Hostname: page-component-8448b6f56d-cfpbc Total loading time: 0 Render date: 2024-04-17T13:14:58.667Z Has data issue: false hasContentIssue false

Synthesis of WO3−x nanomaterials with controlled morphology and composition for highly efficient photocatalysis

Published online by Cambridge University Press:  01 April 2016

Zhenguang Shen
Affiliation:
School of Engineering and Technology, China University of Geosciences, Beijing 100083, People's Republic of China; and State Key Laboratory of Information Photonics and Optical Communications, and School of Science, Beijing University of Posts and Telecommunications, Beijing 100876, People's Republic of China
Zengying Zhao
Affiliation:
School of Science, China University of Geosciences, Beijing 100083, People's Republic of China
Jingwen Qian
Affiliation:
School of Engineering and Technology, China University of Geosciences, Beijing 100083, People's Republic of China; and State Key Laboratory of Information Photonics and Optical Communications, and School of Science, Beijing University of Posts and Telecommunications, Beijing 100876, People's Republic of China
Zhijian Peng*
Affiliation:
School of Engineering and Technology, China University of Geosciences, Beijing 100083, People's Republic of China
Xiuli Fu*
Affiliation:
State Key Laboratory of Information Photonics and Optical Communications, and School of Science, Beijing University of Posts and Telecommunications, Beijing 100876, People's Republic of China
*
a) Address all correspondence to these authors. e-mail: pengzhijian@cugb.edu.cn
b) e-mail: xiulifu@bupt.edu.cn
Get access

Abstract

Tungsten oxide (WO3−x) nanomaterials with controlled morphology and composition were fabricated by thermal evaporation of WO3 and S powders at different temperatures in a vacuum tube furnace. At 850 °C the obtained green particle is still of the same monoclinic WO3 phase as that of the starting powder. At a temperature between 900 and 1100 °C, the resultant dark-blue products are particle-like clusters composed of numerous monoclinic WO2.90 short nanorods, but the clusters became looser and the nanorods grew somewhat longer as the temperature increased. At a temperature between 1150 and 1250 °C, elongated and thoroughly separate purple-red monoclinic W18O49 nanorods were obtained. The growth of the prepared WO3−x nanomaterials was controlled by a gas–solid mechanism. Their photocatalytic degradation on organic contaminants was evaluated by decomposing methylene blue (MB) in aqueous phase under sunlight, in which WO3 particles presented higher photocatalytic activity than its oxygen-deficient counterparts, WO2.90 and W18O49. But the W18O49 nanorods had higher adsorption ability to MB in all the samples.

Type
Articles
Copyright
Copyright © Materials Research Society 2016 

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

Boulova, M., Gaskov, A., and Lucazeau, G.: Tungsten oxide reactivity versus CH4, CO, and NO2 molecules studied by Raman spectroscopy. Sens. Actuators, B 81, 99 (2001).CrossRefGoogle Scholar
Solis, J.L., Saukko, S., Kish, L., Granqvist, C.G., and Lantto, V.: Semiconductor gas sensors based on nanostructured tungsten oxide. Thin Solid Films 391, 255 (2001).CrossRefGoogle Scholar
Polleux, J., Pinna, N., Antonietti, M., and Niederberger, M.: Growth and assembly of crystalline tungsten oxide nanostructures assisted by bioligation. J. Am. Chem. Soc. 127, 15595 (2005).CrossRefGoogle ScholarPubMed
Xi, G.C., Ye, J.H., Ma, Q., Su, N., Bai, H., and Wang, C.: In situ growth of metal particles on 3D urchin-like WO3 nanostructures. J. Am. Chem. Soc. 134, 6508 (2012).CrossRefGoogle ScholarPubMed
Rausch, B., Symes, M.D., Chisholm, G., and Cronin, L.: Decoupled catalytic hydrogen evolution from a molecular metal oxide redox mediator in water splitting. Science 345, 1326 (2014).CrossRefGoogle ScholarPubMed
Kim, J., Lee, C.W., and Choi, W.: Platinized WO3 as an environmental photocatalyst that generates OH radicals under visible light. Environ. Sci. Technol. 44, 6849 (2010).CrossRefGoogle ScholarPubMed
Szilagyi, I.M., Forizs, B., Rosseler, O., Szegedi, A., Nemeth, P., Kiraly, P., Tarkanyi, G., Vajna, B., Varga-Josepovits, K., Laszlo, K., Toth, A.L., Baranyai, P., and Leskela, M.: WO3 photocatalysts: Influence of structure and composition. J. Catal. 294, 119 (2012).CrossRefGoogle Scholar
Li, X.L., Lou, T.J., Sun, X.M., and Li, Y.D.: Highly sensitive WO3 hollow-sphere gas sensors. Inorg. Chem. 43, 5442 (2004).CrossRefGoogle ScholarPubMed
Wei, H.G., Yan, X.R., Wu, S.J., Luo, Z.P., Wei, S.Y., and Guo, Z.H.: Electropolymerized polyaniline stabilized tungsten oxide nanocomposite films: Electrochromic behavior and electrochemical energy storage. J. Phys. Chem. C 116, 25052 (2012).CrossRefGoogle Scholar
Wu, W.T., Wu, J.J., and Chen, J.S.: Resistive switching behavior and multiple transmittance states in solution-processed tungsten oxide. ACS Appl. Mater. Interfaces 3, 2616 (2011).CrossRefGoogle ScholarPubMed
Wen, J.Q., Li, X., Liu, W., Fang, Y.P., Xie, J., and Xu, Y.H.: Photocatalysis fundamentals and surface modification of TiO2 nanomaterials. Chin. J. Catal. 36, 2049 (2015).CrossRefGoogle Scholar
Asahi, R., Morikawa, T., Ohwaki, T., Aoki, K., and Taga, Y.: Visible-light photocatalysis in nitrogen-doped titanium oxides. Science 293, 269 (2001).CrossRefGoogle ScholarPubMed
Liu, L. and Chen, X.B.: Titanium dioxide nanomaterials: Self-structural modifications. Chem. Rev. 114, 9890 (2014).CrossRefGoogle ScholarPubMed
Chen, X.B., Liu, L., and Huang, F.Q.: Black titanium dioxide (TiO2) nanomaterials. Chem. Soc. Rev. 44, 1861 (2015).CrossRefGoogle ScholarPubMed
Cao, S.W., Low, J.X., Yu, J.G., and Jaroniec, M.: Polymeric photocatalysts based on graphitic carbon nitride. Adv. Mater. 27, 2150 (2015).CrossRefGoogle ScholarPubMed
Wang, Z.Y., Guan, W., Sun, Y.J., Dong, F., Zhou, Y., and Ho, W.K.: Water-assisted production of honeycomb-like g-C3N4 with ultralong carrier lifetime and outstanding photocatalytic activity. Nanoscale 7, 2471 (2015).CrossRefGoogle ScholarPubMed
Foeldvary, C.M. and Wojnarovits, L.: The effect of high-energy radiation on aqueous solution of Acid Red 1 textile dye. Radiat. Phys. Chem. 76, 1485 (2007).CrossRefGoogle Scholar
Chen, Y.P., Liu, S.Y., Yu, H.Q., Yin, H., and Li, Q.R.: Radiation-induced degradation of methyl orange in aqueous solutions. Chemosphere 72, 532 (2008).CrossRefGoogle ScholarPubMed
Qamar, M., Gondal, M.A., and Yamani, Z.H.: Synthesis of highly active nanocrystalline WO3 and its application in laser-induced photocatalytic removal of a dye from water. Catal. Commun. 10, 1980 (2009).CrossRefGoogle Scholar
Rauf, M.A. and Ashraf, S.S.: Radiation induced degradation of dyes-an overview. J. Hazard. Mater. 166, 6 (2009).CrossRefGoogle ScholarPubMed
Fu, H.B., Pan, C.S., Yao, W.Q., and Zhu, Y.F.: Visible-light-induced degradation of rhodamine B by nanosized Bi2WO6 . J. Phys. Chem. B 109, 22432 (2005).CrossRefGoogle ScholarPubMed
Fujishima, A., Zhang, X.T., and Tryk, D.A.: TiO2 photocatalysis and related surface phenomena. Surf. Sci. Rep. 63, 515 (2008).CrossRefGoogle Scholar
Bamwenda, G.R., Sayama, K., and Arakawa, H.: The effect of selected reaction parameters on the photoproduction of oxygen and hydrogen from a WO3–Fe2+–Fe3+ aqueous suspension. J. Photochem. Photobiol., A 122, 175 (1999).CrossRefGoogle Scholar
Wan, L., Sheng, J., Chen, H., and Xu, Y.: Different recycle behavior of Cu2+ and Fe3+ ions for phenol photodegradation over TiO2 and WO3 . J. Hazard. Mater. 262, 114 (2013).CrossRefGoogle ScholarPubMed
Li, X., Yu, J.G., Low, J.X., Fang, Y.P., Xiao, J., and Chen, X.B.: Engineering heterogeneous semiconductors for solar water splitting. J. Mater. Chem. A 3, 2485 (2015).CrossRefGoogle Scholar
Szilagyi, I.M., Madarasz, J., Pokol, G., Kiraly, P., Tarkanyi, G., Saukko, S., Mizsei, J., Toth, A.L., Szabo, A., and Varga-Josepovitso, K.: Stability and controlled composition of hexagonal WO3 . Chem. Mater. 20, 4116 (2008).CrossRefGoogle Scholar
DeJournett, T.J. and Spicer, J.B.: Photoinduced silver precursor decomposition for particle modification in tungsten oxide-polymer matrix nanocomposites. J. Phys. Chem. C 118, 9820 (2014).CrossRefGoogle Scholar
Papynov, E.K., Mayorov, V.Y., Palamarchuk, M.S., and Avramenko, V.A.: Peculiarities of formation of phase composition, porous structure, and catalytic properties of tungsten oxide-based macroporous materials fabricated by sol–gel synthesis. Mater. Charact. 88, 42 (2014).CrossRefGoogle Scholar
Ahmadi, M., Younesi, R., and Guinel, M.J.F.: Synthesis of tungsten oxide nanoparticles using a hydrothermal method at ambient pressure. J. Mater. Res. 29, 1424 (2014).CrossRefGoogle Scholar
Yang, C.Z., van der Laak, N.K., Chan, K.Y., and Zhang, X.: Microwave-assisted microemulsion synthesis of carbon supported Pt-WO3 nanoparticles as an electrocatalyst for methanol oxidation. Electrochim. Acta 75, 262 (2012).CrossRefGoogle Scholar
Qian, J.W., Zhao, Z.Y., Shen, Z.G., Zhang, G.L., Peng, Z.J., and Fu, X.L.: A large scale of CuS nano-networks: Catalyst-free morphologically controllable growth and their application as efficient photocatalysts. J. Mater. Res. 30, 3746 (2015).CrossRefGoogle Scholar
Wu, F.J., Liu, W., Qiu, J.L., Li, J.Z., Zhou, W.Y., Fang, Y.P., Zhang, S.T., and Li, X.: Enhanced photocatalytic degradation and adsorption of methylene blue via TiO2 nanocrystals supported on graphene-like bamboo charcoal. Appl. Surf. Sci. 358, 425 (2015).CrossRefGoogle Scholar
Huang, H.W., Liu, K., Chen, K., Zhang, Y.L., Zhang, Y.H., and Wang, S.C.: Ce and F comodification on the crystal structure and enhanced photocatalytic activity of Bi2WO6 photocatalyst under visible light irradiation. J. Phys. Chem. C 118, 14379 (2014).CrossRefGoogle Scholar
Wagner, R.S. and Ellis, W.C.: Vapor–liquid–solid mechanism of single crystal growth. Appl. Phys. Lett. 4, 89 (1964).CrossRefGoogle Scholar
Chen, J.Y., Wiley, B.J., and Xia, Y.N.: One-dimensional nanostructures of metals: Large-scale synthesis and some potential applications. Langmuir 23, 4120 (2007).CrossRefGoogle ScholarPubMed
Trentler, T.J., Hickman, K.M., Geol, S.C., Viano, A.M., Gibbons, P.C., and Buhro, W.E.: Solution–liquid–solid growth of crystalline III–V semiconductors: An analogy to vapor–liquid–solid growth. Science 270, 1791 (1995).CrossRefGoogle Scholar
Palmisano, L., Augugliaro, V., Sclafani, A., and Schiavello, M.: Activity of chromium-ion-doped titania for the dinitrogen photoreduction to ammonia and for the phenol photodegradation. J. Phys. Chem. 92, 6710 (1988).CrossRefGoogle Scholar
Herrmann, J.M., Disdier, J., and Pichat, P.: Effect of chromium doping on the electrical and catalytic properties of powder titania under UV and visible illumination. Chem. Phys. Lett. 108, 618 (1984).CrossRefGoogle Scholar
Carp, O., Huisman, C.L., and Reller, A.: Photoinduced reactivity of titanium dioxide. Prog. Solid State Chem. 32, 33 (2004).CrossRefGoogle Scholar
Lee, S.H., Cheong, H.M., Tracy, C.E., Mascarenhas, A., Benson, D.K., and Deb, S.K.: Raman spectroscopic studies of electrochromic a-WO3 . Electrochim. Acta 44, 3111 (1999).CrossRefGoogle Scholar
Qian, J.W., Zhao, Z.Y., Shen, Z.G., Zhang, G.L., Peng, Z.J., and Fu, X.L.: Oxide vacancies enhanced visible active photocatalytic W19O55 NMRs via strong adsorption. RSC Adv. 6, 8061 (2016).CrossRefGoogle Scholar
Supplementary material: File

Shen supplementary material

Shen supplementary material

Download Shen supplementary material(File)
File 34.9 MB