Next Article in Journal
A Novel Substrate-Binding Site in the X-ray Structure of an Oxidized E. coli Glyceraldehyde 3-Phosphate Dehydrogenase Elucidated by Single-Wavelength Anomalous Dispersion
Previous Article in Journal
Multi-Step Concanavalin A Phase Separation and Early-Stage Nucleation Monitored Via Dynamic and Depolarized Light Scattering
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

A Simple Dip-Coating Method of SnO2-NiO Composite Thin Film on a Ceramic Tube Substrate for Methanol Sensing

1
State Key Laboratory of Integrated Optoelectronics, College of Electronic Science and Engineering, Jilin University, Changchun 130118, China
2
College of Chemistry, Jilin University, Changchun 130118, China
3
College of Electronic Science and Engineering, Jilin University, Changchun 130118, China
*
Authors to whom correspondence should be addressed.
Crystals 2019, 9(12), 621; https://doi.org/10.3390/cryst9120621
Submission received: 2 November 2019 / Revised: 21 November 2019 / Accepted: 25 November 2019 / Published: 26 November 2019

Abstract

:
In this work, SnO2-NiO composite thin film was successfully grown on a ceramic tube substrate directly by a simple dip-coating method combined with annealing. Characterization analysis demonstrates that uniform SnO2 film consists of a great number of nanospheres and NiO grows on SnO2 as an agglomerated block. In comparison to the pure SnO2 sample, the SnO2-NiO composite thin films gas sensor exhibits superior methanol sensing properties at 225 °C. The gas response to 10 ppm methanol reached 15.12 and the response and recovery times were 8 s and 7 s, respectively. The excellent selectivity and recovery rate are explained by the unique properties of the NiO semiconductor and the higher sensor response is attributed to the pivotal heterojunction effect.

Graphical Abstract

1. Introduction

SnO2 is a typical wide band-gap (Eg = 3.6 eV) n-type semiconductor and has been widely used in the gas sensor field owing to the advantage of its low cost, stable electrical properties, excellent physical and chemical stability and unique optical performance [1,2,3]. However, pure SnO2 gas sensors have several limitations such as poor selectivity [4], slow recovery [5] and high working temperature [6]. To date, many researchers have taken measures to improve its gas sensing performance, such as doping various elements and compositing several sensing materials [7,8]. In addition, an effective strategy which has been effectively demonstrated to control the carrier concentration and enhance the response is to construct heterojunctions among different semiconductor oxides [9,10]. Additionally, NiO is a typical p-type semiconductor with a 3.5 eV band gap and has been widely considered as an excellent additive to improve response/recovery speeds and stability [11]. Therefore, the combination of nickel oxide and tin dioxide materials has been investigated in several reports [12,13]. However, coating ceramic tubes with the powders is a necessary step in the fabrication of gas sensors, which leads to poor repeatability because the thickness cannot be controlled. Therefore, a direct growth process on ceramic tubes could solve this problem perfectly.
Methanol is one of the toxic gases among volatile organic compounds, and has been widely used in the production of formaldehyde, colorings and antifreeze [14]. Furthermore, it can be processed into fuel cells in automobiles as a clean form of energy [15]. However, small amounts of methanol may cause blindness and fatal diseases [16]. Therefore, a methanol gas sensor with low detection limit, high response and fast response and recovery rate is essential in daily life. At present, numerous research workers are contributing to research on methanol sensing. G. Korotcenkov et al. fabricated Au-modified SnO2 films through a successive ionic layer deposition method, which exhibits a high response to 50 ppm methanol [17]. A CdS-doped tin oxide thick film made up by Lallan et al. [18] employing a screen-printing technology has a 70 response and 100 s recover time to 5000 ppm methanol. Song et al. [19] reported a hierarchical tin oxide nanoflower structure by hydrothermal synthesis and showed a high response (58) as well as low operating temperature (200 °C) to 100 ppm methanol. Although they have achieved a high response to methanol associated with a fast response rate and low working temperature, their detection limit is more than 50 ppm, which motivates us to prepare a new gas sensor with an outstanding sensing performance combined with a low detection limit.
Herein, we prepared a complete SnO2-NiO composite thin film on a gas sensing ceramic tube directly through a dip-coating method, which avoided the cumbersome coating process and brought a repeatable growth technic. A comparative methanol response examination between SnO2 and SnO2-NiO composite is subsequently performed. The results of gas sensing measurements show that SnO2-NiO composite thin film (SnO2–NiO TF) possesses a prominent enhancement in selectivity, recovery rate (7 s) and gas response (15.12). Compared with previous work on methanol sensors, the SnO2-NiO composite thin film gas sensor exhibited a lower detection limit combined with high response, fast response (18 s) and recovery rate and excellent selectivity. The improvement mechanism regarding the decoration of NiO is also discussed.

2. Materials and Methods

All chemicals (analytical grade reagents) were obtained from Beijing Chemical Works (Beijing, China) without further purification. The preparation process of SnO2-NiO TF on the ceramic substrate was as follows: Firstly, in order to obtain Sn(OH)4 sol, 130 uL of citric acid (0.1 M) was added to a 100 mL solution of SnCl4 (0.13 M) and 0.1 M ammonia solution was added dropwise to raise the solution pH to 2. The resulting product was centrifuged, and the precipitate was added into 50 mL deionized water. The ammonia solution was added drop by drop to the mixed solution under magnetic stirring at 50 °C until the pH value was adjusted to 9. Secondly, the preparation of Ni(OH)2 sol: 3.7326 g Ni(CH3COO)2∙4H2O and 5 g PEG6000 was dissolved in 100 mL ethanol and the mixed solution was stirred continuously for 1 hour at 60 °C. Then, 10 mL of 25% ammonia was added dropwise and the resulting product was stirred at 70 °C for 1 day. Thirdly, ceramic tube substrates were dipped into the Sn(OH)4 sol for approximately 60 s and then, they were removed at a rate of 50 um/s. To avoid cracks in the Sn(OH)4 coating, the specimen was first dried at ambient temperature for 60 minutes and heated at 100 °C in a muffle furnace for about 60 minutes. The dip-coating process was repeated 3 times and the specimen was maintained at 500 °C for 2 hours to obtain tin dioxide thin film. After that, the dip-coating process and parameters of NiO were the same as those of SnO2.
The crystal structure information of the SnO2-NiO TF on a ceramic substrate were characterized through X-ray diffraction (XRD, Rigaku D/Max 2550, Rigaku, Ltd, Tokyo, Japan) with a Cu Kα radiation line of 0.1506 nm at 40 kV, 200 mA. The Field Emission Electron Microscopy (FESEM, JEOL JSM-7500 F, JEOL, Ltd, Tokyo, Japan) was used to determine the microstructure and morphology of the specimen. In addition, the bonding state of SnO2 and NiO was observed by means of the Transmission Electron Microscope (TEM, JEOL JEM-3010, JEOL, Ltd, Tokyo, Japan). In this paper, the Pt wires on as-prepared ceramic tubes and heating wires were welded to the bases and aged at 200 °C for 2 days to form gas sensor devices (Figure 1a). Besides, the static working system was used to measure the response of the gas sensor and is shown in Figure 1b. In detail, the device was placed in a 2.5 L as-prepared air chamber, and the resistance data was transferred to a Fluke 8846A digital multimeter connected to the computer. Furthermore, the heating wire inside the ceramic tube was used in conjunction with GPD3303S and the target gas was injected into the air chamber through a syringe. During the measurement, the Fluke 8846 Arecorded one resistor per second and displayed the data on the computer for real-time monitoring. The gas response was defined as S = Ra/Rg, in which Ra was the electrical resistance in the air chamber and Rg was that in the target gas environment. The response and recovery time were defined as the time taken to reach 90% of the final stable resistance change value.

3. Results and Discussion

The crystallographic information of SnO2-NiO TF on a ceramic tube substrate and pure ceramic substrate has been obtained out by X-ray diffraction analysis. The results are presented in Figure 2. Obviously, all the diffraction peaks of the ceramic substrate can be easily associated to the standard peak values of Al2O3 (JCPDS No.10-173) and SiO2 (JCPDS No.1-438), which indicates that these two oxides are the main component of the substrate. Furthermore, the diffraction peaks appearing at 26.86° and 34.08° can be indexed to SnO2 (JCPDS No.2-1337), which confirms that SnO2 has successfully grown on the ceramic tube substrate. However, the peaks of NiO do not appear in the sample, which is probably due to the low NiO concentration on the surface. For the purpose of confirming the crystal structure of NiO in the sample, as a comparison, pure NiO powder was obtained by heating the Ni(OH)2 sol-gel to a temperature of 500 °C, and X-ray diffraction analysis was carried out. The diffraction peaks of this material are in good agreement with NiO (JCPDS No.4-835) and the diffraction peak at 43.27° is also the observed in the diffraction pattern of Al2O3, which is another reason for that NiO is not detected.
The morphology and structure of pure SnO2 and SnO2-NiO TF on the ceramic tube substrate are shown in Figure 3. Compared with the surface of the pure ceramic tube in Figure 3a, a smooth film structure has appeared on the surface in Figure 3b. Combined with the result of XRD, the thin film consisted of SnO2, and the further details could be seen in Figure 3c, in which tiny SnO2 nanoparticles were found to be tightly stacked together. This aggregation mode may prevent the internal tin dioxide from contacting methanol and only conduct charge transfer, which is not conducive to increasing the sensitivity. The image in Figure 3d indicated that the thickness of SnO2 is about 301.5 nm. Moreover, we could see that NiO grows on SnO2 as an agglomerated block in Figure 3e. The high-magnification picture (Figure 3f) indicates the contact gap of NiO nanoparticles is larger than that of SnO2, which can provide a more reactive site to methanol. The low magnification of SnO2-NiO TF on the ceramic tube observed in Figure 3g exhibits a smooth and uniform surface of SnO2-NiO TF. In order to obtain the particle morphology of tin oxide and nickel oxide, the sample is characterized by using TEM and Figure 3h reveals that the particle diameter of SnO2 and NiO is about 8–9 nm. Further details of the lattice structure can be observed in Figure 3i through the high-resolution TEM image (HRTEM). A lattice spacing of 0.331 nm and 0.264 nm matches the SnO2 planes of (110) and (101), respectively [20,21], and the lattice spacing of 0.240 nm is the same as the NiO plane of (111) [22].
The gas sensors performance under different operation temperatures was also tested. Figure 4a displays the gas response of pure SnO2 and SnO2-NiO TF sensors in 10 ppm methanol at varying operating temperatures from 100 to 375 °C. The two response curves in this chart reveal that the maximum response of SnO2 and SnO2-NiO TF is observed under the optimal operating temperature of 225 °C, with the response values being 4.5 and 15.12, respectively. The two curves also show that the decoration of NiO evidently improves the response to methanol. The selectivity of sensors to 10 ppm of various gases was also investigated. Figure 4b indicates that the pure SnO2 sensor has similar responses to most of the test gases. On the other hand, compared with other gases, the SnO2-NiO TF gas sensor reveals an outstanding response to methanol, which indicates that SnO2-NiO TF has a unique selectivity to methanol and the decoration of NiO must play a vital role in this process. Figure 4c shows the sensor response to 10 ppm methanol at 225 °C with seven reversible cycles. This chart demonstrates both sensors exhibit an outstanding repeatability. Figure 4d indicates that the response time of both sensors to 10 ppm methanol are 18 s and the recovery times of pure SnO2 and SnO2-NiO TF sensors are 7 s and 15 s, respectively. The improvement of the recovery time may be attributed to the heterojunction formed between NiO and SnO2. The dynamic response curves of both gas sensors exposed to methanol gas in the concentration range from 0.1 to 100 ppm at 205 °C is shown in Figure 4e. According to the curve, the response increases rapidly and with good linearity to methanol in the range of 1–100 ppm. The minimum concentration that the sensors are capable of detecting is around 1 ppm. Figure 4f also indicates the responses of pure and SnO2-NiO TF sensors to a wide range of 1–1000 ppm vs. a narrow range of 1–100 ppm of methanol. Above 100 ppm, the response of the sensors increases more slowly, which reveals a saturated trend to methanol. The superb linear relationship under 100 ppm indicates that this sensor is better at detecting low concentrations of methanol. Besides, the long-term stability in Figure 4g indicated an excellent performance of both sensors. In order to offer a contrast with other published results, a comparison of methanol gas-sensing characteristics using other compounds in previous reports is listed in Table 1. Evidently, the SnO2/NiO thin film in this work exhibits a remarkable response to 10 ppm methanol, combined with a lower operating temperature and faster recovery time than the others.
Similarly to other semiconductors, the gas-sensing mechanism of the SnO2-NiO composite thin film to methanol shown in Figure 5a relies on the adsorption and desorption of oxygen, which leads to a change in the resistance [33,34,35]. Furthermore, the good selectivity and fast recovery rate are attributed to the unique properties of p-type semiconductors such as NiO [36]. Specifically, the methanol decomposition temperature on the surface of the NiO material is 240 °C [37], which is slightly higher than our operating temperature, while leading to the accumulation of methanol and good selectivity. The improvement of gas sensitivity achieved by SnO2-NiO TF to methanol is probably attributed to a heterojunction effect and is exhibited in Figure 5b. In detail, the p–n heterojunction is formed between NiO and SnO2. In general, the Fermi energy of SnO2 is different from that of NiO, which leads to a temporary relative movement of charge carriers at the interface of both semiconductors until a new balance is achieved. During this process, the space charge layer is generated and the phenomenon of energy band bending in the depletion layer increases the barrier height and exhibits a higher resistance. When exposed to methanol, the adsorbed oxygen will remove electrons from methanol molecules rapidly and previous electrons will be absorbed back into SnO2 and NiO, which results in an increase of the area of electric transport channels. As a result, the resistance of the SnO2-NiO TF gas sensor in methanol decreased and the gas response increased. Furthermore, a great number of cracks on the surface of nickel oxide also provide more reaction sites for methanol.

4. Conclusions

In summary, a highly sensitive gas sensor is realized by decorating an agglomerated block of NiO on an SnO2 thin film, which delivers outstanding performances for the detection of methanol. Herein, XRD, SEM and TEM patterns indicated that NiO nanoparticles have been successfully attached onto the SnO2 films. As for the gas-sensitive performance to methanol, the SnO2 gas sensor exhibits a better selectivity, recovery time and gas response than pure SnO2. The sensing improvement realized by the NiO composite, in addition to special properties of the NiO semiconductor, can be attributed to the heterojunction effect between NiO and SnO2, resulting in an increase of the gas response.

Author Contributions

Conceptualization, T.L. and X.W.; methodology, L.X.; investigation, Q.L. and Q.C.; writing—original draft preparation, T.L.; writing—review and editing, T.L.; providing the gas sensing mechanism, H.S. and C.Z.

Funding

This work is financially supported by National Natural Science Foundation of China (Grant NOs. 61474056 and 51672103).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Choi, K.-I.; Hübner, M.; Haensch, A.; Kim, H.-J.; Weimar, U.; Barsan, N.; Jong, H.-L. Ambivalent effect of Ni loading on gas sensing performance in SnO2 based gas sensor. Sens. Actuators B Chem. 2013, 183, 401–410. [Google Scholar] [CrossRef]
  2. Ge, Q.; Ma, S.Y.; Xu, Y.B.; Xu, X.L.; Chen, H.; Qiang, Z.; Yang, H.M.; Ma, L.; Zeng, Z. Preparation, characterization and gas sensing properties of Pr-doped ZnO/SnO2 nanoflowers. Mater. Lett. 2017, 191, 5–9. [Google Scholar] [CrossRef]
  3. Rongjun, Z.; Xu, Z.; Sijia, P.; Ping, H.; Tong, Z.; Zidong, W.; Xinxin, X.; Yue, Y.; Yude, W. Shaddock peels as bio-templates synthesis of Cd-doped SnO2 nanofibers: A high performance formaldehyde sensing material. J. Alloy. Compd. 2020, 813, 152170. [Google Scholar]
  4. Park, K.-R.; Cho, H.-B.; Lee, J.; Song, Y.; Kim, W.-B.; Choa, Y.-H. Design of highly porous SnO2-CuO nanotubes for enhancing H2S gas sensor performance. Sens. Actuators B Chem. 2020, 302, 127179. [Google Scholar] [CrossRef]
  5. Yinglin, W.; Chang, L.; Lian, W.; Jie, L.; Bo, Z.; Yuan, G.; Peng, S.; Yanfeng, S.; Tong, Z.; Geyu, L. Horseshoe-shaped SnO2 with annulus-like mesoporous for ethanol gas sensing application. Sens. Actuators B Chem. 2017, 240, 1321–1329. [Google Scholar]
  6. Van Hieu, N.; Kim, H.-R.; Ju, B.-K.; Lee, J.-H. Enhanced performance of SnO2 nanowires ethanol sensor by functionalizing with La2O3. Sens. Actuators B Chem. 2008, 133, 228–234. [Google Scholar] [CrossRef]
  7. Kang, W.; Tianyu, Z.; Gang, L.; Qinqin, Y.; Chunhong, L.; Qilong, W.; Deliang, C. Room temperature CO sensor fabricated from Pt-loaded SnO2 porous nanosolid. Sens. Actuators B Chem. 2013, 184, 33–39. [Google Scholar] [CrossRef]
  8. Miao, Y.-E.; He, S.; Zhong, Y.; Yang, Z.; Tjiu, W.W.; Liu, T. A novel hydrogen peroxide sensor based on Ag/SnO2 composite nanotubes by electrospinning. Electrochim. Acta 2013, 99, 117–123. [Google Scholar] [CrossRef]
  9. Shouli, B.; Ke, T.; Hang, F.; Yongjun, F.; Ruixian, L.; Dianqing, L.; Aifan, C.; Chung, C.L. Novel α-Fe2O3/BiVO4 heterojunctions for enhancing NO2 sensing properties. Sens. Actuators B Chem. 2018, 268, 136–143. [Google Scholar]
  10. Shi, S.; Zhang, F.; Lin, H.; Wang, Q.; Shi, E.; Qu, F. Enhanced triethylamine-sensing properties of P-N heterojunction Co3O4/In2O3 hollow microtubes derived from metal–organic frameworks. Sens. Actuators B Chem. 2018, 262, 739–749. [Google Scholar] [CrossRef]
  11. Xiaoguang, S.; Guodong, Z.; Guosheng, W.; Yanbai, S.; Dan, M.; Yajing, Z.; Fanli, M. Assembly of 3D flower-like NiO hierarchical architectures by 2D nanosheets: Synthesis and their sensing properties to formaldehyde. RSC Adv. 2017, 7, 3540–3549. [Google Scholar]
  12. Jahromi, H.S.; Behzad, M. Construction of 0, 1, 2 and 3 dimensional SnO2 nanostructures decorated by NiO nanopetals: Structures, growth and gas-sensing properties. Mater. Chem. Phys. 2018, 207, 489–498. [Google Scholar] [CrossRef]
  13. Zhang, L.; He, J.; Jiao, W. Synthesis and gas sensing performance of NiO decorated SnO2 vertical-standing nanotubes composite thin films. Sens. Actuators B Chem. 2019, 281, 326–334. [Google Scholar] [CrossRef]
  14. Kun, L.; Yinzhen, W.; Mingpeng, C.; Qian, R.; Zhongqi, Z.; Qingju, L.; Jin, Z. High Methanol Gas-Sensing Performance of Sm2O3/ZnO/SmFeO3 Microspheres Synthesized Via a Hydrothermal Method. Nanoscale Res. Lett. 2019, 14, 57. [Google Scholar]
  15. Qing, L.; Tanyuan, W.; Dana, H.; Hanguang, Z.; Ping, X.; Jiantao, H.; Jaephil, C.; Gang, W. High-Performance Direct Methanol Fuel Cells with Precious-Metal-Free Cathode. Adv. Sci. 2016, 3, 1600140. [Google Scholar]
  16. Mengting, Z.; Yuxin, Y.; Chang, L.; Peiyi, H.; Song, H.; Liancheng, Z.; Xianshun, Z. A colorimetric and fluorometric dual-modal sensor for methanol based on a functionalized pentacenequinone derivative. Chem. Commun. 2018, 54, 8339–8342. [Google Scholar]
  17. Korotcenkov, G.; Brinzari, V.; Gulina, L.B.; Cho, B.K. The influence of gold nanoparticles on the conductivity response of SnO2-based thin film gas sensors. Appl. Surf. Sci. 2015, 353, 793–803. [Google Scholar] [CrossRef]
  18. Yadava, L.; Verma, R.; Dwivedi, R. Sensing properties of CdS-doped tin oxide thick film gas sensor. Sens. Actuators B Chem. 2010, 144, 37–42. [Google Scholar] [CrossRef]
  19. Liming, S.; Anatolii, L.; Denys, B.; Haibo, L.; Junkai, Z.; Ming, F.; Liying, L.; Duo, C.; Klyui, N.I. Facile Synthesis of Hierarchical Tin Oxide Nanoflowers with Ultra-High Methanol Gas Sensing at Low Working Temperature. Nanoscale Res. Lett. 2019, 14, 84. [Google Scholar]
  20. Li, Z.; Yi, J. Enhanced ethanol sensing of Ni-doped SnO2 hollow spheres synthesized by a one-pot hydrothermal method. Sens. Actuators B Chem. 2017, 243, 96–103. [Google Scholar] [CrossRef]
  21. Kim, J.; Mirzaei, H.A.; Kim, H.W.; Kim, S.S. Variation of shell thickness in ZnO-SnO2 core-shell nanowires for optimizing sensing behaviors to CO, C6H6, and C7H8 gases. Sens. Actuators B Chem. 2020, 302, 127150. [Google Scholar] [CrossRef]
  22. Sikai, Z.; Yanbai, S.; Pengfei, Z.; Jin, Z.; Wei, Z.; Xiangxiang, C.; Dezhou, W.; Ping, F.; Yansong, S. Highly selective NO2 sensor based on p-type nanocrystalline NiO thin films prepared by sol–gel dip coating. Ceram. Int. 2018, 44, 753–759. [Google Scholar]
  23. Xianghong, L.; Jun, Z.; Liwei, W.; Taili, Y.; Xianzhi, G.; Shihua, W.; Shurong, W. 3D hierarchically porous ZnO structures and their functionalization by Aunanoparticles for gas sensors. J. Mater. Chem. 2011, 21, 349–356. [Google Scholar]
  24. Li, C.C.; Yin, X.M.; Li, Q.H.; Wang, T.H. Enhanced gas sensing properties of ZnO/SnO2 hierarchical architectures by glucose-induced attachment. CrystEngComm 2011, 13, 1557–1563. [Google Scholar] [CrossRef]
  25. Tang, W.; Wang, J.; Yao, P.; Li, X. Hollow hierarchical SnO2-ZnO composite nanofibers with heterostructure based on electrospinning method for detecting methanol. Sens. Actuators B Chem. 2014, 192, 543–549. [Google Scholar] [CrossRef]
  26. Yang, H.M.; Ma, S.Y.; Yang, G.J.; Jin, W.X.; Wang, T.T.; Jiang, X.H.; Li, W.Q. High sensitive and low concentration detection of methanol by a gas sensor based on one-step synthesis α-Fe2O3 hollow spheres. Mater. Lett. 2016, 169, 73–76. [Google Scholar] [CrossRef]
  27. Tang, W. Sensing mechanism of SnO2/ZnO nanofibers for CH3OH sensors: Heterojunction effects. J. Phys. D Appl. Phys. 2017, 50, 475105. [Google Scholar] [CrossRef]
  28. Ji, H.-F.; Liu, W.-K.; Li, S.; Li, Y.; Shi, Z.-F.; Tian, T.Y.; Li, X.-J. High-performance methanol sensor based on GaN nanostructures grown on silicon nanoporous pillar array. Sens. Actuators B Chem. 2017, 250, 518–524. [Google Scholar] [CrossRef]
  29. Fang, W.; Hairong, L.; Zhaoxin, Y.; Yongzhe, S.; Fangzhi, C.; Heng, D.; Longzhen, X.; Haiyan, L. A highly sensitive gas sensor based on CuO nanoparticles synthetized via a sol–gel method. RSC Adv. 2016, 6, 79343–79349. [Google Scholar]
  30. Ding, X.; Zeng, D.; Xie, C. Controlled growth of SnO2 nanorods clusters via Zn doping and its influence on gas-sensing properties. Sens. Actuators B Chem. 2010, 149, 336–344. [Google Scholar] [CrossRef]
  31. Qin, J.; Cui, Z.; Yang, X.; Zhu, S.; Li, Z.; Liang, Y. Synthesis of three-dimensionally ordered macroporous LaFeO3 with enhanced methanol gas sensing properties. Sens. Actuators B Chem. 2015, 209, 706–713. [Google Scholar] [CrossRef]
  32. Wang, L.L.; Li, Z.J.; Luo, L.; Zhao, C.Z.; Kang, L.P.; Liu, D.W. Methanol sensing properties of honeycomb-like SnO2 grown on silicon nanoporous pillar array. J. Alloy. Compds. 2016, 682, 170–175. [Google Scholar] [CrossRef]
  33. Cai, Z.-X.; Li, H.-Y.; Ding, J.-C.; Guo, X. Hierarchical flowerlike WO3 nanostructures assembled by porous nanoflakes for enhanced NO gas sensing. Sens. Actuators B Chem. 2017, 246, 225–234. [Google Scholar] [CrossRef]
  34. Chen, H.; Wang, Q.; Kou, C.; Sui, Y.; Zeng, Y.; Du, F. One-pot synthesis and improved sensing properties of hierarchical flowerlike SnO2 assembled from sheet and ultra-thin rod subunits. Sens. Actuators B Chem. 2014, 194, 447–453. [Google Scholar] [CrossRef]
  35. Li, L.; Yu, Z.; Guoguang, W.; Shouchun, L.; Lianyuan, W.; Yu, H.; Xiaoxue, J.; Aiguo, W. High toluene sensing properties of NiO-SnO2 composite nanofiber sensors operating at 330 °C. Sens. Actuators B Chem. 2011, 160, 448–454. [Google Scholar] [CrossRef]
  36. Geng, X.; Lahem, D.; Zhang, C.; Li, C.-J.; Olivier, M.-G.; Debliquy, M. Visible light enhanced black NiO sensors for ppb-level NO2 detection at room temperature. Ceram. Int. 2019, 45, 4253–4261. [Google Scholar] [CrossRef]
  37. Badlani, M.; Wachs, I.E. Methanol: A “Smart” Chemical Probe Molecule. Catal. Lett. 2001, 75, 137. [Google Scholar] [CrossRef]
Figure 1. Schematic drawing of (a) as-prepared sensor device (b) measuring system of gas sensors.
Figure 1. Schematic drawing of (a) as-prepared sensor device (b) measuring system of gas sensors.
Crystals 09 00621 g001
Figure 2. XRD pattern of the as-prepared SnO2-NiO thin film (TF), ceramic substrate and NiO powders.
Figure 2. XRD pattern of the as-prepared SnO2-NiO thin film (TF), ceramic substrate and NiO powders.
Crystals 09 00621 g002
Figure 3. FESEM images of (a) the ceramic substrate (bd) pure SnO2 on the ceramic substrate and (eg) SnO2-NiO TF on the ceramic substrate; TEM images of (h,i) the SnO2-NiO composite.
Figure 3. FESEM images of (a) the ceramic substrate (bd) pure SnO2 on the ceramic substrate and (eg) SnO2-NiO TF on the ceramic substrate; TEM images of (h,i) the SnO2-NiO composite.
Crystals 09 00621 g003
Figure 4. (a) Responses of pure and SnO2-NiO TF sensors at various working temperatures to 10 ppm methanol, (b) selectivity to 10 ppm various gases, (c) seven reversible cycles to 10 ppm methanol at 225 °C, (d) dynamic response and recover property to 10 ppm methanol, (e) the dynamic response change toward 1–100 ppm methanol, (f) the response of sensors to 1–1000 ppm methanol and fitting line of the response in range of 1–100 ppm methanol, (g) long-term stability of sensors to 10 ppm methanol.
Figure 4. (a) Responses of pure and SnO2-NiO TF sensors at various working temperatures to 10 ppm methanol, (b) selectivity to 10 ppm various gases, (c) seven reversible cycles to 10 ppm methanol at 225 °C, (d) dynamic response and recover property to 10 ppm methanol, (e) the dynamic response change toward 1–100 ppm methanol, (f) the response of sensors to 1–1000 ppm methanol and fitting line of the response in range of 1–100 ppm methanol, (g) long-term stability of sensors to 10 ppm methanol.
Crystals 09 00621 g004
Figure 5. (a) The changes of the space charge layer of SnO2-NiO TF when exposed to air and methanol. (b) The energy band diagram of contact between NiO and SnO2 in air and methanol.
Figure 5. (a) The changes of the space charge layer of SnO2-NiO TF when exposed to air and methanol. (b) The energy band diagram of contact between NiO and SnO2 in air and methanol.
Crystals 09 00621 g005
Table 1. Comparison of methanol gas-sensing characteristics using other compounds in previous reports.
Table 1. Comparison of methanol gas-sensing characteristics using other compounds in previous reports.
Sensing MaterialConcentration (ppm)Temperature (°C)ResponseResponse/Recovery TimeReference
Au decorated ZnO103002.7~20 s/~15 s[23]
ZnO-SnO2 nanostructure microspheres10300~2~18 s/~35 s[24]
SnO2-ZnO composites nanofibers103508.520 s/40 s[25]
α-Fe2O3 hollow spheres1028025.18 s/9 s[26]
SnO2/ZnO nanofibers10350~8.9~12 s/~31 s[27]
GaN nanostructures10350~1.3~15 s/~30 s[28]
CuO nanoparticles102205.913 s/13 s[29]
Zn doping SnO2 nanorods clusters1027011.5~10 s/~10 s[30]
three-dimensionally LaFeO310190~2~12 s/~15 s[31]
honeycomb-like SnO210320~4~20 s/~20 s[32]
SnO2/NiO thin film1022515.1218 s/7 sThis work

Share and Cite

MDPI and ACS Style

Liu, T.; Xu, L.; Wang, X.; Li, Q.; Cui, Q.; Suo, H.; Zhao, C. A Simple Dip-Coating Method of SnO2-NiO Composite Thin Film on a Ceramic Tube Substrate for Methanol Sensing. Crystals 2019, 9, 621. https://doi.org/10.3390/cryst9120621

AMA Style

Liu T, Xu L, Wang X, Li Q, Cui Q, Suo H, Zhao C. A Simple Dip-Coating Method of SnO2-NiO Composite Thin Film on a Ceramic Tube Substrate for Methanol Sensing. Crystals. 2019; 9(12):621. https://doi.org/10.3390/cryst9120621

Chicago/Turabian Style

Liu, Tianyu, Longxiao Xu, Xiangyue Wang, Qiuyang Li, Qiang Cui, Hui Suo, and Chun Zhao. 2019. "A Simple Dip-Coating Method of SnO2-NiO Composite Thin Film on a Ceramic Tube Substrate for Methanol Sensing" Crystals 9, no. 12: 621. https://doi.org/10.3390/cryst9120621

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