Hostname: page-component-848d4c4894-x5gtn Total loading time: 0 Render date: 2024-05-21T14:31:04.908Z Has data issue: false hasContentIssue false

Photoreduction of Methylviologen in Saponite Clay: Effect of Methylviologen Adsorption Density on the Reaction Efficiency

Published online by Cambridge University Press:  01 January 2024

Takuya Fujimura*
Affiliation:
Department of Materials Chemistry, Graduate School of Natural Science & Technology, Shimane University, 1060, Nishikawatsu-cho, Matsue, Shimane 690-8504, Japan
Tetsuya Shimada
Affiliation:
Department of Applied Chemistry, Graduate School of Urban Environmental Sciences, Tokyo Metropolitan University, 1-1 Minami-osawa, Hachioji, Tokyo 192-0397, Japan Center for Artificial Photosynthesis, Tokyo Metropolitan University, 1-1 Minami-osawa, Hachioji, Tokyo 192-0397, Japan
Ryo Sasai
Affiliation:
Department of Materials Chemistry, Graduate School of Natural Science & Technology, Shimane University, 1060, Nishikawatsu-cho, Matsue, Shimane 690-8504, Japan
Shinsuke Takagi*
Affiliation:
Department of Applied Chemistry, Graduate School of Urban Environmental Sciences, Tokyo Metropolitan University, 1-1 Minami-osawa, Hachioji, Tokyo 192-0397, Japan Center for Artificial Photosynthesis, Tokyo Metropolitan University, 1-1 Minami-osawa, Hachioji, Tokyo 192-0397, Japan
*
*E-mail address of corresponding author: tfujimura@riko.shimane-u.ac.jp
*E-mail address of corresponding author: takagi-shinsuke@tmu.ac.jp

Abstract

To identify the mechanisms for and to estimate the photochemical reaction efficiency of molecules in solid-state host materials is difficult. The objective of the present research was to measure the photogeneration efficiency of the methylviologen cation radical (MV+•) hosted in a semi-transparent hybrid film composed of MV2+ and saponite, a 2:1 clay mineral. MV+• is the one-electron reduced species of MV2+. MV+• was generated by UV irradiation of these films. The fluorescence intensity of MV2+ and the photogeneration efficiency of MV+• depended on the loading level of MV2+. When the loading level of MV2+ was high (75% of the cation exchange capacity (abbreviated as % CEC) of saponite), its fluorescence was reduced considerably because of the self-fluorescence quenching reaction, and the photogeneration efficiency of MV+• was relatively high (quantum yield φ = 3.5×10–2) compared to that of films with low adsorption density (10% CEC, φ = 1.1×10–2). Furthermore, when the loading level of MV2+ was very low (0.13% CEC), a self-fluorescence quenching reaction was not observed and MV+• was not generated. From these observations, one of the principal mechanisms of the self-quenching reaction and MV+• formation in saponite is the electron transfer reaction between excited MV2+ and adjacent MV2+ molecules in the ground state.

Type
Article
Copyright
Copyright © Clay Minerals Society 2019

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

Alvaro, M., García, H., García, S., Márquez, F., & Scaiano, J. C. (1997). Intrazeolite photochemistry. 17. Zeolites as electron donors: Photolysis of methylviologen incorporated within zeolites. The Journal of Physical Chemistry B, 101, 30433051.CrossRefGoogle Scholar
Auerbach, S. M., Carrado, K. A., & Dutta, P. K. (2004). Handbook of Layered Materials, 10. Florida: CRC Press.CrossRefGoogle Scholar
Bahnemann, D. W., Fischer, C.-H., Janata, E., & Henglein, A. (1987). The two-electron oxidation of methyl viologen. Detection and analysis of two fluorescing products. Journal of the Chemical Society, Faraday Transactions 1: Physical Chemistry in Condensed Phases, 83, 2559.CrossRefGoogle Scholar
Bard, A. J., & Fox, M. A. (2002). Artificial photosynthesis: Solar splitting of water to hydrogen and oxygen. Accounts of Chemical Research, 28, 141145.CrossRefGoogle Scholar
Bockman, T. M., & Kochi, J. K. (1990). Isolation and oxidation-reduction of methylviologen cation radicals. Novel disproportionation in charge-transfer salts by X-ray crystallography. The Journal of Organic Chemistry, 55, 41274135.CrossRefGoogle Scholar
Chernia, Z., & Gill, D. (1999). Flattening of tmpyp adsorbed on laponite. Evidence in observed and calculated UV–vis spectra. Langmuir, 15, 16251633.CrossRefGoogle Scholar
Ebbesen, T. W., Manring, L. E., & Peters, K. S. (1984). Picosecond photochemistry of methyl viologen. Journal of the American Chemical Society, 106, 74007404.CrossRefGoogle Scholar
Egawa, T., Watanabe, H., Fujimura, T., Ishida, Y., Yamato, M., Masui, D., Shimada, T., Tachibana, H., Yoshida, H., Inoue, H., & Takagi, S. (2011). Novel methodology to control the adsorption structure of cationic porphyrins on the clay surface using the “size-matching rule”. Langmuir, 27, 1072210729.CrossRefGoogle ScholarPubMed
Inoue, H., Ichiroku, N., Torimoto, T., Sakata, T., Mori, H., & Yoneyama, H. (1994). Photoinduced electron transfer from zinc sulfide microcrystals modified with various alkanethiols to methyl viologen. Langmuir, 10, 45174522.CrossRefGoogle Scholar
Ishida, Y., Masui, D., Shimada, T., Tachibana, H., Inoue, H., & Takagi, S. (2012a). The mechanism of the porphyrin spectral shift on inorganic nanosheets: The molecular flattening induced by the strong host-guest interaction due to the “size-matching rule”. The Journal of Physical Chemistry C, 116, 78797885.CrossRefGoogle Scholar
Ishida, Y., Shimada, T., Tachibana, H., Inoue, H., & Takagi, S. (2012b). Regulation of the collisional self-quenching of fluorescence in clay/porphyrin complex by strong host-guest interaction. The Journal of Physical Chemistry A, 116, 1206512072.CrossRefGoogle Scholar
Ishida, Y., Shimada, T., & Takagi, S. (2014). “Surface-fixation induced emission” of porphyrazine dye by a complexation with inorganic nanosheets. The Journal of Physical Chemistry C, 118, 2046620471.CrossRefGoogle Scholar
Kakegawa, N., Kondo, T., & Ogawa, M. (2003). Variation of electron-donating ability of smectites as probed by photoreduction of methyl viologen. Langmuir, 19, 35783582.CrossRefGoogle Scholar
Kawamata, J., Suzuki, Y., & Tenma, Y. (2010). Fabrication of clay mineral–dye composites as nonlinear optical materials. Philosophical Magazine, 90, 25192527.CrossRefGoogle Scholar
Kodaka, M., & Kubota, Y. (1999). Effect of structures of bipyridinium salts on redox potential and its application to CO2 fixation. Journal of the Chemical Society, Perkin Transactions, 2, 891894.CrossRefGoogle Scholar
Kuykendall, V. G., & Thomas, J. K. (1990). Photophysical investigation of the degree of dispersion of aqueous colloidal clay. Langmuir, 6, 13501356.CrossRefGoogle Scholar
Mao, Y., Breen, N. E., & Thomas, J. K. (1995). Formation of methylviologen radical monopositive cations and ensuing reactions with polychloroalkanes on silica gel surfaces. The Journal of Physical Chemistry, 99, 99099917.CrossRefGoogle Scholar
Matheson, M. S., Lee, P. C., Meisel, D., & Pelizzetti, E. (1983). Kinetics of hydrogen production from methyl viologen radicals on colloidal platinum. The Journal of Physical Chemistry, 87, 394399.CrossRefGoogle Scholar
Miyata, H., Sugahara, Y., Kuroda, K., & Kato, C. (1987). Synthesis of montmorillonite–viologen intercalation compounds and their photo-chromic behaviour. Journal of the Chemical Society, Faraday Transactions 1: Physical Chemistry in Condensed Phases, 83, 1851.CrossRefGoogle Scholar
Miyata, H., Sugahara, Y., Kuroda, K., & Kato, C. (1988). Synthesis of a viologen–tetratitanate intercalation compound and its photochemical behaviour. Journal of the Chemical Society, Faraday Transactions 1: Physical Chemistry in Condensed Phases, 84, 2677.CrossRefGoogle Scholar
Ohtani, Y., Ishida, Y., Ando, Y., Tachibana, H., Shimada, T., & Takagi, S. (2014). Adsorption and photochemical behaviors of the novel cationic xanthene derivative on the clay surface. Tetrahedron Letters, 55, 10241027.CrossRefGoogle Scholar
Palenzuela, J., Vinuales, A., Odriozola, I., Cabanero, G., Grande, H. J., & Ruiz, V. (2014). Flexible viologen electrochromic devices with low operational voltages using reduced graphene oxide electrodes. ACS Applied Materials & Interfaces, 6, 1456214567.CrossRefGoogle ScholarPubMed
Peon, J., Tan, X., Hoerner, J. D., Xia, C., Luk, Y. F., & Kohler, B. (2001). Excited state dynamics of methyl viologen. Ultrafast photo-reduction in methanol and fluorescence in acetonitrile. The Journal of Physical Chemistry A, 105, 57685777.CrossRefGoogle Scholar
Porter, W. W. III, & Vaid, T. P. (2005). Isolation and characterization of phenyl viologen as a radical cation and neutral molecule. The Journal of Organic Chemistry, 70, 50285035.CrossRefGoogle ScholarPubMed
Raupach, M., Emerson, W. W., & Slade, P. G. (1979). The arrangement of paraquat bound by vermiculite and montmorillonite. Journal of Colloid and Interface Science, 69, 398408.CrossRefGoogle Scholar
Rytwo, G., Nir, S., & Margulies, L..(1996). Adsorption and interactions of diquat and paraquat with montmorillonite. Soil Science Society of America Journal, 60, 601.CrossRefGoogle Scholar
Shichi, T., & Takagi, K. (2000). Clay minerals as photochemical reaction fields. Journal of Photochemistry and Photobiology C: Photochemistry Reviews, 1, 113130.CrossRefGoogle Scholar
Solar, S., Solar, W., Getoff, N., Holcman, J., & Sehested, K. (1982). Pulse radiolysis of methyl viologen in aqueous solutions. Journal of the Chemical Society, Faraday Transactions 1: Physical Chemistry in Condensed Phases, 78, 2467.CrossRefGoogle Scholar
Sprick, R. S., Bonillo, B., Clowes, R., Guiglion, P., Brownbill, N. J., Slater, B. J., Blanc, F., Zwijnenburg, M. A., Adams, D. J., & Cooper, A. I. (2016). Visible-light-driven hydrogen evolution using planarized conjugated polymer photocatalysts. Angewandte Chemie International Edition, 55, 17921796.CrossRefGoogle ScholarPubMed
Stevenson, M. J., Marguet, S. C., Schneider, C. R., & Shafaat, H. S. (2017). Light-driven hydrogen evolution by nickel-substituted rubredoxin. ChemSusChem, 10, 17.CrossRefGoogle ScholarPubMed
Suquet, H., Iiyama, J. T., Kodama, H., & Pezerat, H. (1977). Synthesis and swelling properties of saponites with increasing layer charge. Clays and Clay Minerals, 25, 231242.CrossRefGoogle Scholar
Suzuki, Y., Tenma, Y., Nishioka, Y., Kamada, K., Ohta, K., & Kawamata, J. (2011). Efficient two-photon absorption materials consisting of cationic dyes and clay minerals. The Journal of Physical Chemistry C, 115, 2065320661.CrossRefGoogle Scholar
Takagi, S., Tryk, D. A., & Inoue, H. (2002). Photochemical energy transfer of cationic porphyrin complexes on clay surface. Journal of Physical Chemistry B, 106, 54555460.CrossRefGoogle Scholar
Takagi, S., Eguchi, M., Yui, T., & Inoue, H. (2004). Photochemical electron transfer reactions in clay-porphyrin complexes. Clay Science, 12, 8287.Google Scholar
Takagi, S., Shimada, T., Masui, D., Tachibana, H., Ishida, Y., Tryk, D. A., & Inoue, H. (2010). Unique solvatochromism of a membrane composed of a cationic porphyrin-clay complex. Langmuir, 26, 46394641.CrossRefGoogle ScholarPubMed
Takagi, S., Shimada, T., Ishida, Y., Fujimura, T., Masui, D., Tachibana, H., Eguchi, M., & Inoue, H. (2013). Size-matching effect on inorganic nanosheets: Control of distance, alignment, and orientation of molecular adsorption as a bottom-up methodology for nanomaterials. Langmuir, 29, 21082119.CrossRefGoogle ScholarPubMed
Tokieda, D., Tsukamoto, T., Ishida, Y., Ichihara, H., Shimada, T., & Takagi, S. (2017). Unique fluorescence behavior of dyes on the clay minerals surface: Surface fixation induced emission (s-fie). Journal of Photochemistry and Photobiology A: Chemistry, 339, 6779.CrossRefGoogle Scholar
Toshima, N., Kuriyama, M., Yamada, Y., & Hirai, H. (1981). Colloidal platinum catalyst for light-induced hydrogen evolution from water. A particle size effect. Chemistry Letters, 10, 793796.CrossRefGoogle Scholar
Villemure, G., Detellier, C., & Szabo, A. G. (1986). Fluorescence of clay-intercalated methylviologen. Journal of the American Chemical Society, 108, 46584659.CrossRefGoogle Scholar
Villemure, G., Detellier, C., & Szabo, A. G. (1991). Fluorescence of methylviologen intercalated into montmorillonite and hectorite aqueous suspensions. Langmuir, 7, 12151221.CrossRefGoogle Scholar
Wang, Q., Hisatomi, T., Suzuki, Y., Pan, Z., Seo, J., Katayama, M., Minegishi, T., Nishiyama, H., Takata, T., Seki, K., Kudo, A., Yamada, T., & Domen, K. (2017). Particulate photocatalyst sheets based on carbon conductor layer for efficient z-scheme pure-water splitting at ambient pressure. Journal of the American Chemical Society, 139, 16751683.CrossRefGoogle ScholarPubMed
Wasielewski, M. R. (1992). Photoinduced electron transfer in supra-molecular systems for artificial photosynthesis. Chemical Reviews, 92, 435461.CrossRefGoogle Scholar
Watanabe, T., & Honda, K. (1982). Measurement of the extinction coefficient of the methyl viologen cation radical and the efficiency of its formation by semiconductor photocatalysis. The Journal of Physical Chemistry, 86, 26172619.CrossRefGoogle Scholar
Yonemoto, E. H., Riley, R. L., Kim, Y. I., Atherton, S. J., Schmehl, R. H., & Mallouk, T. E. (1992). Photoinduced electron transfer in covalently linked ruthenium tris(bipyridyl)-viologen molecules: Observation of back electron transfer in the Marcus inverted region. Journal of the American Chemical Society, 114, 80818087.CrossRefGoogle Scholar
Supplementary material: File

Fujimura et al. supplementary material
Download undefined(File)
File 1.4 MB