Reduced Graphene Oxide-Loaded Metal-Oxide Nanofiber Gas Sensors: History
Please note this is an old version of this entry, which may differ significantly from the current revision.

Reduced graphene oxide (rGO) is a reduced form of graphene oxide used extensively in gas sensing applications. On the other hand, in its pristine form, graphene has shortages and is generally utilized in combination with other metal oxides to improve gas sensing capabilities. There are different ways of adding rGO to different metal oxides with various morphologies. 

  • gas sensor
  • reduced graphene oxide
  • rGO-loading
  • metal oxide
  • sensing mechanism

1. Introduction

Metal oxide gas sensors are used to sense various toxic gases and vapors [1][2] in many areas [3]. These sensors are quite popular owing to their low preparation costs, high sensitivity, fast dynamics, high stability, simple operation, and small size [4]. The general gas sensing mechanism of metal oxide gas sensors stems from the modulation of electrical resistance in the presence of target gases [5]. In these sensors, the sensing layer is exposed directly to the gas. The interaction between the target gas and sensing layer modulates the electrical resistance of the gas sensor, resulting in the generation of a sensing signal. This conductivity change is due to variations in the width of the electron depletion layer across the exposed area of the sensing layer [6]. Various strategies can be used to enhance the gas sensing characteristics of metal oxide-based gas sensors, such as UV light activation [7], fabrication of p-n, p-p and n-n heterojunctions [8], noble metal decoration [9],surface engineering, and the generation of structural defects [10]. Thus far, different morphologies of metal oxides, such as nanowires [11], nanotubes [12], nanorods [13], nanobelts [14], nanofibers (NFs), nanosheets [15], and hierarchical structures [16], have been used in gas sensing studies. This is because gas adsorption on the surface of gas sensing relates directly to the surface area of gas sensors and higher adsorption of gas means higher sensing signal. Thus, several studies have attempted to synthesize morphologies with a high surface area to increase the sensing performance of gas sensors. In addition, a porous morphology [17] and hollow structures [18][19] can be other useful techniques to increase the gas sensing properties. Therefore, metal oxide NFs are highly popular for sensing studies because of a very high surface area results from the one-dimensional morphology and the presence of nanograins on their surfaces. The generation of depletion layers on nanograins causes the development of potential barriers, which will be modulated during exposure to the target gases, contributing to the sensing signal generation.
Another advantage of NFs is their ease of preparation using a facile electrospinning technique. In general, the electrospinning technique can be used to fabricate continuous fibers with the possibility of control over the fiber diameter [20]. In addition, different NFs, such as porous NFs [21], core-shell NFs [22][23], and complex NFs [24], can be easily synthesized by electrospinning. Furthermore, NFs with controllable alignment can be produced by modifying the electrospinning [25]. Electrospinning also can be used for the mass production of NFs because of the easy handling, possibility of control of the diameter, low cost, simple operation, and high reproducibility [26]. Briefly, the features of electrospun fibers can be controlled by electrospinning parameters, including the solution variables (e.g., surface tension, viscosity, and conductivity), operating variables (e.g., applied voltage, spinning distance, and solution flow rate), and ambient variables (e.g., humidity and temperature) [27].Li et al. [28] ) and Xue et al. [29] reviewed these factors, so this entry does not discuss them further.
The major components of electrospinning include a syringe pump, spinneret, conductive collector, and power supply. The electrospinning technique can be described as follows: (1) charging of the liquid droplet and the formation of a Taylor cone, (2) formation of the charged jet, (3) thinning of the jet by the applied voltage, and (4) collection and solidification of the jet on a grounded collector. A Taylor cone is formed owing to surface tension and the applied electric field, from which a charged jet is ejected. The ejected jet was solidified quickly, resulting in the collection of solid fibers on the collector [30][31].

2. Graphene, Graphene Oxide, and Reduced Graphene Oxide

Graphene is a single-layer comprised of sp2carbon atoms that can be used as a gas sensor owing to its high charge carrier mobility (200,000 cm2 V−1 s−1), high mechanical stiffness, high environmental compatibility, and huge surface area (2630 m2g−1) [32][33]. Schedin et al. [34] introduced the first graphene gas sensor in 2007. Because atoms in a single layer of graphene can be considered surface atoms, graphene can interact with even a single molecule [34]. Even though nowadays pristine graphene can be synthesized on a large scale [35] with good water solubility [36][37], it can be easily agglomerated in solution due to surface interactions [38]. Moreover, graphene has no bandgap or functional groups, limiting its gas sensors applications, particularly in the pristine form [39]. Therefore, reduced graphene oxide (rGO), which is synthesized by the reduction of graphene oxide (GO), is a better choice for gas sensing applications because it has many functional groups and defects [40]. GO has also been used for gas sensing studies [41]. On the other hand, GO has very high resistance due to the presence of alkoxy (C-O-C), hydroxyl (-OH), carboxylic acid (-COOH), carbonyl (C=O), and other oxygen-based functional groups [42][43]. rGO has more defects and dangling bonds than graphite, resulting in better sensing properties [44]. GO is widely prepared using either Hummers [45] or Brodie [46] methods. In these methods, differences are in both the acid used (nitric or sulfuric acid), and the type of salt used (potassium chlorate or potassium permanganate). By subsequent reduction of GO, rGO can be obtained. In fact, rGO can be prepared easily from GO by chemical reduction, thermal reduction, and UV light reduction [47]
Furthermore, rGO has high thermal stability, and the total weight loss of rGO was reported to be only 11% up to 800 °C, which was attributed to the absence of most oxygen functional groups. [48]. Pristine rGO gas sensors have a long response and recovery times, and incorporating rGO with metal oxides can be a good strategy to increase the sensing capabilities of rGO-based gas sensors [49]. The synthesis and properties of rGO have been reviewed comprehensively [38].
Table 1 lists the gas sensing characteristics of rGO-loaded metal oxide NFs reported in literature. Different oxidizing (NO2) and reducing (C6H6, H2, CO, H2S, C3H6O, and C7H8) gases, have been successfully detected by these gas sensors. The sensing temperatures ranging from room temperature up to 400 °C have been reported. This demonstrates the possibility of a high sensing temperature of rGO-loaded gas sensors. In some cases, such as Ref. [50], a very high response to target gas can be obtained. Finally, in some cases, high responses to low concentrations of gases have been reported. Overall, the data in Table 1 shows the promising effects of rGO to be used along with metal oxide NFs for gas sensing studies.
Table 1. Gas sensing characteristics of rGO-loaded metal oxide NFs.
Also, Table 2 summarizes precursors, NF diameter, surface area and porosity of different rGO-loaded NF gas sensors reported in the literature. In all cases, initially, a viscous solution of precursor materials was prepared and then it was loaded into a syringe and subsequently NFs were produced by electrospinning process. It should be noted that in most cases, rGO was reduced from the synthesized GO (via Hummers method). In most cases, the surface area is not mentioned, but it could reach to 78.57 m2/g. In addition, in most cases the NFs with different diameters and mesoporous nature can be obtained by electrospinning.
Table 2. Precursors, NF diameter, surface area and porosity nature of rGO-loaded NF gas sensors reported in literature.

Sensing Material

Precursors

NF Diameter (nm)

Surface Area (m2/g)

Porosity Type

Ref.

rGO-loaded ZnO NFs

Zinc acetate, PVA

190

NA

NA

[50]

rGO-loaded ZnO NFs

Zinc acetate,

polyvinyl alcohol (PVA)

~150

NA

NA

[51]

Au and Pd/rGO-loaded ZnO NFs

Zinc acetate, PVA, HAuCl4⋅nH2O, PdCl2

~200

NA

Mesoporous

[52]

0.44 wt.% rGO-loaded SnO2 NFs

SnCl2.2H2O, polyvinyl acetate (PVAc)

~180

7.0574

Mesoporous

[53]

rGO-loaded-SnO2 NFs under UV light

SnCl2.2H2O, PVP, dimethyl formamide (DMF)

80–250

NA

NA

[54]

0.01 wt.% rGO-loaded-SnO2 NFs

PVP, polymethylmethacrylate (PMMA) + tin(IV) acetate, acetic acid

370

NA

NA

[55]

rGO/Pt and Pd co-loaded SnO2 NFs

SnCl2.2H2O, PVAc, PdCl2, H2PtCl6.nH2O, DMF

NA

NA

Mesoporous

[56]

1 wt.% rGO-loaded Fe2O3 NFs

Ferric acetylacetonte

PVP

100

NA

NA

[57]

1 wt.% rGO-loaded Fe2O3 NFs

PVA, Fe(NO3)3.9H2O

50–100

NA

NA

[58]

rGO-loaded In2O3 NFs

InCl3, PVP, Monomethylamine (MMA), Trimethylamine (TMA), Triethylamine (TEA), and N,N-Dimethylformamide (DMF)

50

NA

Mesoporous

[59]

2.2 wt.% rGO-loaded In2O3 NFs

PVP, DMF and (In(NO3)3·4.5H2O

76

39.82

Mesoporous

[60]

1wt.% rGO–Co3O4NFs

Co(NO3)2·6H2O, PVP, Ethanol

200–300

78.57

Mesoporous

[61]

0.5 wt.% rGO-loaded CuO NFs

PVA, (Cu (CH3CO2)2)

∼50

NA

NA

[62]

rGO-loaded ZnFe2O4 NFs

Zn(CH3COO)2·2H2O, Fe(NO3)3·9H2O and PVA

50–100

NA

Mesoporous

[63]

NA: not available.

This entry is adapted from the peer-reviewed paper 10.3390/s21041352

References

  1. Velmathi, G.; Mohan, S.; Henry, R. Analysis and review of tin oxide-based chemoresistive gas sensor. IETE Tech. Rev. 2016, 33, 323–331.
  2. Wetchakun, K.; Samerjai, T.; Tamaekong, N.; Liewhiran, C.; Siriwong, C.; Kruefu, V.; Wisitsoraat, A.; Tuantranont, A.; Phanichphant, S. Semiconducting metal oxides as sensors for environmentally hazardous gases. Sens. Actuators B Chem. 2011, 160, 580–591.
  3. Feng, S.; Farha, F.; Li, Q.; Wan, Y.; Xu, Y.; Zhang, T.; Ning, H. Review on smart gas sensing technology. Sensors 2019, 19, 3760.
  4. Majhi, S.M.; Mirzaei, A.; Kim, H.W.; Kim, S.S.; Kim, T.W. Recent advances in energy-saving chemiresistive gas sensors: A review. Nano Energy 2021, 79, 105369.
  5. Mirzaei, A.; Kim, J.-H.; Kim, H.W.; Kim, S.S. Resistive-based gas sensors for detection of benzene, toluene and xylene (btx) gases: A review. J. Mater. Chem. C 2018, 6, 4342–4370.
  6. Nazemi, H.; Joseph, A.; Park, J.; Emadi, A. Advanced micro- and nano-gas sensor technology: A review. Sensors 2019, 19, 1285.
  7. Wang, J.; Shen, H.; Xia, Y.; Komarneni, S. Light-activated room-temperature gas sensors based on metal oxide nanostructures: A review on recent advances. Ceram. Inter. 2020. Corrected Proof.
  8. Miller, D.R.; Akbar, S.A.; Morris, P.A. Nanoscale metal oxide-based heterojunctions for gas sensing: A review. Sens. Actuators B Chem. 2014, 204, 250–272.
  9. Singhal, A.V.; Charaya, H.; Lahiri, I. Noble metal decorated graphene-based gas sensors and their fabrication: A review. Crit. Rev. Solid State Mater. Sci. 2017, 42, 499–526.
  10. Kim, J.-H.; Kim, J.-Y.; Lee, J.-H.; Mirzaei, A.; Kim, H.W.; Hishita, S.; Kim, S.S. Indium-implantation-induced enhancement of gas sensing behaviors of SnO2 nanowires by the formation of homo-core–shell structure. Sens. Actuators B Chem. 2020, 321, 128475.
  11. Chen, X.; Wong, C.K.Y.; Yuan, C.A.; Zhang, G. Nanowire-based gas sensors. Sens. Actuators B Chem. 2013, 177, 178–195.
  12. Rezaie, S.; Bafghi, Z.G.; Manavizadeh, N. Carbon-doped zno nanotube-based highly effective hydrogen gas sensor: A first-principles study. Int. J. Hydrogen Energy 2020, 45, 14174–14182.
  13. Gao, R.; Cheng, X.; Gao, S.; Zhang, X.; Xu, Y.; Zhao, H.; Huo, L. Highly selective detection of saturated vapors of abused drugs by ZnO nanorod bundles gas sensor. Appl. Surf. Sci. 2019, 485, 266–273.
  14. Suman, P.; Felix, A.; Tuller, H.; Varela, J.; Orlandi, M. Comparative gas sensor response of SnO2, SnO and Sn3O4 nanobelts to SnO2 and potential interferents. Sens. Actuators B Chem. 2015, 208, 122–127.
  15. Choi, S.-J.; Kim, I.-D. Recent developments in 2D nanomaterials for chemiresistive-type gas sensors. Electron. Mater. Lett. 2018, 14, 221–260.
  16. Lee, J.-H. Gas sensors using hierarchical and hollow oxide nanostructures: Overview. Sens. Actuators B Chem. 2009, 140, 319–336.
  17. Korotcenkov, G.; Cho, B.K. Porous semiconductors: Advanced material for gas sensor applications. Crit. Rev. Solid State Mater. Sci. 2010, 35, 1–37.
  18. Lu, Z.; Zhou, Q.; Wei, Z.; Xu, L.; Peng, S.; Zeng, W. Synthesis of hollow nanofibers and application on detecting SF6 decomposing products: A mini review. Front. Mater. 2019, 6, 183.
  19. Hanh, N.H.; Van Duy, L.; Hung, C.M.; Van Duy, N.; Heo, Y.-W.; Van Hieu, N.; Hoa, N.D. Voc gas sensor based on hollow cubic assembled nanocrystal Zn2SnO4 for breath analysis. Sens. Actuators A Phys. 2020, 302, 111834.
  20. Aruna, S.T.; Balaji, L.; Kumar, S.S.; Prakash, B.S.J. Electrospinning in solid oxide fuel cells—A review. Renew. Sustain. Energy Rev. 2017, 67, 673–682.
  21. Phuoc, P.H.; Hung, C.M.; Van Toan, N.; Van Duy, N.; Hoa, N.D.; Van Hieu, N. One-step fabrication of SnO2 porous nanofiber gas sensors for sub-ppm H2S detection. Sens. Actuators A. Phys. 2020, 303, 111722.
  22. 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.
  23. Yoon, J.; Yang, H.S.; Lee, B.S.; Yu, W.R.J.A.M. Recent progress in coaxial electrospinning: New parameters, various structures, and wide applications. Adv. Mater. 2018, 30, 1704765.
  24. Yu, D.G.; Wang, M.; Li, X.; Liu, X.; Zhu, L.M.; Annie Bligh, S.W. Multifluid electrospinning for the generation of complex nanostructures. Wiley Interdisciplinary Reviews: Nanomed. Nanobiotechnol. 2020, 12, e1601.
  25. Kishan, A.P.; Cosgriff-Hernandez, E.M. Recent advancements in electrospinning design for tissue engineering applications: A review. J. Biomed. Mater. Res. Part A 2017, 105, 2892–2905.
  26. 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.
  27. Liu, Q.; Zhu, J.; Zhang, L.; Qiu, Y. Recent advances in energy materials by electrospinning. Renew. Sustain. Energy Rev. 2018, 81, 1825–1858.
  28. Li, X.; Chen, W.; Qian, Q.; Huang, H.; Chen, Y.; Wang, Z.; Chen, Q.; Yang, J.; Li, J.; Mai, Y.-W. Electrospinning-based strategies for battery materials. Adv. Energy Mater. 2020, 11, 2000845.
  29. Xue, J.; Wu, T.; Dai, Y.; Xia, Y. Electrospinning and electrospun nanofibers: Methods, materials, and applications. Chem. Rev. 2019, 119, 5298–5415.
  30. Xue, J.; Xie, J.; Liu, W.; Xia, Y. Electrospun nanofibers: New concepts, materials, and applications. Acc. Chem. Res. 2017, 50, 1976–1987.
  31. Wen, P.; Zong, M.-H.; Linhardt, R.J.; Feng, K.; Wu, H. Electrospinning: A novel nano-encapsulation approach for bioactive compounds. Trends Food Sci. Technol. 2017, 70, 56–68.
  32. Li, X.; Yu, J.; Wageh, S.; Al-Ghamdi, A.A.; Xie, J. Graphene in photocatalysis: A review. Small 2016, 12, 6640–6696.
  33. Rao, C.N.R.; Sood, A.K.; Subrahmanyam, K.S.; Govindaraj, A. Graphene: The new two-dimensional nanomaterial. Angew. Chem. Int. Ed. 2009, 48, 7752–7777.
  34. Schedin, F.; Geim, A.K.; Morozov, S.V.; Hill, E.; Blake, P.; Katsnelson, M.; Novoselov, K.S. Detection of individual gas molecules adsorbed on graphene. Nat. Mater. 2007, 6, 652–655.
  35. Wang, X.; You, H.; Liu, F.; Li, M.; Wan, L.; Li, S.; Li, Q.; Xu, Y.; Tian, R.; Yu, Z.; et al. Large-scale synthesis of few-layered graphene using CVD. Chem. Vap. Depos. 2009, 15, 53–56.
  36. Si, Y.; Samulski, E.T. Synthesis of water soluble graphene. Nano Lett. 2008, 8, 1679–1682.
  37. Zhang, M.; Bai, L.; Shang, W.; Xie, W.; Ma, H.; Fu, Y.; Fang, D.; Sun, H.; Fan, L.; Han, M.; et al. Facile synthesis of water-soluble, highly fluorescent graphene quantum dots as a robust biological label for stem cells. J. Mater. Chem. 2012, 22, 7461–7467.
  38. Smith, A.T.; LaChance, A.M.; Zeng, S.; Liu, B.; Sun, L. Synthesis, properties, and applications of graphene oxide/reduced graphene oxide and their nanocomposites. Nano Mater. Sci. 2019, 1, 31–47.
  39. Gupta Chatterjee, S.; 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.
  40. Kumar, R.; Kaur, A. Chemiresistive gas sensors based on thermally reduced graphene oxide for sensing sulphur dioxide at room temperature. Diam. Relat. Mater. 2020, 109, 108039.
  41. Toda, K.; Furue, R.; Hayami, S. Recent progress in applications of graphene oxide for gas sensing: A review. Anal. Chim. Acta 2015, 878, 43–53.
  42. Pendolino, F.; Armata, N. Graphene Oxide in Environmental Remediation Process; Springer: Cham, Switzerland, 2017; pp. 5–21.
  43. Guex, L.G.; Sacchi, B.; Peuvot, K.F.; Andersson, R.L.; Pourrahimi, A.M.; Ström, V.; Farris, S.; Olsson, R.T. Experimental review: Chemical reduction of graphene oxide (GO) to reduced graphene oxide (rGO) by aqueous chemistry. Nanoscale 2017, 9, 9562–9571.
  44. Xiao, Y.; Yang, Q.; Wang, Z.; Zhang, R.; Gao, Y.; Sun, P.; Sun, Y.; Lu, G. Improvement of SnO2 gas sensing performance based on discoid tin oxide modified by reduced graphene oxide. Sensors Actuators B Chem. 2016, 227, 419–426.
  45. Hummers, W.S.; Offeman, R.E. Preparation of graphitic oxide. J. Am. Chem. Soc. 1958, 80, 1339.
  46. Brodie, B.C. Hydration behavior and dynamics of water molecules in graphite oxide. Ann. Chim. Phys. 1860, 59, 466–472.
  47. Mohan, V.B.; Lau, K.-T.; Hui, D.; Bhattacharyya, D. Graphene-based materials and their composites: A review on production, applications and product limitations. Comp. Part B. Eng. 2018, 142, 200–220.
  48. Cui, P.; Lee, J.; Hwang, E.; Lee, H. One-pot reduction of graphene oxide at subzero temperatures. Chem. Commun. 2011, 47, 12370–12372.
  49. Drmosh, Q.A.; Yamani, Z.H.; Hendi, A.H.; Gondal, M.A.; Moqbel, R.A.; Saleh, T.A.; Khan, M.Y. A novel approach to fabricating a ternary rGO/ZnO/Pt system for high-performance hydrogen sensor at low operating temperatures. Appl. Surf. Sci. 2019, 464, 616–626.
  50. Abideen, Z.U.; Kim, H.W.; Kim, S.S. An ultra-sensitive hydrogen gas sensor using reduced graphene oxide-loaded ZnO nanofibers. Chem. Commun. 2015, 51, 15418–15421.
  51. Abideen, Z.U.; Katoch, A.; Kim, J.-H.; Kwon, Y.J.; Kim, H.W.; Kim, S.S. Excellent gas detection of ZnO nanofibers by loading with reduced graphene oxide nanosheets. Sens. Actuators B Chem. 2015, 221, 1499–1507.
  52. Abideen, Z.U.; Kim, J.-H.; Mirzaei, A.; Kim, H.W.; Kim, S.S. Sensing behavior to ppm-level gases and synergistic sensing mechanism in metal-functionalized rgo-loaded ZnO nanofibers. Sens. Actuators B Chem. 2018, 255, 1884–1896.
  53. Lee, J.-H.; Katoch, A.; Choi, S.-W.; Kim, J.-H.; Kim, H.W.; Kim, S.S. Extraordinary improvement of gas-sensing performances in SnO2 nanofibers due to creation of local p–n heterojunctions by loading reduced graphene oxide nanosheets. ACS Appl. Mater. Interfaces 2015, 7, 3101–3109.
  54. Li, W.; Guo, J.; Cai, L.; Qi, W.; Sun, Y.; Xu, J.-L.; Sun, M.; Zhu, H.; Xiang, L.; Xie, D.; et al. UV light irradiation enhanced gas sensor selectivity of NO2 and SO2 using rgo functionalized with hollow SnO2 nanofibers. Sens. Actuators B Chem. 2019, 290, 443–452.
  55. Choi, S.-J.; Jang, B.-H.; Lee, S.-J.; Min, B.K.; Rothschild, A.; Kim, I.-D. Selective detection of acetone and hydrogen sulfide for the diagnosis of diabetes and halitosis using SnO2 nanofibers functionalized with reduced graphene oxide nanosheets. ACS Appl. Mater. Interfaces 2014, 6, 2588–2597.
  56. Kim, J.-H.; Zheng, Y.; Mirzaei, A.; Kim, H.W.; Kim, S.S. Synthesis and selective sensing properties of rgo/metal-coloaded SnO2 nanofibers. J. Electron. Mater. 2017, 46, 3531–3541.
  57. Guo, L.; Kou, X.; Ding, M.; Wang, C.; Dong, L.; Zhang, H.; Feng, C.; Sun, Y.; Gao, Y.; Sun, P.; et al. Reduced graphene oxide/α-Fe2O3 composite nanofibers for application in gas sensors. Sens. Actuators B Chem. 2017, 244, 233–242.
  58. Van Hoang, N.; Hung, C.M.; Hoa, N.D.; Van Duy, N.; Van Toan, N.; Hong, H.S.; Hong Van, P.T.; Sơn, N.T.; Yoon, S.-G.; Van Hieu, N. Enhanced H2S gas-sensing performance of α-Fe2O3 nanofibers by optimizing process conditions and loading with reduced graphene oxide. J. Alloys Comp. 2020, 826, 154169.
  59. Andre, R.S.; Mercante, L.A.; Facure, M.H.M.; Mattoso, L.H.C.; Correa, D.S. Enhanced and selective ammonia detection using In2O3/reduced graphene oxide hybrid nanofibers. App. Surf. Sci. 2019, 473, 133–140.
  60. Yan, C.; Lu, H.; Gao, J.; Zhang, Y.; Guo, Q.; Ding, H.; Wang, Y.; Wei, F.; Zhu, G.; Yang, Z. Improved NO2 sensing properties at low temperature using reduced graphene oxide nanosheet–In2O3 heterojunction nanofibers. J. Alloys Comp. 2018, 741, 908–917.
  61. Feng, Q.; Li, X.; Wang, J.; Gaskov, A.M. Reduced graphene oxide (rGO) encapsulated Co3O4 composite nanofibers for highly selective ammonia sensors. Sens. Actuators B Chem. 2016, 222, 864–870.
  62. Kim, J.-H.; Mirzaei, A.; Zheng, Y.; Lee, J.-H.; Kim, J.-Y.; Kim, H.W.; Kim, S.S. Enhancement of H2S sensing performance of p-Cuo nanofibers by loading p-reduced graphene oxide nanosheets. Sens. Actuators B Chem. 2019, 281, 453–461.
  63. Van Hoang, N.; Hung, C.M.; Hoa, N.D.; Van Duy, N.; Park, I.; Van Hieu, N. Excellent detection of H2S gas at ppb concentrations using ZnFe2O4 nanofibers loaded with reduced graphene oxide. Sens. Actuators B Chem. 2019, 282, 876–884.
More
This entry is offline, you can click here to edit this entry!
Video Production Service