption-desorption isotherms, respectively. The presence of oxygen vacancies was determined using electron paramagnetic resonance (EPR, Bruker A300, German).

    1.4 Gas sensing test

    Specialized CGS-1 TP intelligent gas sensing analyzing system (Beijing Elite Tech Co., Ltd, Beijing, China) was adopted in this work and the fabrication detail and testing principal could be found in our previous report[17]. The gas sensing response (S) was defined as the following formula:

    $S=\frac{|{{R}_{\text{a}}}-{{R}_{\text{g}}}|}{{{R}_{\text{a}}}}\times 100\%$

    where Ra and Rg represent the resistance of sensing materials in air and target gas, respectively. The response and recovery time were labelled as the time to reach 90% of the total resistance change.

    2 Results and discussion

    2.1 Characterization

    In order to identity the crystal structure, the intermediate product and the as-prepared CeO2 nanosheets were carefully investigated by XRD. The patterns of intermediate product perfectly matches with CeCO3OH (JCPDF 32-0189) as shown in Fig. S1. XRD patterns of CeO2- 0.5min NSs, CeO2-2min NSs and CeO2-5min NSs show sharp diffraction peaks, which demonstrates high crystallinity as shown in Fig. 2. These diffraction peaks can be well indexed to cubic fluorite cerium oxide (JCPDF 34-0394). The typical peaks located at 2θ=28.6°, 33.0°, 47.5°, 56.3°, 59.1°, 69.4°, 76.7°, 79.0°, correspond to (111), (200), (220), (311), (222), (400), (331), (420) crystal planes of fluorite phase of CeO2, respectively. However, the width of main diffraction peak of (111) at 2θ=28.6° is quite different. With flame annealing time increasing, the diffraction peak width become narrower, demonstrating the increment of crystal sizes. The crystal sizes based on (111) crystal plane of CeO2-0.5min NSs, CeO2-2min NSs and CeO2-5min NSs were calculated to be 12.12, 17.02 and 20.32 nm, respectively, according to the Scherer equation.

    XRD patterns of the flame-annealed CeO2 nanosheets

    Figure 2.XRD patterns of the flame-annealed CeO2 nanosheets

    The element compositions and surface chemical states could be detected by XPS, and the details are given in Fig. 3. For O element, the O1s core level spectra of CeO2- 0.5min NSs, CeO2-2min NSs and CeO2-5min NSs can be deconvolved into three peaks, i.e. lattice oxygen (OL, ca. 529.0 eV), deficient oxygen (OV, ca. 529.5 eV) and sorbed oxygen (OS, ca. 531.2 eV) in Fig. 3(a, b, c), respectively[18]. As is well-known, OV and OS (O2-, O- or O2-) play absolutely important roles in gas sensing, contributing to improving response[19]. Table S1 provides the ratio of each oxygen species to the whole oxygen extent. In this study, CeO2-2min NSs possesses the largest amount of OV (43.2%), and these for CeO2-0.5min NSs and CeO2-5min NSs are relatively smaller. For Ce element, the Ce3d core level spectra of CeO2-0.5min NSs, CeO2-2min NSs and CeO2-5min NSs can be deconvolved into ten peaks in Fig. 3(d, e, f), respectively. The peaks at about 916.4, 907.1, 902.4 and 900.5 eV are designated to Ce3d3/2, and that at about 898.2, 897.8, 888.5, 884.5, 882.1 and 881.4 eV pertain to Ce3d5/2[20]. Thereinto, the peaks at about 881.4, 884.5, 897.8 and 902.4 eV are assigned to Ce3+, and those at around 882.1, 888.5, 898.2, 900.5, 907.1 and 916.4 eV belong to Ce4+[21]. Moreover, the ratio of Ce3+ to the total extent of Ce is presented in Table S1. It was reported that there is a transformation between Ce4+ and Ce3+ which determines by the oxygen partial pressure. That means that with more Ce3+, more OV exists in the samples[8]. In this case, CeO2-2min NSs holds the largest part of Ce3+ (28.6%), consistently with the largest part of OV (43.2%).

    XPS spectra of the CeO2 NSs (a, b, c) high resolution spectra of O 1s of CeO2-0.5min NSs, CeO2-2min NSs and CeO2-5min NSs; (d, e, f) high resolution spectra of Ce 3d of CeO2-0.5min NSs, CeO2-2min NSs and CeO2-5min NSs

    Figure 3.XPS spectra of the CeO2 NSs (a, b, c) high resolution spectra of O 1s of CeO2-0.5min NSs, CeO2-2min NSs and CeO2-5min NSs; (d, e, f) high resolution spectra of Ce 3d of CeO2-0.5min NSs, CeO2-2min NSs and CeO2-5min NSs

    The morphology and structure were characterized by TEM. The spectra of CeO2-0.5min NSs, CeO2-2min NSs and CeO2-5min NSs are given in Fig. 4(a, b, c), respectively, which are transparent nanosheets with pores (labelled by red dotted circles). Nevertheless, there is a difference in pore diameters that with the increase of flame time, the pore diameter becomes larger. This result is consistent with the tendency of nitrogen adsorption-desorption discussed later, which may influence the gas sensing performance. More importantly, with the increment of flame annealing time, the grain sizes become larger, which is consistent with the results of XRD. The crystal structure and crystal lattice spacing of the samples can be observed by HRTEM, as shown in Fig. 4(d,e,f), respectively. The distance of two adjacent crystal planes is 0.31 nm, corresponding to (111) planes of the CeO2 fluorite phase. Obviously, crystallinity increases along with a longer flame annealing time.

    Morphologies and structure characterizations of the as-prepared samples(a, b, c) TEM images of CeO2-0.5min NSs, CeO2-2min NSs and CeO2-5min NSs (pores are labelled by red dotted circles); (d, e, f) HRTEM images of CeO2-0.5min NSs, CeO2-2min NSs and CeO2-5min NSs

    Figure 4.Morphologies and structure characterizations of the as-prepared samples(a, b, c) TEM images of CeO2-0.5min NSs, CeO2-2min NSs and CeO2-5min NSs (pores are labelled by red dotted circles); (d, e, f) HRTEM images of CeO2-0.5min NSs, CeO2-2min NSs and CeO2-5min NSs

    Nitrogen adsorption-desorption isothermal curves were performed to characterize the aspect surface area and porosity of the samples, as shown in Fig. 5. It is observed from Fig. 5(a) that all samples exhibit the isotherm feature of the IV-type with a hysteresis loop according to the IUPAC classification, which indicates the existence of a mesoporous structure in nanosheets[22]. The surface area of CeO2-0.5min NSs, CeO2-2min NSs and CeO2-5min NSs are 73.344, 46.804 and 40.836 m2∙g-1, respectively, summarized in Table S2. It is noticed that with the flame annealing time increasing, the surface area becomes smaller, which may be contributed to congregation and growth of small grains during the annealing process. Nevertheless, there is a variance in pore size distribution.As shown in Fig. 5(b), CeO2-2min NSs (15.5 nm) and CeO2-5min NSs (15.8 nm) possess larger pore diameter than CeO2-0.5min NSs (10.1 nm), which indicates gas molecules easily adsorb and desorb on the surface of CeO2-2min NSs and CeO2-5min NSs as well as a fast response/recovery time.

    (a) Nitrogen adsorption-desorption isotherms and (b) pore size distributions of CeO2 NSs

    Figure 5.(a) Nitrogen adsorption-desorption isotherms and (b) pore size distributions of CeO2 NSs

    To further verify the presence of oxygen vacancies, EPR were performed to ascertain relative amount of oxygen vacancies in each sample, as shown in Fig. 6. The value of g represents different species which contain unpaired electron. In this case, all samples have the same value of g(2.0038), which typically demonstrates the presence of oxygen vacancies in the samples[19]. The intensity of CeO2-2min NSs and CeO2-5min NSs is almost identical and far larger than that for CeO2-0.5min NSs, which confirms that oxygen vacancies in CeO2-2min NSs and CeO2-5min NSs are abundant and much higher than that in CeO2-0.5min NSs, consistent with the results of XPS.

    EPR spectra of CeO2 NSs

    Figure 6.EPR spectra of CeO2 NSs

    2.2 Gas sensing properties

    We systematically investigated the gas sensing performances of the samples, as shown in Fig. 7. First of all, the response of the sensors towards 500 μL/L CO were measured from 300 ℃ to 500 ℃ with the interval of 50 ℃ to determine the optimal temperature. For CeO2-2min NSs and CeO2-5min NSs, the optimal temperature is 450 ℃, and that for CeO2-0.5min NSs is 400 ℃ in Fig. 7(a). However, the responses of CeO2-2min NSs (10%) and CeO2-5min NSs (9%) are far higher than that for CeO2- 0.5min NSs (4%). The responses of CeO2-2min NSs and CeO2-5min NSs are considerable, which may be ascribed to the comparable OV content demonstrated by XPS (Table S1, 43.2% and 42.7%, respectively) and EPR (Fig. 6). For a little higher response of CeO2-2min NSs than CeO2-5min NSs, the difference between surface area (Table S2) and the content of Ce3+ (Table S1) may account for it. However, there is a large gap between the response of CeO2-2min NSs and CeO2-0.5min NSs, which may be resulted from the discrepancy of the amount of OV (Table S1 and Fig. 6) and pore diameters (Fig. 5(b)). Therefore, CeO2-2min NSs was chosen to proceed further testing at its optimal temperature(450 ℃).

    Gas sensing properties of CeO2 NSs (a) Response of CeO2 NSs at different temperatures from 300 ℃ to 500 ℃ with an interval of 50 ℃ towards 500 μL/L CO; (b) Response of CeO2-2min NSs towards 500 μL/L CO at 450 ℃ for seven periods; (c) Transient response of CeO2-2min NSs towards different concentration from 10 μL/L to 10000 μL/L at 450 ℃; (d) Fitting of CO concentration and its corresponding response; (e) Determination of response/recovery time of CeO2-2min NSs towards 500 μL/L CO at 450 ℃; (f) Selectivity of CeO2-2min NSs towards CH4, H2, NH3, NO2 and CO at 450 ℃

    Figure 7.Gas sensing properties of CeO2 NSs (a) Response of CeO2 NSs at different temperatures from 300 ℃ to 500 ℃ with an interval of 50 ℃ towards 500 μL/L CO; (b) Response of CeO2-2min NSs towards 500 μL/L CO at 450 ℃ for seven periods; (c) Transient response of CeO2-2min NSs towards different concentration from 10 μL/L to 10000 μL/L at 450 ℃; (d) Fitting of CO concentration and its corresponding response; (e) Determination of response/recovery time of CeO2-2min NSs towards 500 μL/L CO at 450 ℃; (f) Selectivity of CeO2-2min NSs towards CH4, H2, NH3, NO2 and CO at 450 ℃

    Reproducibility is a basic requirement for gas sensors to practical application. 500 μL/L CO at 450 ℃ was adopted to test reproducibility. In Fig. 7(b), it is found that the response of 7-time testing is stable without any decrease. It demonstrates that CeO2-2min NSs possesses well reproducible ability in CO gas sensing.

    Relationship between CO concentration and its response is important to determine real CO concentration in CO gas sensor applications. Different concentration of CO was subject to test at 450 ℃ as shown in Fig. 7(c). The highest concentration of the test is 10000 μL/L, and with the decrease of the concentration, the response also declines. Meanwhile, the lowest concentration tested is 10 μL/L with the response of 1.5%. Furthermore, the response and the concentration of CO (10-10000 μL/L) is fitted as shown in Fig. 7(d) and it keeps functional relationship of y=1.09x0.45 with R2=0.9908. It proves that CeO2-2min NSs is a promising material in CO gas sensing with a wide range. In addition, it is impressive that the response/recovery time are ultrafast in the sensing procedure, which is expected for a gas sensor.

    To investigate response/recovery time carefully, we chose a cycle of transient response in reproducibility test for further research. As can be seen in Fig. 7(e), the response and recovery time are both 2 s, which is very short in gas sensing. As discussed above, carriers on nanosheets move quickly because the morphology constraints its pathway in planes. In this case, electrons move in planes of CeO2 nanosheets, which promises ultrafast response/recovery time. Therefore, it is reasonable to deem that CeO2-2min NSs is advantageous in CO gas sensing. Moreover, recent progresses in CO gas sensing are summarized in Table 1, and it is concluded that CeO2-2min NSs outperforms in response/recovery time.

    OS/% OV/% Ce3+/% Ce3+/Ce4+
    CeO2-0.5min NSs 19.040.026.30.357
    CeO2-2min NSs 14.843.228.60.401
    CeO2-5min NSs 15.142.723.80.312

    Table 1.

    Percentage of each element species to the whole element content

    Selectivity is a valuable index to evaluate gas sensor. We also studied the selectivity of CeO2-2min NSs in gas sensing was studied as presented in Fig. 7(f). 500 μL/L of CH4, H2, NH3, NO2 and CO are subject to test, and it is found that the response of CO is much higher than that to other gases. Therefore, CeO2-2min NSs is very promising in CO gas sensing.

    In addition, it is noticed that humidity has obvious impact on CeO2-2min NSs response (Fig. S2), namely high humidity for low response, which may be arisen from the water molecules occupy the active sites on CeO2 NSs at high humidity. Besides, long-term stability of CeO2-2min was also performed as shown in Fig. S3. Although the response has a small decrease in two weeks, long-term stability is improved by doping or coating in followup work.

    2.3 Gas sensing mechanism

    The gas sensing mechanism of CeO2 NSs follows a surface charge model, which can be explained by the change of resistance in different target gases, as shown in Fig. 8[5,28]. In oxygen atmosphere (ambient air), oxygen molecules are adsorbed on the surface of the sensing materials by capturing free electrons, forming adsorbed oxygen anion species depended on the temperature[29,30]. In this study, when CeO2 NSs are exposed to air, oxygen molecules are adsorbed on the surface trapping electrons from them, which forms depletion layer and an increased resistance. Subsequently, oxygen anions change as the sequence of O2-, O- and O2- along with the temperature rising. At last, when reducing gas CO is introduced, the CO gas molecules adsorb on the surface of CeO2 NSs and then react with O2- (the operating temperature is 450 ℃). The whole procedure may proceed as following equations:

    ${{\text{O}}_{2(\text{ads})}}+{{\text{e}}^{-}}\to {{\text{O}}_{2}}{{^{-}}_{(\text{ads})}}$
    ${{\text{O}}_{2}}{{^{-}}_{(\text{ads})}}+{{\text{e}}^{-}}\to 2{{\text{O}}^{-}}_{(\text{ads})}$
    ${{\text{O}}^{-}}_{(\text{ads})}+{{\text{e}}^{-}}\to {{\text{O}}^{2-}}_{(\text{ads})}$
    $\text{CO}+{{\text{O}}^{2-}}_{(\text{ads})}\to \text{C}{{\text{O}}_{2}}+2{{\text{e}}^{-}}$

    Schematic illustration of CO gas sensing mechanism Yellow spheres represent O atoms; Orange spheres represent Ce atoms; Red rings represent the position of oxygen vacancies

    Figure 8.Schematic illustration of CO gas sensing mechanism Yellow spheres represent O atoms; Orange spheres represent Ce atoms; Red rings represent the position of oxygen vacancies

    During this reaction, the release of electrons into the surface of CeO2 NSs increases the concentration of carriers, leading to a thinner depletion layer as well as a decrease of resistance.

    In this case, CeO2-2min NSs possesses the largest part of OV, which is the active site for facilitating oxygen adsorption and dissociation by providing unpaired electrons[13]. Besides, CeO2-2min NSs holds largest pore diameter, which guarantees that CO molecules penetrate into the inside of sensing materials to touch more active sites[31]. Therefore, CeO2-2min NSs deserves the highest response as discussed above. On the contrary, the small amount of as well as small pore diameters of OV of CeO2-0.5min NSs, which pays for a low response. In addition, nanosheets can constrain carriers to transport in planes, which make carriers possess ultrafast mobility. On the other hand, oxygen vacancies can also increase mobility of carriers with easy movement between vacancies. These factors all contribute to ultrafast response/recovery time in this study.

    3 Conclusion

    In summary, CeO2 NSs via flame annealing intermediate product CeCO3OH nanosheets was synthesized by hydrothermal method with different flame time. Wherein, CeO2-2min NSs showed superb gas sensing performances. The response/recovery time were only 2 s/2 s, which was ultrafast in gas sensing field. In addition, CO concentration and its response kept well functional relationship at a wide detection range. Moreover, reproducibility and selectivity were also decent. The good performances are attributed to more oxygen vacancies and porous nanosheets morphology. As a result, this CeO2 NSs is reliable to apply in CO gas sensing. Meanwhile, flame annealing provides a novel method to prepare nanomaterials with simplicity and low cost.

    Supporting materials

    XRD pattern of the intermediate product

    Figure S1.XRD pattern of the intermediate product

    Response of CeO2-2min NSs towards 500 μL/L CO at 450 ℃ under different humidities

    Figure S2.Response of CeO2-2min NSs towards 500 μL/L CO at 450 ℃ under different humidities

    Long-term stability of CeO2-2min NSs towards 500 μL/L CO at 450 ℃ for two weeks

    Figure S3.Long-term stability of CeO2-2min NSs towards 500 μL/L CO at 450 ℃ for two weeks

    Surface area/(m2∙g-1) Pore volume/ (cm3∙g-1) Average pore diameter/nm
    CeO2-0.5min NSs 73.3440.17210.1
    CeO2-2min NSs 46.8040.18115.5
    CeO2-5min NSs 40.8360.16315.8

    Table 2.

    Summary of surface area and pore volume of CeO2 NSs

    Table Infomation Is Not Enable

    References

    [1] Z YANG, D ZHANG, D WANG. Carbon monoxide gas sensing properties of metal-organic frameworks-derived tin dioxide nanoparticles/molybdenum diselenide nanoflowers. Sensors and Actuators B: Chemical, 304, 127369(2019).

    [2] K BASU A, S CHAUHAN P, M AWASTHI et al. α-Fe2O3 loaded rGO nanosheets based fast response/recovery CO gas sensor at room temperature. Applied Surface Science, 465, 56-66(2019).

    [3] WHO Air Quality Guidelines-Global Update 2005. World Health Organization, Copenhagen(2006).

    [4] S YANG, C JIANG, H WEI S. Gas sensing in 2D materials. Applied Physics Reviews, 4, 021304(2017).

    [5] A DEY. Semiconductor metal oxide gas sensors: a review. Materials Science and Engineering: B, 229, 206-217(2018).

    [6] S MAHAJAN, S JAGTAP. Metal-oxide semiconductors for carbon monoxide (CO) gas sensing: a review. Applied Materials Today, 18, 100483(2020).

    [7] K GOVARDHAN, N GRACE A. Metal/Metal oxide doped semiconductor based metal oxide gas sensors-a review. Sensor Letters, 14, 741-750(2016).

    [8] C SUN, H LI, L CHEN. Nanostructured ceria-based materials: synthesis, properties, and applications. Energy & Environmental Science, 5, 8475-8505(2012).

    [9] Y LIU, Y LEI. Pt-CeO2 nanofibers based high-frequency impedancemetric gas sensor for selective CO and C3H8 detection in high- temperature harsh environment. Sensors and Actuators B: Chemical, 188, 1141-1147(2013).

    [10] D MAJUMDER, S ROY. Development of low-ppm CO sensors using pristine CeO2 nanospheres with high surface area. ACS Omega, 3, 4433-4440(2018).

    [11] M DURRANI S, M F AL-KUHAILI, I A BAKHTIARI et al. Investigation of the carbon monoxide gas sensing characteristics of tin oxide mixed cerium oxide thin films. Sensors (Basel), 12, 2598-2609(2012).

    [12] X LIU, T MA, N PINNA et al. Two-dimensional nanostructured materials for gas sensing. Advanced Functional Materials, 27, 1702168(2017).

    [13] F LIU, X WANG, X CHEN et al. Porous ZnO ultrathin nanosheets with high specific surface areas and abundant oxygen vacancies for acetylacetone gas sensing. ACS Applied Materials Interfaces, 11, 24757-24763(2019).

    [14] J MIAO, C CHEN, L MENG et al. Self-assembled monolayer of metal oxide nanosheet and structure and gas-sensing property relationship. ACS Sensors, 4, 1279-1290(2019).

    [15] G CHOI P, N IZU, N SHIRAHATA et al. SnO2 nanosheets for selective alkene gas sensing. ACS Applied Nano Materials, 2, 1820-1827(2019).

    [16] P LI, B WANG, C QIN et al. Band-gap-tunable CeO2 nanoparticles for room-temperature NH3 gas sensors. Ceramics International, 46, 19232-19240(2020).

    [17] P LI, B WANG, W LI et al. Effect of annealing atmosphere with different oxygen concentration on CO gas sensing performances for CeO2 nanoparticles. Materials Letters, 284, 129000(2021).

    [18] T SAMERJAI, D CHANNEI, C KHANTA et al. Flame-spray- made ZnInO alloyed nanoparticles for NO2 gas sensing. Journal of Alloys and Compounds, 680, 711-721(2016).

    [19] H YUAN, S ALJNEIBI, J YUAN et al. ZnO nanosheets abundant in oxygen vacancies derived from metal-organic frameworks for ppb-level gas sensing. Advanced Materials, 31, e1807161(2019).

    [20] H ZHAO, Y DONG, P JIANG et al. Highly dispersed CeO2 on TiO2 nanotube: a synergistic nanocomposite with superior peroxidase-like activity. ACS Applied Materials Interfaces, 7, 6451-6461(2015).

    [21] C DENG, Q HUANG, X ZHU et al. The influence of Mn-doped CeO2 on the activity of CuO/CeO2 in CO oxidation and NO+CO model reaction. Applied Surface Science, 389, 1033-1049(2016).

    [22] L YAO, Y LI, Y RAN et al. Construction of novel Pd-SnO2 composite nanoporous structure as a high-response sensor for methane gas. Journal of Alloys and Compounds, 826, 154063(2020).

    [23] C BUSACCA, A DONATO, M LO FARO et al. CO gas sensing performance of electrospun Co3O4 nanostructures at low operating temperature. Sensors and Actuators B: Chemical, 303, 127193(2020).

    [24] Q WANG, C WANG, H SUN et al. Microwave assisted synthesis of hierarchical Pd/SnO2 nanostructures for CO gas sensor. Sensors and Actuators B: Chemical, 222, 257-263(2016).

    [25] R MOHAMMADI M, J FRAY D. Nanostructured TiO2-CeO2 mixed oxides by an aqueous Sol-Gel process: effect of Ce:Ti molar ratio on physical and sensing properties. Sensors and Actuators B: Chemical, 150, 631-640(2010).

    [26] D TRUNG DO, D HOA N, V TONG P et al. Effective decoration of Pd nanoparticles on the surface of SnO2 nanowires for enhancement of CO gas-sensing performance. Journal of Hazard Materials, 265, 124-132(2014).

    [27] D KHOANG N, S HONG H, D TRUNG D et al. On-chip growth of wafer-scale planar-type ZnO nanorod sensors for effective detection of CO gas. Sensors and Actuators B: Chemical, 181, 529-536(2013).

    [28] C QIN, B WANG, N WU. Metal-organic frameworks derived porous Co3O4 dodecahedeons with abundant active Co3+ for ppb- level CO gas sensing. Applied Surface Science, 506, 144900(2020).

    [29] M XU J, P CHENG J. The advances of Co3O4 as gas sensing materials: a review. Journal of Alloys and Compounds, 686, 753-768(2016).

    [30] F YAN, G SHEN, X YANG et al. Low operating temperature and highly selective NH3 chemiresistive gas sensors based on Ag3PO4 semiconductor. Applied Surface Science, 479, 1141-1147(2019).

    [31] Z WANG, R YU. Hollow micro/nanostructured ceria-based materials: synthetic strategies and versatile applications. Advanced Materials, 31, e1800592(2019).

    Pengpeng LI, Bing WANG, Yingde WANG. Ultrafast CO Sensor Based on Flame-annealed Porous CeO2 Nanosheets for Environmental Application [J]. Journal of Inorganic Materials, 2021, 36(11): 1223
    Download Citation