Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 2822 2023-04-29 05:02:56 |
2 format correct + 16 word(s) 2838 2023-05-04 04:54:38 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Mirzaei, A.; Ansari, H.R.; Shahbaz, M.; Kim, J.; Kim, H.W.; Kim, S.S. Metal Oxide Semiconductor Gas Sensors with Different Morphologies. Encyclopedia. Available online: https://encyclopedia.pub/entry/43644 (accessed on 27 July 2024).
Mirzaei A, Ansari HR, Shahbaz M, Kim J, Kim HW, Kim SS. Metal Oxide Semiconductor Gas Sensors with Different Morphologies. Encyclopedia. Available at: https://encyclopedia.pub/entry/43644. Accessed July 27, 2024.
Mirzaei, Ali, Hamid Reza Ansari, Mehrdad Shahbaz, Jin-Young Kim, Hyoun Woo Kim, Sang Sub Kim. "Metal Oxide Semiconductor Gas Sensors with Different Morphologies" Encyclopedia, https://encyclopedia.pub/entry/43644 (accessed July 27, 2024).
Mirzaei, A., Ansari, H.R., Shahbaz, M., Kim, J., Kim, H.W., & Kim, S.S. (2023, April 29). Metal Oxide Semiconductor Gas Sensors with Different Morphologies. In Encyclopedia. https://encyclopedia.pub/entry/43644
Mirzaei, Ali, et al. "Metal Oxide Semiconductor Gas Sensors with Different Morphologies." Encyclopedia. Web. 29 April, 2023.
Metal Oxide Semiconductor Gas Sensors with Different Morphologies
Edit

There is an increasing need for the development of low-cost and highly sensitive gas sensors for environmental, commercial, and industrial applications in various areas, such as hazardous gas monitoring, safety, and emission control in combustion processes. Considering this, resistive-based gas sensors using metal oxide semiconductors (MOSs) have gained special attention owing to their high sensing performance, high stability, and low cost of synthesis and fabrication. The relatively low final costs of these gas sensors allow their commercialization; consequently, they are widely used and available at low prices. Different morphologies of metal oxide gas sensors are discussed.

semiconductor metal oxide morphology gas sensor sensing mechanism

1. Quantum-Dot-Based Gas Sensors Using Metal Oxide Semiconductors

Morphology engineering is an important approach for enhancing the gas-sensing features of metal oxide semiconductor (MOS)-based gas sensors, and MOS materials with different morphologies have been explored for gas-sensing purposes. The main goal of using novel morphologies for sensing studies is to increase the surface area. The larger the surface area, the greater the availability of adsorption sites, resulting in the adsorption of more gas molecules on the surface of the gas sensor, along with the appearance of a higher sensing signal. For sensing studies, zero-, one-, two-, and three-dimensional morphologies have been explored. Quantum dots (QDs) are zero-dimensional (0D) semiconductor nanoparticles, where the motion of charge carriers is confined in all three directions. Thus, the energy difference between energy bands changes depending on the size of the particle, and it is possible to tune the band gap of QDs by changing their size. Accordingly, the electrical properties of QDs can be engineered by controlling the dot size [1]. Owing to the extremely small size of QDs, when the QDs are exposed to air, they are depleted of electrons, and subsequent exposure to the target gas leads to extensive modulation of the electrical resistance, leading to a high sensing signal. In addition, owing to their large resistance variation and good electrical properties, most QD-based gas sensors can work at low temperatures.

2. Nanowire-Based Nanomaterial Gas Sensors

Among different morphologies, nanowires (NWs) with aspect ratios greater than 20 [2] are highly popular for sensing applications, owing to their high surface area, straightforward preparation methods, ease of gas sensor fabrication, high and rapid response to the target gas, long-term stability, and high crystallinity [3][4]. Accordingly, several studies have realized the enhanced gas-sensing capacity of NW-based gas sensors [5][6]. In [7], various methods to synthesize NWs are discussed in detail.
Among different MOSs with NW morphology, SnO2 NWs are very popular for sensing purposes [8][9][10][11]. SnO2 has a band gap of 3–4 eV, depending on the synthesis method, size, surface states, and impurities [12]. However, most studies have reported a band gap of 3.7 eV for this MOS [13]. SnO2 is the most stable n-type oxide, and is extensively used in the preparation of semiconductor gas sensors. Its high electron mobility (~160 cm2/V·s) makes it a key choice for sensing applications [14][15]. In addition, SnO2 NWs exhibit good optical transparency and conductivity [16]. Wang et al. [17] performed one of the first studies on SnO2 NWs for the detection of H2 gas, and compared the results with those of SnO2 nanorods (NRs). It was reported that owing to the higher surface area of NWs, size effects when the diameter of the NWs was less than the Debye length of SnO2, and the small electrode gap of the SnO2 NW gas sensor, a larger response was observed for the SnO2 NWs at 300 °C than that for SnO2 NRs. Notably, not only SnO2 NWs, but also other MOS NWs, have been employed as gas sensors [18][19]. The construction of n–n and p–n junctions is a commonly employed strategy for improving the sensitivity of NW-based gas sensors [20]. Owing to the generation of these heterojunctions, the resistance of gas sensors is higher than that of the pure sensing material, and the subsequent modulation of resistance in the presence of target gases generates a high sensing signal for heterojunction gas sensors [21].
Branched NWs (e.g., nanoforests or nanotrees) are a special type of NWs that have recently been used for sensing applications. Their 3D morphology and numerous homo- or heterojunctions make the branching NWs highly popular for sensing studies. With high surface area, branched NWs can detect extremely low amounts of gases. Branched NWs also offer an increased number of carrier paths and improved conduction between the NW branches and backbones [22][23][24].

3. Nanofiber-Based Metal Oxide Semiconductor Gas Sensors

Nanofibers (NFs) are another type of 1D material used in MOS gas sensors. Similar to NWs, NFs have an aspect ratio of more than 20. However, their surfaces have many nanograins, which result in a further increase in the surface area [25]. There are various methods to synthesize NFs; however, NFs are commonly produced using the electrospinning (ES) technique. The ES technique consists of a spinneret section, a high-voltage power supply, and a metal collector [26]. During electrospinning, NFs are formed from a liquid jet and subsequently elongated under a high voltage. The NF formation process consists of the following: (i) onset of jetting and development of a rectilinear jet, (ii) bending deformation along with solvent evaporation to form solidified NFs, and (iii) NF collection [27]. In a simplified view, a sol is initially prepared in this technique, and is subsequently poured into a syringe. Upon applying a high voltage to the sol with a defined viscosity, it is possible to collect extremely fine and long NFs on the collector placed at a fixed distance from the syringe. After calcination of the synthesized NFs at an appropriate temperature for sufficient time, the final diameter decreases due to evaporation of the solvent and organic materials [26].
By optimization of the ES technique parameters—such as the applied voltage, distance between the needle tip and the collector, flow rate, and temperature—it is possible to tune the diameter of the resulting NFs [28][29]. In particular, the viscosity of the solution can significantly affect the formation of long NFs [30]. If the viscosity of the ES solution is not adjusted to the desired value, NFs with beads are formed [30].
Furthermore, with a special type of ES technique, known as coaxial ES, it is possible to directly synthesize C–S NFs [31][32]. Many MOSs have been successfully synthesized using the ES technique for gas-sensing studies [33]. Similar to C–S NWs, the thickness of the shell layer should be adjusted to obtain the maximum response to the target gas. For example, Kim et al. [34] reported that in SnO2–Cu2O C–S NF gas sensors with different shell thicknesses, the sensor with a shell thickness of 30 nm gave the highest response to CO gas—it showed a response to 5–10 ppm of CO gas at 300 °C.

4. Nanotube-Based Metal Oxide Semiconductor Gas Sensors

Metal oxide nanotubes (NTs) have been studied less compared to their NW and NF counterparts [35]. This may be due to the difficulty in the synthesis of MOS NTs. Various techniques can be used for synthesizing MOS NTs. However, the most common method is the use of a hard or soft sacrificial template. Hard templates have been prepared using silica, carbon, and polystyrene beads, while soft templates have been fabricated from bubbles, polymer vesicles, emulsion droplets, and surfactant micelles. The core can be removed by dissolving in a solvent or calcination to generate hollow structures [36]. The main features of gas sensors based on hollow structures such as NTs are (i) wall permeability, (ii) wall thickness, and (iii) wall chemistry [37]. Although thinner NTs are expected to provide greater surface area and more adsorption sites for gas molecules [38], the presence of defects can sometimes alter the sensing properties. Hazra et al. investigated the effects of shell thickness on the sensing characteristics of TiO2 NTs. Interestingly, the sensor with the thinnest wall and largest surface area did not show the highest response to the target gas, due to the lower amount of oxygen vacancies. The device with a thicker wall and sufficient oxygen vacancies exhibited the highest response [39]. In conclusion, MOS-based NTs for gas-sensing studies offer a high surface area owing to the hollowness of their structure, which can provide many adsorption sites for gas molecules.

5. Nanorod-Based Metal Oxide Semiconductor Gas Sensors

Nanorod-based gas sensors are another popular category of MOS materials for gas-sensing studies, owing to their unique electrical properties and high surface area provided by the NR morphology. The hydrothermal method provides a convenient, fast, versatile, and low-cost route for constructing well-ordered nanorod arrays [40][41][42][43]. However, other methods—such as chemical bath deposition (CBD), which is economical, straightforward, and scalable—can also be used to prepare NRs from different materials [44]. Chemical vapor deposition [45][46], metal–organic chemical vapor deposition [47][48], and pulsed laser deposition [49] are other methods for obtaining NRs in MOSs. NR-based MOS materials are popular for sensing studies because of their ease of production, high surface area, and compositional versatility.

6. Nanosheet-Based Metal Oxide Semiconductor Gas Sensors

Two-dimensional MOS nanostructured materials with a sheet-like morphology are popular for gas-sensing studies due to their high surface area. Only a few 2D MOS (MoO3, WO3, and SnO2) analogs of graphene are known. Two-dimensional (2D) layered MOSs have strong in-plane bonds and weak van der Waals forces between layers, and gas molecules can be adsorbed on these sites. Thus, 2D layered MOSs can be employed for the production of gas sensors [50]. Apart from the naturally 2D layered MOSs, other MOSs can also be synthesized in a nanosheet-like morphology. For instance, Choi et al. [51] synthesized porous ZnO nanosheets using a solvothermal technique for NO2 detection. The sensor exhibited a high response to 74.68–10 ppm of NO2 gas at 200 °C. The width of the EDL increased in the presence of NO2 gas, contributing to the sensing signal. Moreover, homojunctions were formed, and changes in the height of the homojunctions led to resistance modulation in the presence of NO2 gas. Furthermore, the highly porous morphology and high specific surface area (11.51 m2/g) contributed to the increased availability of active sites for NO2 gas molecules [51].

7. Three-Dimensional Metal Oxide Semiconductor Gas Sensors

Hierarchical nanostructures with three-dimensional (3D) morphology are composed of zero-, one-, two-, and three-dimensional morphologies, such as nanorods, nanotubes, or nanosheets [52]. Hierarchical nanostructures with high surface area provide more adsorption sites for target gases. In addition, with pores or channels in their structure, the diffusion of gases is facilitated, resulting in better interaction of the target gas with deep parts of the sensing material [53]. Vapor-phase growth and hydrothermal techniques are the two most-used methods to synthesize hierarchical nanostructures [52]. Owing to these advantages, hierarchical morphologies of MOSs are widely used for sensing studies [54][55].
For instance, Guo et al. [56] used a 3D hierarchical hollow structure of CuO/WO3 for p-xylene-sensing studies. The sensor showed a good response to 6.36–50 ppm of xylene gas at 260 °C. The sensing mechanism was related to the unique 3D hierarchical structure with a high surface area of 23.4962 m2·g−1, the formation of defects in the interfaces between CuO and WO3, and the formation of p–n heterojunctions between CuO and WO3. Recently, Yu et. al. [57] reviewed MOSs with hierarchical structures for gas-sensing studies.

8. Noble-Metal-Decorated Metal Oxide Semiconductor Gas Sensors

Gas-sensing properties can be improved by loading MOSs with small amounts of appropriately chosen noble metals. Noble metals such as Au, Pt, Pd, Ag, and Ru are widely used for the decoration or functionalization of the MOS surfaces to improve the overall performance of the resulting gas sensors.
It should be noted that the noble metals need to be dispersed as finely as possible on the surface of the MOSs. Agglomeration of noble metals on the surface of the sensing layer can lead to a poor response of the gas sensor. Generally, noble metals affect the gas-sensing performance via two well-known mechanisms: The chemical sensitization mechanism, which occurs via the spillover effect, is a commonly known phenomenon in catalytic chemistry. In this case, the noble metal activates the target gas to facilitate catalytic oxidation of gas. Noble metals increase the gas sensitivity as they increase the rate of chemical processes [58]. In electronic sensitization, because noble metals generally have a larger work function than MOSs, electrons are transferred from the MOS to the noble metals, leading to the formation of Schottky barriers and contraction of the EDL or HAL inside the MOS. In the target gas atmosphere, the width of the EDL and the HAL changes, leading to significant modulation of the sensor resistance. Among noble metals, Pd is well known for H2 gas detection [59], owing to its excellent catalytic activation ability for hydrogen through the hydrogen spillover effect [60]. When a sensor decorated with Pd NPs is placed in an atmosphere containing H2 molecules, hydrogen can easily dissociate into hydrogen atoms on the Pd surface; in the spillover effect, hydrogen atoms move to the neighboring MOS surfaces, leading to additional reactions between atomic hydrogen and adsorbed oxygen [61]. Other noble metals also exhibit good catalytic activities toward different gases. For example, Ag is often used for H2S detection because of the generation of Ag2S upon exposure to H2S gas [62]. Pt and Pd also have good catalytic activities toward C7H8 and C6H6 gases, respectively [63][64].

9. Hybrid Metal Oxide Semiconductor Gas Sensors

A good strategy to enhance the gas-sensing characteristics of MOS-based gas sensors is using another material in combination with the MOS. Hybrid MOS nanocomposites are either (i) composites of MOS and carbon materials such as graphene, graphene oxide (GO), reduced graphene oxide (rGO), and carbon nanotubes, or (ii) composites of MOS and conductive polymers (CPs). Graphene, which comprises sp2 carbon atoms, is utilized in gas sensors due to its high charge carrier mobility (200,000 cm2 V−1 s−1) and large surface area (2630 m2 g −1) [65]. The first ever graphene gas sensor was introduced in 2007 [66]. Owing to its single-layer or few-layer nature, graphene can even interact with a single molecule. Pristine graphene can easily agglomerate owing to the surface interactions. Furthermore, graphene has no band gap, hindering its gas-sensing usage [67]. Thus, rGO, with its many functional groups—such as -OH and –O—as well as defects, is a better choice for gas-sensing studies. Although GO can be used for sensing applications, it has a very high resistance owing to the presence of oxygen-based functional groups, limiting its use in sensing studies [68]. rGO can be synthesized from GO using chemical reduction, thermal reduction, and UV light reduction methods [69][70]. Many papers on gas sensors with hybrids of MOSs and carbon allotropes have been reported. However, for sensing studies, rGO is the most effective allotrope of carbon in its hybrid form. This is because of its large surface area, presence of many defects, and high concentrations of charge carriers with high mobility—all of which are beneficial for sensing. Accordingly, several studies have been reported on hybrid MOS–rGO composites with the ability to operate at room temperature [71][72].
CPs have tunable conductivity and high flexibility in synthesis and processing. However, because of the high affinity of CPs for moisture, they are unstable, and generally exhibit poor sensitivity and selectivity for different gases in their pristine form. The use of hybrid nanocomposites with CPs and MOSs could result in the development of room-temperature gas sensors [73]. Accordingly, hybrids of CPs and MOSs have been used to enhance the sensitivity of nanostructured sensors at low or room temperature [74][75]. One application of hybrids of MOSs and CPs is the development of flexible and wearable gas sensors. Flexible gas sensors are generally deposited on flexible substrates, and need to work at low temperatures, e.g., room temperature. Since composites of MOs and CPs can function at low temperatures, they are good candidates for such applications.

10. Comparison of Performance of Gas Sensors with Different Morphologies

Table 1 summarizes the performance of gas sensors with various morphologies. Different morphologies and compositions have been used for the detection of various gases at different temperatures. Moreover, pristine gas sensors show lower responses and higher sensing temperatures, while composite or decorated sensing materials show better performance at lower temperatures.
Table 1. Gas-sensing properties of different sensors with various morphologies.

References

  1. Vardan, G. Quantum dots: Perspectives in next-generation chemical gas sensors—A review. Anal. Chim. Acta 2021, 1152, 238192.
  2. Murphy, C.J.; Jana, N.R. Controlling the aspect ratio of inorganic nanorods and nanowires. Adv. Mater. 2002, 14, 80–82.
  3. Comini, E.; Sberveglieri, G. Metal oxide nanowires as chemical sensors. Mater. Today 2010, 13, 36–44.
  4. Ramgir, N.; Datta, N.; Kaur, M.; Kailasaganapathi, S.; Debnath, A.K.; Aswal, D.K.; Gupta, S.K. Metal oxide nanowires for chemiresistive gas sensors: Issues, challenges and prospects. Colloids Surf. A Physicochem. Eng. Asp. 2013, 439, 101–116.
  5. Yuan, K.; Wang, C.-Y.; Zhu, L.-Y.; Cao, Q.; Yang, J.-H.; Li, X.-X.; Huang, W.; Wang, Y.-Y.; Lu, H.-L.; Zhang, D.W. Fabrication of a micro-electromechanical system-based acetone gas sensor using CeO2 nanodot-decorated WO3 nanowires. ACS Appl. Mater. Interfaces 2020, 12, 14095–14104.
  6. Tonezzer, M.; Kim, J.-H.; Lee, J.-H.; Iannotta, S.; Kim, S.S. Predictive gas sensor based on thermal fingerprints from Pt-SnO2 nanowires. Sens. Actuators B Chem. 2019, 281, 670–678.
  7. Galstyan, V.; Moumen, A.; Kumarage, G.W.; Comini, E. Progress towards chemical gas sensors: Nanowires and 2D semiconductors. Sens. Actuators B-Chem. 2022, 357, 131466.
  8. Park, S.; Jung, Y.W.; Ko, G.M.; Jeong, D.Y.; Lee, C. Enhanced NO2 gas sensing performance of the In2O3-decorated SnO2 nanowire sensor. Appl. Phys. A 2021, 127, 898.
  9. Van Tong, P.; Minh, L.H.; Van Duy, N.; Hung, C.M. Porous In2O3 nanorods fabricated by hydrothermal method for an effective CO gas sensor. Mater. Res. Bull. 2021, 137, 111179.
  10. Choi, Y.-J.; Hwang, I.-S.; Park, J.-G.; Choi, K.J.; Park, J.-H.; Lee, J.-H. Novel fabrication of an SnO2 nanowire gas sensor with high sensitivity. Nanotechnology 2008, 19, 95508.
  11. Hoa, N.D.; Le, D.T.T.; Tam, P.D.; Le, A.-T.; Van Hieu, N. On-chip fabrication of SnO2-nanowire gas sensor: The effect of growth time on sensor performance. Sens. Actuators B Chem. 2010, 146, 361–367.
  12. Arlinghaus, F.J. Energy bands in stannic oxide (SnO2). J. Phys. Chem. Solids 1974, 35, 931–935.
  13. Yadav, K.K.; Guchhait, S.K.; Sood, K.; Mehta, S.K.; Ganguli, A.K.; Jha, M. Mechanistic insights of enhanced photocatalytic efficiency of SnO2-SnS2 heterostructures derived from partial sulphurization of SnO2. Sep. Purif. Technol. 2020, 242, 116835.
  14. Yamazoe, N.; Sakai, G.; Shimanoe, K. Oxide semiconductor gas sensors. Catal. Surv. Asia 2003, 7, 63–75.
  15. Gondal, M.A.; Drmosh, Q.A.; Saleh, T.A. Preparation and characterization of SnO2 nanoparticles using high power pulsed laser. Appl. Surf. Sci. 2010, 256, 7067–7070.
  16. Das, S.; Jayaraman, V. SnO2: A comprehensive review on structures and gas sensors. Prog. Mater. Sci. 2014, 66, 112–255.
  17. Wang, B.; Zhu, L.F.; Yang, Y.H.; Xu, N.S.; Yang, G.W. Fabrication of a SnO2 nanowire gas sensor and sensor performance for hydrogen. J. Phys. Chem. C 2008, 112, 6643–6647.
  18. Ahn, M.-W.; Park, K.-S.; Heo, J.-H.; Park, J.-G.; Kim, D.-W.; Choi, K.J.; Lee, J.-H.; Hong, S.-H. Gas sensing properties of defect-controlled ZnO-nanowire gas sensor. Appl. Phys. Lett. 2008, 93, 263103.
  19. Yi, S.; Tian, S.; Zeng, D.; Xu, K.; Zhang, S.; Xie, C. An In2O3 nanowire-like network fabricated on coplanar sensor surface by sacrificial CNTs for enhanced gas sensing performance. Sens. Actuators B Chem. 2013, 185, 345–353.
  20. Kaur, N.; Zappa, D.; Maraloiu, V.; Comini, E. Novel Christmas Branched Like NiO/NiWO4/WO3 (p–p–n) Nanowire Heterostructures for Chemical Sensing. Adv. Funct. Mater. 2021, 31, 2104416.
  21. Liu, H.; Wang, Z.; Cao, G.; Pan, G.; Yang, X.; Qiu, M.; Sun, C.; Shao, J.; Li, Z.; Zhang, H. Construction of hollow NiO/ZnO pn heterostructure for ultrahigh performance toluene gas sensor. Mater. Sci. Semicond. Process. 2022, 141, 106435.
  22. Cheng, C.; Fan, H.J. Branched nanowires: Synthesis and energy applications. Nano Today 2012, 7, 327–343.
  23. Wan, Q.; Huang, J.; Xie, Z.; Wang, T.; Dattoli, E.N.; Lu, W. Branched SnO2 nanowires on metallic nanowire backbones for ethanol sensors application. Appl. Phys. Lett. 2008, 92, 102101.
  24. An, S.; Park, S.; Ko, H.; Jin, C.; Lee, W.I.; Lee, C. Enhanced gas sensing properties of branched ZnO nanowires. Thin Solid Film. 2013, 547, 241–245.
  25. Kou, X.; Xie, N.; Chen, F.; Wang, T.; Guo, L.; Wang, C.; Wang, Q.; Ma, J.; Sun, Y.; Zhang, H. Superior acetone gas sensor based on electrospun SnO2 nanofibers by Rh doping. Sens. Actuators B Chem. 2018, 256, 861–869.
  26. Li, Y.; Zhu, J.; Cheng, H.; Li, G.; Cho, H.; Jiang, M.; Gao, Q.; Zhang, X. Developments of advanced electrospinning techniques: A critical review. Adv. Mater. Technol. 2021, 6, 2100410.
  27. Rodríguez-Tobías, H.; Morales, G.; Grande, D. Comprehensive review on electrospinning techniques as versatile approaches toward antimicrobial biopolymeric composite fibers. Mater. Sci. Eng. C 2019, 101, 306–322.
  28. Abideen, Z.U.; Kim, J.-H.; Lee, J.-H.; Kim, J.-Y.; Mirzaei, A.; Kim, H.W.; Kim, S.S. Electrospun metal oxide composite nanofibers gas sensors: A review. J. Korean Ceram. Soc. 2017, 54, 366–379.
  29. Islam, M.S.; Ang, B.C.; Andriyana, A.; Afifi, A.M. A review on fabrication of nanofibers via electrospinning and their applications. SN Appl. Sci. 2019, 1, 1–16.
  30. Haider, A.; Haider, S.; Kang, I.-K. A comprehensive review summarizing the effect of electrospinning parameters and potential applications of nanofibers in biomedical and biotechnology. Arab. J. Chem. 2018, 11, 1165–1188.
  31. Tsai, P.; Chen, J.; Ercan, E.; Chueh, C.; Tung, S.; Chen, W. Uniform luminous perovskite nanofibers with color-tunability and improved stability prepared by one-step core/shell electrospinning. Small 2018, 14, 1704379.
  32. Huang, B.; Zhang, Z.; Zhao, C.; Cairang, L.; Bai, J.; Zhang, Y.; Mu, X.; Du, J.; Wang, H.; Pan, X. Enhanced gas-sensing performance of In2O3 shell nanofibers prepared by coaxial electrospinning. Sens. Actuators B Chem. 2018, 255, 2248–2257.
  33. Ding, B.; Wang, M.; Yu, J.; Sun, G. Gas sensors based on electrospun nanofibers. Sensors 2009, 9, 1609–1624.
  34. Kim, J.H.; Lee, J.H.; Kim, J.Y.; Mirzaei, A.; Kim, H.W.; Kim, S.S. Enhancement of CO and NO2 sensing in n-SnO2-p-Cu2O core-shell nanofibers by shell optimization. J. Hazard. Mater. 2019, 376, 68–82.
  35. Cho, I.; Kang, K.; Yang, D.; Yun, J.; Park, I. Localized liquid-phase synthesis of porous SnO2 nanotubes on MEMS platform for low-power, high performance gas sensors. ACS Appl. Mater. Interfaces 2017, 9, 27111–27119.
  36. Jankiewicz, B.J.; Jamiola, D.; Choma, J.; Jaroniec, M. Silica–metal core–shell nanostructures. Adv. Colloid Interface Sci. 2012, 170, 28–47.
  37. Lee, J.-H. Gas sensors using hierarchical and hollow oxide nanostructures: Overview. Sens. Actuators B Chem. 2009, 140, 319–336.
  38. Choi, H.-J.; Kwon, S.-H.; Lee, W.-S.; Im, K.-G.; Kim, T.-H.; Noh, B.-R.; Park, S.; Oh, S.; Kim, K.-K. Ultraviolet photoactivated room temperature NO2 gas sensor of ZnO hemitubes and nanotubes covered with TiO2 nanoparticles. Nanomaterials 2020, 10, 462.
  39. Hazra, A.; Bhowmik, B.; Dutta, K.; Chattopadhyay, P.P.; Bhattacharyya, P. Stoichiometry, length, and wall thickness optimization of TiO2 nanotube array for efficient alcohol sensing. ACS Appl. Mater. Interfaces 2015, 7, 9336–9348.
  40. Liu, F.-T.; Gao, S.-F.; Pei, S.-K.; Tseng, S.-C.; Liu, C.-H.J. ZnO nanorod gas sensor for NO2 detection. J. Taiwan Inst. Chem. Eng. 2009, 40, 528–532.
  41. Rai, P.; Song, H.-M.; Kim, Y.-S.; Song, M.-K.; Oh, P.-R.; Yoon, J.-M.; Yu, Y.-T. Microwave assisted hydrothermal synthesis of single crystalline ZnO nanorods for gas sensor application. Mater. Lett. 2012, 68, 90–93.
  42. Wei, Y.; Hu, M.; Yan, W.; Wang, D.; Yuan, L.; Qin, Y. Hydrothermal synthesis porous silicon/tungsten oxide nanorods composites and their gas-sensing properties to NO2 at room temperature. Appl. Surf. Sci. 2015, 353, 79–86.
  43. Da Silva, L.F.; Catto, A.C.; Avansi, W., Jr.; Cavalcante, L.S.; Mastelaro, V.R.; Andres, J.; Aguir, K.; Longo, E. Acetone gas sensor based on α-Ag2WO4 nanorods obtained via a microwave-assisted hydrothermal route. J. Alloys Compd. 2016, 683, 186–190.
  44. Zhang, H.; Feng, J.; Wang, J.; Zhang, M. Preparation of ZnO nanorods through wet chemical method. Mater. Lett. 2007, 61, 5202–5205.
  45. Liu, Y.; Liu, M. Ordered ZnO nanorods synthesized by combustion chemical vapor deposition. J. Nanosci. Nanotechnol. 2007, 7, 4529–4533.
  46. Chien, F.S.-S.; Wang, C.-R.; Chan, Y.-L.; Lin, H.-L.; Chen, M.-H.; Wu, R.-J. Fast-response ozone sensor with ZnO nanorods grown by chemical vapor deposition. Sens. Actuators B Chem. 2010, 144, 120–125.
  47. Cha, H.G.; Kim, C.W.; Kim, Y.H.; Jung, M.H.; Ji, E.S.; Das, B.K.; Kim, J.C.; Kang, Y.S. Preparation and characterization of α-Fe2O3 nanorod-thin film by metal–organic chemical vapor deposition. Thin Solid Film. 2009, 517, 1853–1856.
  48. Pradhan, S.K.; Reucroft, P.J.; Yang, F.; Dozier, A. Growth of TiO2 nanorods by metalorganic chemical vapor deposition. J. Cryst. Growth 2003, 256, 83–88.
  49. Liu, B.; Hu, Z.; Che, Y.; Allenic, A.; Sun, K.; Pan, X. Growth of ZnO nanoparticles and nanorods with ultrafast pulsed laser deposition. Appl. Phys. A 2008, 93, 813–818.
  50. Yang, T.; Yu, H.; Xiao, B.; Li, Z.; Zhang, M. Enhanced 1-butylamine gas sensing characteristics of flower-like V2O5 hierarchical architectures. J. Alloys Compd. 2017, 699, 921–927.
  51. Sik, C.M.; Young, K.M.; Mirzaei, A.; Kim, H.-S.; Kim, S.-I.; Baek, S.-H.; Won, C.D.; Jin, C.; Hyoung, L.K. Selective, sensitive, and stable NO2 gas sensor based on porous ZnO nanosheets. Appl. Surf. Sci. 2021, 568, 150910.
  52. Kim, H.-J.; Lee, J.-H. Highly sensitive and selective gas sensors using p-type oxide semiconductors: Overview. Sens. Actuators B Chem. 2014, 192, 607–627.
  53. Xia, Y.; Wang, J.; Chen, R.; Zhou, D.; Xiang, L. A review on the fabrication of hierarchical ZnO nanostructures for photocatalysis application. Crystals 2016, 6, 148.
  54. Liu, J.; Zhang, L.; Fan, J.; Zhu, B.; Yu, J. Triethylamine gas sensor based on Pt-functionalized hierarchical ZnO microspheres. Sens. Actuators B-Chem. 2021, 331, 129425.
  55. Cheng, P.; Dang, F.; Wang, Y.; Gao, J.; Xu, L.; Wang, C.; Lv, L.; Li, X.; Zhang, B.; Liu, B. Gas sensor towards n-butanol at low temperature detection: Hierarchical flower-like Ni-doped Co3O4 based on solvent-dependent synthesis. Sens. Actuators B-Chem. 2021, 328, 129028.
  56. Guo, M.; Luo, N.; Chen, Y.; Fan, Y.; Wang, X.; Xu, J. Fast-response MEMS xylene gas sensor based on CuO/WO3 hierarchical structure. J. Hazard. Mater. 2022, 429, 127471.
  57. Yu, H.; Guo, C.; Zhang, X.; Xu, Y.; Cheng, X.; Gao, S.; Huo, L. Recent Development of Hierarchical Metal Oxides Based Gas Sensors: From Gas Sensing Performance to Applications. Adv. Sustain. Syst. 2022, 6, 2100370.
  58. Yamazoe, N. New approaches for improving semiconductor gas sensors. Sens. Actuators B Chem. 1991, 5, 7–19.
  59. Mirzaei, A.; Yousefi, H.R.; Falsafi, F.; Bonyani, M.; Lee, J.-H.; Kim, J.-H.; Kim, H.W.; Kim, S.S. An overview on how Pd on resistive-based nanomaterial gas sensors can enhance response toward hydrogen gas. Int. J. Hydrogen Energy 2019, 44, 20552–20571.
  60. Iordache, S.M.; Ionete, E.I.; Iordache, A.M.; Tanasa, E.; Stamatin, I.; Grigorescu, C.E.A. Pd-decorated CNT as sensitive material for applications in hydrogen isotopes sensing-Application as gas sensor. Int. J. Hydrogen Energy 2021, 46, 11015–11024.
  61. Li, G.; Fan, Y.; Hu, Q.; Zhang, D.; Ma, Z.; Cheng, Z.; Wang, X.; Xu, J. Morphology and size effect of Pd nanocrystals on formaldehyde and hydrogen sensing performance of SnO2 based gas sensor. J. Alloys Compd. 2022, 906, 163765.
  62. Shen, S.-K.; Cui, X.-L.; Guo, C.-Y.; Dong, X.; Zhang, X.-F.; Cheng, X.-L.; Huo, L.-H.; Xu, Y.-M. Sensing mechanism of Ag/α-MoO3 nanobelts for H2S gas sensor. Rare Met. 2021, 40, 1545–1553.
  63. Kim, J.-H.; Kim, S.S. Realization of ppb-scale toluene-sensing abilities with Pt-functionalized SnO2–ZnO core–shell nanowires. ACS Appl. Mater. Interfaces 2015, 7, 17199–17208.
  64. 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.
  65. Li, X.; Yu, J.; Wageh, S.; Al-Ghamdi, A.A.; Xie, J. Graphene in photocatalysis: A review. Small 2016, 12, 6640–6696.
  66. Schedin, F.; Geim, A.K.; Morozov, S.V.; Hill, E.W.; Blake, P.; Katsnelson, M.I.; Novoselov, K.S. Detection of individual gas molecules adsorbed on graphene. Nat. Mater. 2007, 6, 652–655.
  67. Chatterjee, S.G.; Chatterjee, S.; Ray, A.K.; Chakraborty, A.K. Graphene–metal oxide nanohybrids for toxic gas sensor: A review. Sens. Actuators B Chem. 2015, 221, 1170–1181.
  68. Pendolino, F.; Armata, N. Graphene Oxide in Environmental Remediation Process; Springer: Berlin/Heidelberg, Germany, 2017; Volume 11.
  69. Hummers, W.S., Jr.; Offeman, R.E. Preparation of graphitic oxide. J. Am. Chem. Soc. 1958, 80, 1339.
  70. Brodie, B.C. Hydration behavior and dynamics of water molecules in graphite oxide. Ann. Chim Phys. 1860, 59, 466–472.
  71. Shewale, P.S.; Yun, K.-S. Synthesis and characterization of Cu-doped ZnO/RGO nanocomposites for room-temperature H2S gas sensor. J. Alloys Compd. 2020, 837, 155527.
  72. Liu, Z.; Yu, L.; Guo, F.; Liu, S.; Qi, L.; Shan, M.; Fan, X. Facial development of high performance room temperature NO2 gas sensors based on ZnO nanowalls decorated rGO nanosheets. Appl. Surf. Sci. 2017, 423, 721–727.
  73. Jia, A.; Liu, B.; Liu, H.; Li, Q.; Yun, Y. Interface design of SnO2@PANI nanotube with enhanced sensing performance for ammonia detection at room temperature. Front. Chem. 2020, 8, 383.
  74. Amiri, V.; Roshan, H.; Mirzaei, A.; Neri, G.; Ayesh, A.I. Nanostructured metal oxide-based acetone gas sensors: A review. Sensors 2020, 20, 3096.
  75. Sharma, H.J.; Salorkar, M.A.; Kondawar, S.B. H2 and CO gas sensor from SnO2/polyaniline composite nanofibers fabricated by electrospinning. Adv. Mater. Proc. 2017, 2, 61–66.
  76. Lee, J.-H.; Mirzaei, A.; Kim, J.H.; Kim, J.-Y.; Nasriddinov, A.F.; Marina Rumyantseva, N.; Kim, H.W.; Kim, S.S. Gas-sensing behaviors of TiO2-layer-modified SnO2 quantum dots in self-heating mode and effects of the TiO2 layer. Sens. Actuators B-Chem. 2020, 310, 127870.
  77. Liu, J.; Lv, J.; Xiong, H.; Wang, Y.; Jin, G.; Zhai, Z.; Fu, C.; Zhang, Q. Size effect and comprehensive mathematical model for gas-sensing mechanism of SnO2 thin film gas sensors. J. Alloys Compd. 2022, 898, 162875.
  78. Zhu, L.; Wang, M.; Lam, T.K.; Zhang, C.; Du, H.; Li, B.; Yao, Y. Fast microwave-assisted synthesis of gas-sensing SnO2 quantum dots with high sensitivity. Sens. Actuators B-Chem. 2016, 236, 646–653.
  79. Yang, X.; Zhang, S.; Zhao, L.; Sun, P.; Wang, T.; Liu, F.; Yan, X.; Gao, Y.; Liang, X.; Zhang, S. One step synthesis of branched SnO2/ZnO heterostructures and their enhanced gas-sensing properties. Sens. Actuators B Chem. 2019, 281, 415–423.
  80. Woo, H.-S.; Kwak, C.-H.; Chung, J.-H.; Lee, J.-H. Highly selective and sensitive xylene sensors using Ni-doped branched ZnO nanowire networks. Sens. Actuators B Chem. 2015, 216, 358–366.
  81. Rafiee, Z.; Roshan, H.; Sheikhi, M.H. Low concentration ethanol sensor based on graphene/ZnO nanowires. Ceram. Int. 2021, 47, 5311–5317.
  82. Bang, J.H.; Choi, M.S.; Mirzaei, A.; Kwon, Y.J.; Kim, S.S.; Kim, T.W.; Kim, H.W. Selective NO2 sensor based on Bi2O3 branched SnO2 nanowires. Sens. Actuators B Chem. 2018, 274, 356–369.
  83. Liao, L.; Mai, H.X.; Yuan, Q.; Lu, H.B.; Li, J.C.; Liu, C.; Yan, C.H.; Shen, Z.X.; Yu, T. Single CeO2 nanowire gas sensor supported with Pt nanocrystals: Gas sensitivity, surface bond states, and chemical mechanism. J. Phys. Chem. C 2008, 112, 9061–9065.
  84. Tonezzer, M.; Hieu, N.V. Size-dependent response of single-nanowire gas sensors. Sens. Actuators B Chem. 2012, 163, 146–152.
  85. Tonezzer, M. Selective gas sensor based on one single SnO2 nanowire. Sens. Actuators B Chem. 2019, 288, 53–59.
  86. Xu, Y.; Tian, X.; Fan, Y.; Sun, Y. A formaldehyde gas sensor with improved gas response and sub-ppm level detection limit based on NiO/NiFe2O4 composite nanotetrahedrons. Sens. Actuators B Chem. 2020, 309, 127719.
  87. Hwang, I.-S.; Kim, S.-J.; Choi, J.-K.; Choi, J.; Ji, H.; Kim, G.-T.; Cao, G.; Lee, J.-H. Synthesis and gas sensing characteristics of highly crystalline ZnO–SnO2 core–shell nanowires. Sens. Actuators B Chem. 2010, 148, 595–600.
  88. Chen, I.; Lin, S.-S.; Lin, T.-J.; Hsu, C.-L.; Hsueh, T.J.; Shieh, T.-Y. The assessment for sensitivity of a NO2 gas sensor with ZnGa2O4/ZnO core-shell nanowires—A novel approach. Sensors 2010, 10, 3057–3072.
  89. Park, S.; Hong, T.; Jung, J.; Lee, C. Room temperature hydrogen sensing of multiple networked ZnO/WO3 core–shell nanowire sensors under UV illumination. Curr. Appl. Phys. 2014, 14, 1171–1175.
  90. Jang, Y.-G.; Kim, W.-S.; Kim, D.-H.; Hong, S.-H. Fabrication of Ga2O3/SnO2 core–shell nanowires and their ethanol gas sensing properties. J. Mater. Res. 2011, 26, 2322–2327.
  91. Kim, J.H.; Mirzaei, A.; Kim, H.W.; Kim, S.S. Extremely sensitive and selective sub-ppm CO detection by the synergistic effect of Au nanoparticles and core–shell nanowires. Sens. Actuators B-Chem. 2017, 249, 177–188.
  92. Zhu, L.F.; She, J.C.; Luo, J.Y.; Deng, S.Z.; Chen, J.; Ji, X.W.; Xu, N.S. Self-heated hydrogen gas sensors based on Pt-coated W18O49 nanowire networks with high sensitivity, good selectivity and low power consumption. Sens. Actuators B Chem. 2011, 153, 354–360.
  93. Ngoc, T.M.; Van Duy, N.; Hoa, N.D.; Hung, C.M.; Nguyen, H.; Van Hieu, N. Effective design and fabrication of low-power-consumption self-heated SnO2 nanowire sensors for reducing gases. Sens. Actuators B Chem. 2019, 295, 144–152.
  94. Lim, S.K.; Hwang, S.-H.; Chang, D.; Kim, S. Preparation of mesoporous In2O3 nanofibers by electrospinning and their application as a CO gas sensor. Sens. Actuators B Chem. 2010, 149, 28–33.
  95. Lin, Y.; Wei, W.; Li, Y.; Li, F.; Zhou, J.; Sun, D.; Chen, Y.; Ruan, S. Preparation of Pd nanoparticle-decorated hollow SnO2 nanofibers and their enhanced formaldehyde sensing properties. J. Alloys Compd. 2015, 651, 690–698.
  96. Gao, X.; Li, F.; Wang, R.; Zhang, T. A formaldehyde sensor: Significant role of pn heterojunction in gas-sensitive core-shell nanofibers. Sens. Actuators B Chem. 2018, 258, 1230–1241.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , , , ,
View Times: 509
Revisions: 2 times (View History)
Update Date: 04 May 2023
1000/1000
Video Production Service