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 -- 2579 2023-07-07 09:16:29 |
2 format Meta information modification 2579 2023-07-10 03:54:11 |

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.
Li, S.; Zhou, S.; Zhao, S.; Jin, T.; Zhong, M.; Cen, Z.; Gao, P.; Yan, W.; Ling, M. Resistive H2 Sensor Operating at Room Temperature. Encyclopedia. Available online: https://encyclopedia.pub/entry/46556 (accessed on 25 July 2024).
Li S, Zhou S, Zhao S, Jin T, Zhong M, Cen Z, et al. Resistive H2 Sensor Operating at Room Temperature. Encyclopedia. Available at: https://encyclopedia.pub/entry/46556. Accessed July 25, 2024.
Li, Sixun, Shiyu Zhou, Shuaiyin Zhao, Tengfei Jin, Maohua Zhong, Zhuhao Cen, Peirong Gao, Wenjun Yan, Min Ling. "Resistive H2 Sensor Operating at Room Temperature" Encyclopedia, https://encyclopedia.pub/entry/46556 (accessed July 25, 2024).
Li, S., Zhou, S., Zhao, S., Jin, T., Zhong, M., Cen, Z., Gao, P., Yan, W., & Ling, M. (2023, July 07). Resistive H2 Sensor Operating at Room Temperature. In Encyclopedia. https://encyclopedia.pub/entry/46556
Li, Sixun, et al. "Resistive H2 Sensor Operating at Room Temperature." Encyclopedia. Web. 07 July, 2023.
Resistive H2 Sensor Operating at Room Temperature
Edit

Lithium-ion batteries (LIBs) have become one of the most competitive energy storage technologies. However, the “thermal runaway” of LIBs leads to serious safety issues. Early safety warning of LIBs is a prerequisite for the widely applications of power battery and large-scale energy storage systems. As reported, hydrogen (H2) could be generated due to the reaction of lithium metal and polymers inside the battery. The generation of H2 is some time earlier than the “thermal runaway”. Therefore, the rapid detection of trace hydrogen is the most effective method for early safety warning of LIBs. Resistive hydrogen sensors have attracted attention in recent years. In addition, they could be placed inside the LIB package for the initial hydrogen detection.

resistive hydrogen sensor room temperature

1. Introduction

For Lithium-ion batteries (LIBs), a high internal temperature probably induces additional violent exothermic chemical reactions for further heat generation, resulting in thermal runaway of LIBs. Hence, room temperature H2 sensors applied inside of LIB is necessary. Resistive H2 sensors operating at room temperature mainly rely on the catalytic adsorption of H2 molecule on noble metal (for example Pd, Au, Pt). Among of them, Pd-based H2 sensors are the current state-of-the-art resistive H2 sensors. Beyond that, synergistic effect of heterojunctions, metallization effect of special metal oxide [1][2], and efficient H2 sensing capacity of emerging materials (i.e., MXene, TMDs) at low temperatures have also been investigated to develop RT H2 sensor.

2. Pd-Based H2 Sensors

Basic H2 Sensing Mechanism
Generally, the H2 sensing mechanism of Pd-based H2 sensors are catalytic adsorption of H2 on Pd, which induces the dissociation of H2 molecule into H atoms and the formation of PdHx (Equations (1) and (2)) [3][4].
H2(gas)H2(ads)
Pd+x2H2(ads)PdHx
Numerous resistive Pd-based H2 sensors have been developed, including pure Pd nanostructures and Pd-based composites, such as Pd-hetero metal, Pd-MOX, Pd-carbon material, and Pd-TMD.
Pure Pd materials. According to Equations (1) and (2), PdHx is formed when H2 molecules adsorb on Pd surface. Conductivity of PdHx is poorer than Pd, which generates H2 response in resistance of Pd [3][5]. Plenty of resistive H2 sensors based on pure Pd materials have been reported. With the significant development of nanoscience, various Pd nanostructures with further H2 sensing mechanisms were developed to enhance the H2 sensing capacities [6][7][8][9].
Kim et al. [6] and Jung et al. [7] investigated the nanograin effects and the α-to-β phase transition of Pd on H2 sensing. When the Pd crystallites < 10 nm, abundant nanogaps were created between the nanograins (<2 nm). The nanogaps between Pd nanograins intrinsically stem the electrical conduction. Exposed to different concentration of H2, different kinds of conducting pathways were formed in Pd, yielding switchable H2 sensing behaviors [7]. Similar phenomena were demonstrated in Ref. [3]: (i) when exposed to low concentrations of H2 (2.5 to 100 ppm), α-PdHx was formed expanding Pd nanograins, which closed the narrow nanogaps and further producing a new conduction pathway. Obvious negative resistance response was observed. (ii) Upon exposed to moderate concentrations of H2 (250 to 2500 ppm), most of the nanogaps were closed, resulting in saturation of the negative resistance response. (iii) Upon exposed to high concentrations of H2 (0.5 to 3%), all nanogaps were closed and β-PdHx were formed, while surface electron scattering generated by β-PdHx dominantly the electrical conductions, leading to significant positive resistance response. Consequently, ultrasmall grain size of Pd are critical for H2 detection at low concentrations, α-to-β phase transition of Pd could detect a wide range of H2 concentrations from sub-ppm to 4%. In addition, negative effect of oxygen on Pd-based H2 sensing has been demonstrated in earlier literatures [10][11].
Pd-metal alloy. Recently, many researchers developed Pd-based alloys to improve the H2 sensing properties, such as limit detection and response/recovery behavior [12][13][14][15]. Jung et al. [13] developed Pd/Pt and Pd/Au high surface-to-volume ratio nanopatterns. Compared to Pd nanopattern, Pd/Pt nanopattern showed 45.5-fold higher response to 1% H2, Pd/Au nanopattern showed about 73-fold and 4.6-fold enhancement in the response and recovery speed, respectively. The dramatic improvements in response and recovery behaviors to H2 are attributed to the ultrasmall size (<5 nm), ultrathin nanopattern (<15 nm), grain interfaces and the lower adsorption/dissociation energy of H2 on Pd/Pt and Pd/Au surfaces. Kim et al. [15] developed hollow Pd-Sn alloy nanotubes with high surface area of 223.0 m2/g for H2 sensing and ultrafine grain size. Interestingly, the Pd-Sn alloy effectively prevented degradation of H2 sensing performances caused by the α-β transition of Pd. As a result, the hollow Pd-Sn nanotubes exhibited outstanding sensing properties to a wide concentration range of H2 (50 ppm to 3%), especially fast response/recovery rate to high concentration of H2 (20 s/17 s to 2% H2). The noteworthy H2 sensing mechanism was proposed: (i) The abundant interface gaps between the grains owing to larger atom size of Sn (0.141 nm) and smaller atom size of Pd (0.138 nm) dominated low concentration of H2 (0.5 ppm to 0.02%) sensing behavior, due to the nanogap effect. (ii) For high concentration of H2 (0.05% to 3%), the formation of β-PdHx governed the H2 sensing properties. The H2 sensing mechanism agreed with the previous reports [6][7].
Pd-MOXs. MOXs are the classical gas sensing semiconductor materials. However, their H2 sensing properties are limited at room temperature. Previously, many researchers developed Pd-MOX composites to achieve good H2 sensing at room temperature.
Zhang et al. [9] investigated the interconvertible effect on H2 sensing of catalyst nanoparticles and semiconductor support in Pd-decorated PdO hollow shells. They prepared PdO hollow shells, which were subsequently treated by NaBH4 to be partially reduced into Pd on the PdO surface. The Pd nanoparticles were discretely and physically inlaid on the surface of PdO with a ultrasmall size of ~ 2 nm. The catalytic effect of Pd, as well as the Schottky-junction between Pd and PdO, enhanced the H2 sensing performances even at 1 ppm. Notably, inlaid Pd in PdO shells prevented the agglomeration of Pd nanoparticles, which generated long-term stability.
In addition, Zhang et al. [16] designed an unique conduction model by inserting a high-conductive metallic core Au into less-conductive p-type PdO to boost the RT H2 sensing performances. As a result, Pd decorating Au@PdO demonstrated an ~90 times larger in H2 response than Pd decorating Pd@PdO. And the boosted H2 response helped the Pd-Au@PdO sensor showed ultralow LOD of 0.1 ppm. Pd-ZnO nanoflowers has been demonstrated RT H2 sensing in ppb level with experimental LOD of 300 ppb, due to the change in channel conductance of ZnO nanoflowers based on the incorporation of Pd [17].
Pd-carbon materials. Carbon nanotubes, graphene, and their derivatives are very promising in the room temperature gas sensor fields, due to their good electrical conductivity at RT. However, H2 molecule is very difficult to adsorb on the surface of carbon materials, due to weak H2 adsorption capacity. In regard of this case, lots of Pd-carbon material composites have been previously investigated for RT H2 sensors. Even so, low LOD and quick recovery behavior remain big challenges. As is well known, Pd particles at ultrasmall size are very helpful for the RT H2 sensing. However, they are very easy to aggregate to further cause degradation of overall H2 response.
During the synthesis process of one-pot solution method, DNA suppressed the stacking of graphene layers due to π-π interaction between DNA and graphene, meanwhile in-situ anchored PdO2 subnanoscale clusters on the exfoliated single layer graphene. The design showed mimic wrinkled morphology and sensing mechanism of natural olfactory neuroepithelium, being named BONe. The BONe boosted H2 sensing performance (25 s/35 s at 5 ppm H2 with experimental LOD of 50 ppb) at RT with yearlong durability. High surface area of the BONe, good conductivity of graphene, as well as the subnanoscale of PdO2, generated the ultra-sensitivity to ppb-level H2 at RT; subnanoscale of PdO2 anchored on graphene yielded yearlong stability. Notably, it was calculated that PdO2 exhibited a d-band downshift of −2.58 eV, which was a much further downshift than that of PdO (−2.16 eV) and Pd (−0.179 eV), governing the complete and fast recovery behavior during H2 detection. A similar d-band theory was proposed previously, that tuning of d-band energy level balanced the adsorption and desorption capacities, and the lower the d-band level, the weaker the adsorption [18].
Pd-other materials. With the emerging of new materials (metal-organic frame MOF, Mxene, et al.), composites of Pd and kind of new materials were developed for H2 sensing. The sensor based on flexible Ti3C2Tx MXene@Pd nanoclusters, reported in Ref. [19] delivered a response/recovery time of 32/161 s and sensitivity of 23% to 4% H2 at RT. The strong H2 adsorption into lattice of Pd nanoclusters induced electrons doping in MXene, generating fast response behavior. However, the limit of detection (0.5%) needs to be improved. The Pd-decorated sodium titanate nanoribbons (Pd-NTO NRs) developed by Zhang et al. [20] exhibited ultrafast response to 1% H2 within 1.1 s at RT, and a wide detection range (0.8 ppm to 10% H2). The excellent H2 sensing capacity benefited from the laterally paralleled morphology and abundant oxygen vacancies on edge sites of nanoribbons, as well as monodispersed Pd nanoparticles in size of ~3.5 nm. Oxygen in air could block the reactive sites, leading to depress the H2 sensitivity and retard the response/recovery rate [10][11]. For that, Kim and Penner et al. [21] designed patterned Pd nanowires covering ZIF-8 membrane (Pd NWs@ZIF-8) for H2 sensing. Although the ZIF-8 membrane reduced the H2 response slightly, 20-fold faster response/recovery rate (3.5% at 10/7 s to 1% H2 versus 5.9% at 164/229 s of Pd nanowires) was achieved due to the molecular sieving and acceleration effects of ZIF-8 since the pore size of ZIF-8 is 0.34 nm, which is larger than the diffusion kinetic diameter of H2 molecule (0.289 nm), but a little smaller than that of O2 molecule (0.345 nm).

3. Other Noble Metal-Based H2 Sensors

Similar to H2 adsorption property of Pd, other noble metals also exhibit H2 adsorption behavior [22]. Moreover, Au and Pt-based H2 sensors has been demonstrated due to the formation of MHx [23][24][25]. Guha et al. [23] reported Pt-functionalized rGO for excellent H2 sensing at RT, 65 s/230 s against 5000 ppm H2 with LOD of 200 ppm in air ambience. Interestingly, for this Pt-rGO composite, H2 response was larger in N2 environment than that in air ambience; however, sensor recovered faster in air than in N2. In N2 and air environments, H2 first physisorbed on Pt-rGO surface and then dissociated to from Pt-H (Equation (3)). But in air, H2 also reacted with the adsorbed oxygen on the Pt-rGO surface to form Pt-H and H2O (Equations (4) and (5)).
Pt+12H2PtH
Pt+12O2PtO
PtO+32H2PtH+H2O
The H2O competed with H2 molecules to adsorb on the Pt-rGO surface, reducing the H2 response. On the contrary, the adsorbed H2 and the dissociated H-atoms reacted with adsorbed oxygen to convert to water vapor while purging air, exhibiting fast recovery in air ambience. Similar effect of oxygen on Pd- and Pd@Pt-based H2 sensing has been demonstrated in earlier literatures [10][11].

4. MOXs-Based H2 Sensors

Generally, most MOXs-based resistive gas sensors operated at high temperature (>150 °C). However, recent approaches were developed to realize RT H2 sensing performances of MOXs, such as nanostructure design [26][27][28], composites of MOXs [29][30], and surface metallization [17]. Huang et al. [27] demonstrated the well-aligned MoO3 nanoribbon arrays exhibited great H2 response/recovery behavior at RT, with a response/recovery time of 3 s/16 s at 100 ppm H2 which is much shorter than 59 s/151 s of randomly arranged MoO3 nanoribbons. The accelerated H2 response/recovery rate was due to the fact that the high alignment of nanoribbons could increase the surface activity of MoO3 and suppress the nanojunction effect. On the contrary, in the randomly arranged MoO3, the interface diffusion of adsorbed oxygen caused serious nanojunction effect, resulting in much slower H2 response/recovery rate.
Yun et al. [17] designed holey engineered 2D ZnO-nanosheets for supersensitive H2 sensing. The sensor exhibited 115% response to 100 ppm H2 with short response/recovery time of 9 s/6 s at room temperature. And the experimental LOD was 5 ppm. Upon exposure to H2, surface of the ZnO nanosheets became metallic Zn. Metallization of Zn on the ZnO surface governed the gas sensing mechanism about high response and great selectivity to H2 at room temperature. Moreover, the synergetic effect of 2D nanosheets and interconnected holey/porous network of ZnO generated the excellent H2 sensing performances by offering abundant active sites for H2 molecules.

5. MoS2-Based H2 Sensors

In recent year, MoS2 has shown great H2 sensing potential due to its 2D van der Waals structures, H2 adsorption capacity and enable RT operation [31][32]. Unfortunately, sluggish response/recovery limited the applications of MoS2-based H2 sensors. In regard of this case, some composites based on MoS2 were investigated for H2 sensing. Huang and Chen at al [33] designed hybrid interlinked MoS2-ZnO nanotubes for RT H2 sensing of 51.1% to 500 ppm with 14/19 s response/recovery and LOD of 10 ppm, due to the increased oxygen vacancies and surface-active sites. Hollow MoS2/Pt-based chemiresistors, designed by Kim et al. [31], exhibited great H2 sensing performance with fast response/recovery rate at RT (8.1/16 s for 1%, and 2.7/16 s for 4% H2), due to the catalytic H2 spillover of Pt, as well as sufficiently permeable pathways and maximized active sites for H2 produced by the hollow MoS2. Similarly, vertically aligned edge-oriented Pd/MoS2 nanofilm was reported for great H2 sensing properties at RT, response of 33.7% to 500 ppm with response/recovery of 16 s/38 s, and LOD of 50 ppm [32]. Even so, selectivity of RT H2 sensors based on MoS2 has not been investigated widely and deeply.

6. Other-Based H2 Sensors

With the development of nanoscience and nanotechnology, new composites for RT H2 sensors have been developed. Xu and Ou et al. [34] explored ultrathin nickel oxysulfide which exhibited a selective and fully reversible response to H2 at RT for a wide range from 0.25% to 1%, due to the physisorption of H2 on the surface. Dash et al. [35] developed RT H2 sensors based on rGO-ZnFe2O4-Pd nanocomposite, showing high sensitivity and fast response/recovery rate (11.43% to 200 ppm with 18/29 s response/recovery) to a wide range of H2 (50–1000 ppm), due to the synergistic effect of rGO, ZnFe2O4 and Pd nanoparticles.
Up to now, there are lots of efforts to develop room temperature resistive H2 sensors. Unfortunately, no H2 sensors integrated inside of LIB cells have been reported. Table 1 summarizes the recent and representative room temperature resistive H2 sensors possibly integrated inside of LIB cells. Although new gas sensing materials are emerging all of the time, Pd and Pd-based materials have been the current state-of-the-art resistive H2 sensing material operating at room temperature. In conclusion, Pd is still the most excellent catalyst for RT H2 sensing with great selectivity due to the formation of PdHx. But the LOD (ppm level) and response/recovery behavior (tens of seconds) of Pd-based H2 sensors should be further improved. Sensors based on Pd-carbon materials should be state-of-the-art RT H2 sensors displaying LOD of ppb-level (due to the catalytic H2 adsorption/dissociation on Pd surface, as well as high surface area and RT good conductivity of carbon materials). However, they still suffer from slow response/recovery rate (tens of seconds) to ppb-H2 at RT. According to previous investigations, atomically dispersed sub-nano clusters (perhaps single atom) of Pd are in favor of ppb-level H2 detection at RT; optimal d-band energy level of Pd could yield complete and fast recovery; construction of conduction channel in Pd-based sensing material could generate fast response and recovery rate. Although some composites of MOXs exhibited good RT H2 sensing property, they inherently showed cross-sensitivity to other gases (such as NO2, CO and ethanol) and susceptibility to humidity. Despite some new structures and materials (ternary composites, MoS2, and MXene) emerging for RT H2 sensing, fast response in 1 s to ppb-H2 with special selectivity is still a critical issue.

References

  1. Kumar, M.; Bhatt, V.; Kim, J.; Abhyankar, A.C.; Chung, H.-J.; Singh, K.; Bin Cho, Y.; Yun, Y.J.; Lim, K.S.; Yun, J.-H. Holey engineered 2D ZnO-nanosheets architecture for supersensitive ppm level H2 gas detection at room temperature. Sens. Actuators B Chem. 2021, 326, 128839.
  2. Wang, C.; Zhou, G.; Li, J.; Yan, B.; Duan, W. Hydrogen-induced metallization of zinc oxide (2 1 1 0) surface and nanowires: The effect of curvature. Phys. Rev. B. 2008, 77, 245303.
  3. Favier, F.; Walter, E.C.; Zach, M.P.; Benter, T.; Penner, R.M. Hydrogen sensors and switches from electrodeposited palladium mesowire arrays. Science 2001, 293, 2227–2231.
  4. Sakintuna, B.; Lamari-Darkrim, F.; Hirscher, M. Metal hydride materials for solid hydrogen storage: A review. Int. J. Hydrogen Energy 2007, 32, 1121–1140.
  5. Xu, T.; Zach, M.; Xiao, Z.; Rosenmann, D.; Welp, U.; Kwok, W.; Crabtree, G. Self-assembled monolayer-enhanced hydrogen sensing with ultrathin palladium films. Appl. Phys. Lett. 2005, 86, 203104.
  6. Kim, D.H.; Kim, S.J.; Shin, H.; Koo, W.T.; Jang, J.S.; Kang, J.Y.; Jeong, Y.J.; Kim, I.D. High-Resolution, Fast, and Shape-Conformable Hydrogen Sensor Platform: Polymer Nanofiber Yarn Coupled with Nanograined ACS Nano 2019, 13, 6071–6082.
  7. Cho, S.Y.; Ahn, H.; Park, K.; Choi, J.; Kang, H.; Jung, H.T. Ultrasmall Grained Pd Nanopattern H2 Sensor. ACS Sens. 2018, 3, 1876–1883.
  8. Lim, M.A.; Kim, D.H.; Park, C.O.; Lee, Y.W.; Han, S.W.; Li, Z.Y.; Williams, R.S.; Park, I. A New Route toward Ultrasensitive, Flexible Chemical Sensors: Metal Nanotubes by Wet-Chemical Synthesis along Sacrificial Nanowire Templates. ACS Nano 2012, 6, 598–608.
  9. Gao, Z.; Wang, T.; Li, X.; Li, Q.; Zhang, X.; Cao, T.; Li, Y.; Zhang, L.; Guo, L.; Fu, Y. Pd-Decorated PdO Hollow Shells: A H2-Sensing System in Which Catalyst Nanoparticle and Semiconductor Support are Interconvertible. ACS Appl. Mater. Interfaces 2020, 12, 42971–42981.
  10. Li, X.W.; Liu, Y.; Hemminger, J.C.; Penner, R.M. Catalytically Activated Nanowires for Accelerated Hydrogen Gas Detection. ACS Nano 2015, 9, 3215–3225.
  11. Yang, F.; Kung, S.-C.; Cheng, M.; Hemminger, J.C.; Penner, R.M. Smaller is faster and more sensitive: The effect of wire size on the detection of hydrogen by single palladium nanowires. ACS Nano 2010, 4, 5233–5244.
  12. Kumar, A.; Zhao, Y.; Abraham, S.R.; Thundat, T.; Swihart, M.T. Pd Alloy Nanosheet Inks for Inkjet-Printable H2 Sensors on Paper. Adv. Mater. Interfaces 2022, 9, 2200363.
  13. Jung, W.B.; Cho, S.Y.; Suh, B.L.; Yoo, H.W.; Jeon, H.J.; Kim, J.; Jung, H.T. Polyelemental nanolithography via plasma ion bombardment: From fabrication to superior H2 sensing application. Adv. Mater. 2019, 31, 1805343.
  14. Hassan, K.; Chung, G.S. Fast and reversible hydrogen sensing properties of Pd-capped Mg ultra-thin films modified by hydrophobic alumina substrates. Sens. Actuators B Chem. 2017, 242, 450–460.
  15. Song, L.; Ahn, J.; Kim, D.-H.; Shin, H.; Kim, I.-D. Porous Pd-Sn Alloy Nanotube-Based Chemiresistor for Highly Stable and Sensitive H2 Detection. ACS Appl. Mater. Interfaces 2022, 14, 28378–28388.
  16. Yang, S.; Li, Q.; Li, C.; Cao, T.; Wang, T.; Fan, F.; Zhang, X.; Fu, Y. Enhancing the Hydrogen-Sensing Performance of p-Type PdO by Modulating the Conduction Model. ACS Appl. Mater. Interfaces 2021, 13, 52754–52764.
  17. Jeon, J.-Y.; Park, S.-J.; Ha, T.-J. Functionalization of Zinc Oxide Nanoflowers with Palladium Nanoparticles via Microwave Absorption for Room Temperature-Operating Hydrogen Gas Sensors in the ppb Level. ACS Appl. Mater. Interfaces 2021, 13, 25082–25091.
  18. Sun, S.; Zhou, X.; Cong, B.; Hong, W.; Chen, G. Tailoring the d-Band Centers Endows (NixFe1–x)2P Nanosheets with Efficient Oxygen Evolution Catalysis. ACS Catal. 2020, 10, 9086–9097.
  19. Zhu, Z.; Liu, C.; Jiang, F.; Liu, J.; Ma, X.; Liu, P.; Xu, J.; Wang, L.; Huang, R. Flexible and lightweight Ti3C2Tx colloidal nanoclusters paper film as novel H2 sensor. J. Hazard. Mater. 2020, 399, 123054.
  20. Wu, J.; Guo, Y.; Wang, Y.; Zhu, H.; Zhang, X. Ti3C2 MXene-derived sodium titanate nanoribbons for conductometric hydrogen gas sensors. Sens. Actuators B Chem. 2022, 361, 131693.
  21. Koo, W.T.; Qiao, S.; Ogata, A.F.; Jha, G.; Jang, J.S.; Chen, V.T.; Kim, I.D.; Penner, R.M. Accelerating Palladium Nanowire H(2) Sensors Using Engineered Nanofiltration. ACS Nano 2017, 11, 9276–9285.
  22. Greeley, J.; Mavrikakis, M. Surface and subsurface hydrogen: Adsorption properties on transition metals and near-surface alloys. J. Phys. Chem. B. 2005, 109, 3460–3471.
  23. Ghosh, R.; Santra, S.; Ray, S.K.; Guha, P.K. Pt-functionalized reduced graphene oxide for excellent hydrogen sensing at room temperature. Appl. Phys. Lett. 2015, 107, 153102.
  24. Kim, Y.; Choi, Y.S.; Park, S.Y.; Kim, T.; Hong, S.P.; Lee, T.H.; Moon, C.W.; Lee, J.H.; Lee, D.; Hong, B.H.; et al. Au decoration of a graphene microchannel for self-activated chemoresistive flexible gas sensors with substantially enhanced response to hydrogen. Nanoscale 2019, 11, 2966–2973.
  25. Phan, D.-T.; Youn, J.-S.; Jeon, K.-J. High-sensitivity and fast-response hydrogen sensor for safety application using Pt nanoparticle-decorated 3D graphene. Renew. Energy 2019, 144, 167–171.
  26. Kadhim, I.H.; Abu Hassan, H.; Ibrahim, F.T. Hydrogen gas sensing based on nanocrystalline SnO2 thin films operating at low temperatures. Int. J. Hydrogen Energy 2020, 45, 25599–25607.
  27. Yang, P.; Li, X.; Huang, H.; Yang, S.; Zhang, X.; Hu, Y.; Wang, Z.; Gu, H. Hydrogen sensing kinetics of laterally aligned MoO3 nanoribbon arrays with accelerated response and recovery performances at room temperature. Int. J. Hydrogen Energy 2020, 45, 23841–23850.
  28. Haidry, A.A.; Xie, L.; Wang, Z.; Li, Z. Hydrogen sensing and adsorption kinetics on ordered mesoporous anatase TiO2 surface. Appl. Surf. Sci. 2020, 500, 144219.
  29. Li, H.; Wu, C.-H.; Liu, Y.-C.; Yuan, S.-H.; Chiang, Z.-X.; Zhang, S.; Wu, R.-J. Mesoporous WO3-TiO2 heterojunction for a hydrogen gas sensor. Sens. Actuators B Chem. 2021, 341, 130035.
  30. Abdullah, Q.N.; Ahmed, A.R.; Ali, A.M.; Yam, F.K.; Hassan, Z.; Bououdina, M. Novel SnO2-coated beta-Ga2O3 nanostructures for room temperature hydrogen gas sensor. Int. J. Hydrogen Energy 2021, 46, 7000–7010.
  31. Park, C.H.; Koo, W.-T.; Lee, Y.J.; Kim, Y.H.; Lee, J.; Jang, J.-S.; Yun, H.; Kim, I.-D.; Kim, B.J. Hydrogen Sensors Based on MoS2 Hollow Architectures Assembled by Pickering Emulsion. ACS Nano 2020, 14, 9652–9661.
  32. Jaiswal, J.; Tiwari, P.; Singh, P.; Chandra, R. Fabrication of highly responsive room temperature H2 sensor based on vertically aligned edge-oriented MoS2 nanostructured thin film functionalized by Pd nanoparticles. Sens. Actuators B Chem. 2020, 325, 128800.
  33. Vivekanandan, A.K.; Huang, B.-R.; Kathiravan, D.; Saravanan, A.; Prasannan, A.; Tsai, H.-C.; Chen, S.-H. Effect of MoS2 solution on reducing the wall thickness of ZnO nanotubes to enhance their hydrogen gas sensing properties. J. Alloys Compd. 2021, 854, 157102.
  34. Ha, N.; Xu, K.; Cheng, Y.; Ou, R.; Ma, Q.; Hu, Y.; Vien, T.; Ren, G.; Yu, H.; Zhang, L.; et al. Self-Assembly of Ultrathin Nickel Oxysulfide for Reversible Gas Sensing at Room Temperature. Chemosensors 2022, 10, 372.
  35. Achary, L.S.K.; Maji, B.; Kumar, A.; Ghosh, S.P.; Kar, J.P.; Dash, P. Efficient room temperature detection of H2 gas by novel ZnFe2O4-Pd decorated rGO nanocomposite. Int. J. Hydrogen Energy 2020, 45, 5073–5085.
More
Information
Subjects: Electrochemistry
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: 280
Revisions: 2 times (View History)
Update Date: 10 Jul 2023
1000/1000
Video Production Service