Hostname: page-component-6b989bf9dc-wj8jn Total loading time: 0 Render date: 2024-04-14T12:01:25.035Z Has data issue: false hasContentIssue false

Enhancing microwave absorption of TiO2 nanocrystals via hydrogenation

Published online by Cambridge University Press:  09 September 2014

Ting Xia*
Affiliation:
Department of Chemistry, University of Missouri – Kansas City, Kansas City, Missouri 64110, USA
Chi Zhang
Affiliation:
Department of Chemistry, University of Missouri – Kansas City, Kansas City, Missouri 64110, USA
Nathan A. Oyler
Affiliation:
Department of Chemistry, University of Missouri – Kansas City, Kansas City, Missouri 64110, USA
Xiaobo Chen*
Affiliation:
Department of Chemistry, University of Missouri – Kansas City, Kansas City, Missouri 64110, USA
*
a)Address all correspondence to this author. e-mail: chenxiaobo@umkc.edu
Get access

Abstract

TiO2 has attracted tremendous research interest for photocatalytic water splitting, solar hydrogen generation, environmental pollution removal, dye-sensitized solar cells, lithium-ion batteries, supercapacitors, and field emission. Microwave absorption materials (MAMs) play important roles in many military (e.g., the stealth coating on the B-2 bomber) and civil (e.g., telecommunications, noise reduction, information security, signal, and data protection) applications. However, TiO2 is not a good MAM due to its poor absorption in the microwave region. Here, we report that via hydrogenation excellent and tunable microwave absorption is achieved with hydrogenated TiO2 nanocrystals. After hydrogenation, 4.3x and 103x improvements have been obtained in storing and dissipating the electric energy of the microwave electromagnetic field. Their permittivity values are higher than those of the current carbonaceous MAMs. Instead of relying on the dipole rotation or ferromagnetic resonance mechanisms for traditional MAMs, the hydrogenated TiO2 nanocrystals work as good MAMs based on a newly proposed collective-movement-of-interfacial-dipole (CMID) mechanism. Although there is still no direct physical evidence of the interface effects of the CMID mechanism, the CMID as a hypothesis at this point successfully explained the origin of the enhanced microwave absorption of the hydrogenated TiO2 nanoparticles. This study thus may open new applications for TiO2 nanocrystals and also stimulate new approaches for new MAM development.

Type
Articles
Copyright
Copyright © Materials Research Society 2014 

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

REFERENCES

Fujishima, A. and Honda, K.: Electrochemical photolysis of water at a semiconductor electrode. Nature 238, 37 (1972).CrossRefGoogle Scholar
Wang, X. and Shi, J.: Evolution of titanium dioxide one-dimensional nanostructures from surface-reaction-limited pulsed chemical vapor deposition. J. Mater. Res. 28, 270 (2013).CrossRefGoogle Scholar
Joo, J.B., Zhang, Q., Dahl, M., Zaera, F., and Yin, Y.: Synthesis, crystallinity control, and photocatalysis of nanostructured titanium dioxide shells. J. Mater. Res. 28, 362 (2013).Google Scholar
Chen, X., Li, C., Grätzel, M., Kostecki, R., and Mao, S.S.: Nanomaterials for renewable energy production and storage. Chem. Soc. Rev. 41, 7909 (2012).CrossRefGoogle ScholarPubMed
Chen, X., Shen, S., Guo, L., and Mao, S.S.: Semiconductor-based photocatalytic hydrogen generation. Chem. Rev. 110, 6503 (2010).Google Scholar
Ma, Y., Xu, Q., Chong, R., and Li, C.: Photocatalytic H2 production on TiO2 with tuned phase structure via controlling the phase transformation. J. Mater. Res. 28, 394 (2013).Google Scholar
Wu, X., Lu, G., and Wang, L.: The effect of photoanode thickness on the performance of dye-sensitized solar cells containing TiO2 nanosheets with exposed reactive {001} facets. J. Mater. Res. 28, 475 (2013).Google Scholar
Chen, X. and Mao, S.S.: Titanium dioxide nanomaterials: Synthesis, properties, modifications, and applications. Chem. Rev. 107, 2891 (2007).CrossRefGoogle ScholarPubMed
Liang, F., Zhang, J., Zheng, L., Tsang, C-K., Li, H., Shu, S., Cheng, H., and Li, Y.Y.: Selective electrodeposition of Ni into the intertubular voids of anodic TiO2 nanotubes for improved photocatalytic properties. J. Mater. Res. 28, 405 (2013).CrossRefGoogle Scholar
Chen, X., Liu, L., Yu, P.Y., and Mao, S.S.: Increasing solar absorption for photocatalysis with black hydrogenated titanium dioxide nanocrystals. Science 331, 746 (2011).Google Scholar
Choi, W., Termin, A., and Hoffmann, M.R.: The role of metal ion dopants in quantum-sized TiO2: Correlation between photoreactivity and charge carrier recombination dynamics. J. Phys. Chem. 98, 13669 (1994).Google Scholar
Chen, X. and Burda, C.: The electronic origin of the visible-light absorption properties of C-, N- and S-doped TiO2 nanomaterials. J. Am. Chem. Soc. 130, 5018 (2008).CrossRefGoogle Scholar
Baker, D.R. and Kamat, P.V.: Photosensitization of TiO2 nanostructures with CdS quantum dots: Particulate versus tubular support architectures. Adv. Funct. Mater. 19, 805 (2009).Google Scholar
Xia, T. and Chen, X.: Revealing the structural properties of hydrogenated black TiO2 nanocrystals. J. Mater. Chem. A 1, 2983 (2013).Google Scholar
Chen, X., Liu, L., Liu, Z., Marcus, M.A., Wang, W-C., Oyler, N.A., Grass, M.E., Mao, B., Glans, P-A., Yu, P.Y., Guo, J., and Mao, S.S.: Properties of disorder-engineered black titanium dioxide nanoparticles through hydrogenation. Sci. Rep. 3, 1510 (2013).CrossRefGoogle ScholarPubMed
Liu, L., Yu, P.P., Chen, X., Mao, S.S., and Shen, D.Z.: Hydrogenation and disorder in engineered black TiO2 . Phys. Rev. Lett. 111, 065505 (2013).CrossRefGoogle ScholarPubMed
Shen, L., Uchaker, E., Zhang, X., and Cao, G.: Hydrogenated Li4Ti5O12 nanowire arrays for high rate lithium ion batteries. Adv. Mater. 24, 6502 (2012).Google Scholar
Xia, T., Zhang, W., Li, W., Oyler, N.A., Liu, G., and Chen, X.: Hydrogenated surface disorder enhances lithium ion battery performance. Nano Energy 2, 826 (2013).CrossRefGoogle Scholar
Lu, X., Wang, G., Zhai, T., Yu, M., Gan, J., Tong, Y., and Li, Y.: Hydrogenated TiO2 nanotube arrays for supercapacitors. Nano Lett. 12, 1690 (2012).Google Scholar
Zhang, C., Yu, H., Li, Y., Gao, Y., Zhao, Y., Song, W., Shao, Z., and Yi, B.: Supported noble metals on hydrogen-treated TiO2 nanotube arrays as highly ordered electrodes for fuel cells. ChemSusChem 6, 659 (2013).Google Scholar
Zhu, W-D., Wang, C-W., Chen, J-B., Li, D-S., Zhou, F., and Zhang, H-L.: Enhanced field emission from hydrogenated TiO2 nanotube arrays. Nanotechnology 23, 455204 (2012).Google Scholar
Micheli, D.: Radar Absorbing Materials and Microwave Shielding Structures Design (LAP LAMBERT Academic Publishing, Germany, 2012).Google Scholar
Zhao, D-L., Li, X., and Shen, Z-M.: Microwave absorbing property and complex permittivity and permeability of epoxy composites containing Ni-coated and Ag filled carbon nanotubes. Compos. Sci. Technol. 68, 2902 (2008).Google Scholar
Bowler, N.: Designing dielectric loss at microwave frequencies using multi-layered filler particles in a composite. IEEE Trans. Dielectr. Electr. Insul. 13, 703 (2006).Google Scholar
Petrov, V.M. and Gagulin, V.V.: Microwave absorbing materials. Inorg. Mater. 37, 93 (2001).Google Scholar
Feng, Y-B., Qiu, T., Shen, C-Y., and Li, X.Y.: Electromagnetic and absorption properties of carbonyl iron/rubber radar absorbing materials. IEEE Trans. Magn. 42, 363 (2006).Google Scholar
Peng, C-H., Wang, H-W., Kan, S-W., Shen, M-Z., Wei, Y-M., and Chen, S-Y.: Microwave absorbing materials using Ag–NiZn ferrite core–shell nanopowders as fillers. J. Magn. Magn. Mater. 284, 113 (2004).CrossRefGoogle Scholar
Xia, T., Zhang, C., Oyler, N.A., and Chen, X.: Hydrogenated TiO2 nanocrystals: A novel microwave absorbing material. Adv. Mater. 25, 6905 (2013).Google Scholar
Yu, P.Y. and Cardona, M.: Fundamentals of Semiconductors: Physics and Materials Properties (Springer, Berlin, 2001).Google Scholar
Diebold, U.: The surface science of titanium dioxide. Surf. Sci. Rep. 48, 53 (2003).Google Scholar
Xia, T., Otto, J.W., Dutta, T., Murowchick, J., Caruso, A.N., Peng, Z., and Chen, X.: Formation of TiO2 nanomaterials via titanium ethylene glycolide decomposition. J. Mater. Res. 28, 326 (2013).Google Scholar
Marchioro, A., Teuscher, J., Friedrich, D., Kunst, M., van de Krol, R., Moehl, T., Grätzel, M., and Moser, J-E.: Unravelling the mechanism of photoinduced charge transfer processes in lead iodide perovskite solar cells. Nat. Photonics 8, 250 (2014), doi: 10.1038/nphoton.2013.374.Google Scholar
Shin, J-Y., Joo, J.H., Samuelis, D., and Maier, J.: Oxygen-deficient TiO2−δ nanoparticles via hydrogen reduction for high rate capability lithium batteries. Chem. Mater. 24, 543 (2012).Google Scholar
Xia, T., Zhang, W., Murowchick, J., Liu, G., and Chen, X.: Built-in electric field-assisted surface-amorphized nanocrystals for high-rate lithium-ion battery. Nano Lett. 13, 5289 (2013).Google Scholar
Spurr, R.A. and Myers, H.: Quantitative analysis of anatase-rutile mixtures with an x-ray diffractometer. Anal. Chem. 29, 760 (1957).Google Scholar
Jenkins, R. and Snyder, R.L.: Introduction to X-ray Powder Diffractometry (John Wiley & Sons Inc., New York, 1996).Google Scholar
Ortega, J., Kodas, T.T., Chadda, S., Smith, D.M., Ciftcioglu, M., and Brennan, J.E.: Formation of dense barium calcium titanate (Ba0.86Ca0.14TiO3) particles by aerosol decomposition. Chem. Mater. 3, 746 (1991).Google Scholar
Hurum, D.C., Agrios, A.G., Gray, K.A., Rajh, T., and Thurnauer, M.C.: Explaining the enhanced photocatalytic activity of Degussa P25 mixed-phase TiO2 using EPR. J. Phys. Chem. B 107, 4545 (2003).Google Scholar
Zhang, J., Xu, Q., Feng, Z., Li, M., and Li, C.: Importance of the relationship between surface phases and photocatalytic activity of TiO2 . Angew. Chem. Int. Ed. 47, 1766 (2008).Google Scholar
Zhang, Y., Harris, C.X., Wallenmayer, P., Murowchick, J., and Chen, X.: Asymmetric lattice vibrational characteristics of rutile TiO2 as revealed by laser power dependent Raman spectroscopy. J. Phys. Chem. C 117, 24015 (2013).Google Scholar
Xia, T., Li, N., Zhang, Y., Kruger, M.B., Murowchick, J., Selloni, A., and Chen, X.: Directional heat dissipation across the interface in anatase–rutile nanocomposites. ACS Appl. Mater. Interfaces 5, 9883 (2013).Google Scholar
Matsumoto, M. and Miyata, Y.: Thin electromagnetic wave absorber for quasi-microwave band containing aligned thin magnetic metal particles. IEEE Trans. Magn. 33, 4459 (1997).CrossRefGoogle Scholar
Li, G., Li, L., Boerio-Goates, J., and Woodfield, B.F.: High purity anatase TiO2 nanocrystals: Near room-temperature synthesis, grain growth kinetics, and surface hydration. J. Am. Chem. Soc. 127, 8659 (2005).CrossRefGoogle ScholarPubMed
Jonsen, P.: A 1H NMR study of reduced Ru/TiO2 and TiO2 . Colloids Surf. 36, 127 (1989).CrossRefGoogle Scholar
Spahr, E.J., Wen, L., Stavola, M., Boatner, L.A., Feldman, L.C., Tolk, N.H., and Lüpke, G.: Giant enhancement of hydrogen transport in rutile TiO2 at low temperatures. Phys. Rev. Lett. 104, 205901 (2010).Google Scholar
Bityurin, N., Kuznetsov, A.I., and Kanaev, A.: Kinetics of UV induced darkening of titanium-oxide gels. Appl. Surf. Sci. 248, 86 (2005).Google Scholar
Komaguchi, K., Maruoka, T., Nakano, H., Imae, I., Ooyama, Y., and Harima, Y.: Electron-transfer reaction of oxygen species on TiO2 nanoparticles induced by sub-band-gap illumination. J. Phys. Chem. C 114, 1240 (2010).CrossRefGoogle Scholar
Zuo, F., Wang, L., Wu, T., Zhang, Z., Borchardt, D., and Feng, P.: Self-doped Ti3+ enhanced photocatalyst for hydrogen production under visible light. J. Am. Chem. Soc. 132, 11856 (2010).Google Scholar
Supplementary material: File

Supplementary Material

Supplementary information supplied by authors.

Download Supplementary Material(File)
File 927.2 KB