リケラボ論文検索は、全国の大学リポジトリにある学位論文・教授論文を一括検索できる論文検索サービスです。

リケラボ 全国の大学リポジトリにある学位論文・教授論文を一括検索するならリケラボ論文検索大学・研究所にある論文を検索できる

リケラボ 全国の大学リポジトリにある学位論文・教授論文を一括検索するならリケラボ論文検索大学・研究所にある論文を検索できる

大学・研究所にある論文を検索できる 「Identification of novel canonical strigolactones produced by tomato」の論文概要。リケラボ論文検索は、全国の大学リポジトリにある学位論文・教授論文を一括検索できる論文検索サービスです。

コピーが完了しました

URLをコピーしました

論文の公開元へ論文の公開元へ
書き出し

Identification of novel canonical strigolactones produced by tomato

Wakabayashi, Takatoshi Moriyama, Daisuke Miyamoto, Ayumi Okamura, Hironori Shiotani, Nanami Shimizu, Nobuhiro Mizutani, Masaharu Takikawa, Hirosato Sugimoto, Yukihiro 神戸大学

2022.12.14

概要

Canonical strigolactones (SLs), such as orobanchol, consist of a tricyclic lactone ring (ABC-ring) connected to a methylbutenolide (D-ring). Tomato plants have been reported to produce not only orobanchol but also various canonical SLs related to the orobanchol structure, including orobanchyl acetate, 7-hydroxyorobanchol isomers, 7-oxoorobanchol, and solanacol. In addition to these, structurally unidentified SL-like compounds known as didehydroorobanchol isomers (DDHs), whose molecular mass is 2 Da smaller than that of orobanchol, have been found. Although the SL biosynthetic pathway in tomato is partially characterized, structural elucidation of DDHs is required for a better understanding of the entire biosynthetic pathway. In this study, three novel canonical SLs with the same molecular mass as DDHs were identified in tomato root exudates. The first was 6,7-didehydroorobanchol, while the other two were not in the DDH category. These two SLs were designated phelipanchol and epiphelipanchol because they induced the germination of Phelipanche ramosa, a noxious root parasitic weed of tomato. We also proposed a putative biosynthetic pathway incorporating these novel SLs from orobanchol to solanacol.

この論文で使われている画像

参考文献

Abe, S., Sado, A., Tanaka, K., Kisugi, T., Asami, K., Ota, S., et al. (2014). Carlactone is converted to carlactonoic acid by MAX1 in Arabidopsis and its methyl ester can directly interact with AtD14 in vitro. Proc. Natl. Acad. Sci. 111, 18084–18089. doi: 10.1073/pnas.1410801111

Akiyama, K., Matsuzaki, K., and Hayashi, H. (2005). Plant sesquiterpenes induce hyphal branching in arbuscular mycorrhizal fungi. Nature 435, 824–827. doi: 10.1038/nature03608

Akiyama, K., Ogasawara, S., Ito, S., and Hayashi, H. (2010). Structural requirements of strigolactones for hyphal branching in AM fungi. Plant Cell Physiol. 51, 1104–1117. doi: 10.1093/pcp/pcq058

Alder, A., Jamil, M., Marzorati, M., Bruno, M., Vermathen, M., Bigler, P., et al. (2012). The path from b-carotene to carlactone, a strigolactone-like plant hormone. Science 335, 1348–1351. doi: 10.1126/science.1218094

Aliche, E. B., Screpanti, C., De Mesmaeker, A., Munnik, T., and Bouwmeester, H. J. (2020). Science and application of strigolactones. New Phytol. 227, 1001–1011. doi: 10.1111/nph.16489

Aquino, B., Bradley, J. M., and Lumba, S. (2021). On the outside looking in: roles of endogenous and exogenous strigolactones. Plant J. 105, 322–334. doi: 10.1111/tpj.15087

Cook, C. E., Whichard, L. P., Turner, B., Wall, M. E., and Egley, G. H. (1966).

Germination of witchweed (Striga lutea lour.): Isolation and properties of a potent stimulant. Science 154, 1189–1190. doi: 10.1126/science.154.3753.1189 Gomez-Roldan, V., Fermas, S., Brewer, P. B., Puech-Pagès, V., Dun, E. A., Pillot, J.-P., et al. (2008). Strigolactone inhibition of shoot branching. Nature 455, 189–194. doi: 10.1038/nature07271

Khetkam, P., Xie, X., Kisugi, T., Kim, H. I., Yoneyama, K., Uchida, K., et al. (2014). 7a- and 7b-hydroxyorobanchyl acetate as germination stimulants for root parasitic weeds produced by cucumber. J. Pestic. Sci. 39, 121–126. doi: 10.1584/ jpestics.D14-038

Kohlen, W., Charnikhova, T., Bours, R., López-Ráez, J. A., and Bouwmeester, H. (2013). Tomato strigolactones: A more detailed look. Plant Signal. Behav. 8, e22785. doi: 10.4161/psb.22785

Koltai, H., LekKala, S. P., Bhattacharya, C., Mayzlish-Gati, E., Resnick, N., Wininger, S., et al. (2010). A tomato strigolactone-impaired mutant displays aberrant shoot morphology and plant interactions. J. Exp. Bot. 61, 1739–1749. doi: 10.1093/jxb/erq041

López-Ráez, J. A., Charnikhova, T., Mulder, P., Kohlen, W., Bino, R., Levin, I., et al. (2008). Susceptibility of the tomato mutant high pigment-2dg (hp-2dg) to Orobanche spp. infection. J. Agric. Food Chem. 56, 6326–6332. doi: 10.1021/jf800760x

Matsui, J., Yokota, T., Bando, M., Takeuchi, Y., and Mori, K. (1999). Synthesis and structure of orobanchol, the germination stimulant for Orobanche minor. Eur. J. Org. Chem. 1999, 2201–2210. doi: 10.1002/(SICI)1099-0690(199909) 1999:9<2201::AID-EJOC2201>3.0.CO;2-Q

Moriyama, D., Wakabayashi, T., Shiotani, N., Yamamoto, S., Furusato, Y., Yabe, K., et al. (2022). Identification of 6-epi-heliolactone as a biosynthetic precursor of avenaol in Avena strigosa. Biosci. Biotechnol. Biochem. 86, 998–1003. doi: 10.1093/bbb/zbac069

Nomura, S., Nakashima, H., Mizutani, M., Takikawa, H., and Sugimoto, Y. (2013). Structural requirements of strigolactones for germination induction and inhibition of Striga gesnerioides seeds. Plant Cell Rep. 32, 829–838. doi: 10.1007/ s00299-013-1429-y

Parker, C. (2013). “The parasitic weeds of the Orobanchaceae,” in Parasitic Orobanchaceae. Eds. D. M. Joel, J. Gressel and L. J. Musselman (Berlin, Heidelberg: Springer Berlin Heidelberg), 313–344. doi: 10.1007/978-3-642-38146-1_18

Sato, D., Awad, A. A., Chae, S. H., Yokota, T., Sugimoto, Y., Takeuchi, Y., et al. (2003). Analysis of strigolactones, germination stimulants for Striga and Orobanche, by high-performance liquid chromatography/tandem mass spectrometry. J. Agric. Food Chem. 51, 1162–1168. doi: 10.1021/jf025997z

Seto, Y., Sado, A., Asami, K., Hanada, A., Umehara, M., Akiyama, K., et al. (2014). Carlactone is an endogenous biosynthetic precursor for strigolactones. Proc. Natl. Acad. Sci. 111, 1640–1645. doi: 10.1073/pnas.1314805111

Shiotani, N., Wakabayashi, T., Ogura, Y., Sugimoto, Y., and Takikawa, H. (2021). Studies on strigolactone BC-ring formation: Chemical conversion of an 18-hydroxycarlactonoate derivative into racemic 4-deoxyorobanchol/5- deoxystrigol via the acid-mediated cascade cyclization. Tetrahedron. Lett. 68, 152922. doi: 10.1016/j.tetlet.2021.152922

Sugimoto, Y., Ali, A. M., Yabuta, S., Kinoshita, H., Inanaga, S., and Itai, A. (2003). Germination strategy of Striga hermonthica involves regulation of ethylene biosynthesis. Physiol. Plant 119, 137–145. doi: 10.1034/j.1399-3054.2003.00162.x

Tokunaga, T., Hayashi, H., and Akiyama, K. (2015). Medicaol, a strigolactone identified as a putative didehydro-orobanchol isomer, from Medicago truncatula. Phytochemistry 111, 91–97. doi: 10.1016/j.phytochem.2014.12.024

Ueno, K., Fujiwara, M., Nomura, S., Mizutani, M., Sasaki, M., Takikawa, H., et al. (2011). Structural requirements of strigolactones for germination induction of Striga gesnerioides seeds. J. Agric. Food Chem. 59, 9226–9231. doi: 10.1021/jf202418a

Ueno, K., Furumoto, T., Umeda, S., Mizutani, M., Takikawa, H., Batchvarova, R., et al. (2014). Heliolactone, a non-sesquiterpene lactone germination stimulant for root parasitic weeds from sunflower. Phytochemistry 108, 122–128. doi: 10.1016/ j.phytochem.2014.09.018

Umehara, M., Hanada, A., Yoshida, S., Akiyama, K., Arite, T., Takeda-Kamiya, N., et al. (2008). Inhibition of shoot branching by new terpenoid plant hormones. Nature 455, 195–200. doi: 10.1038/nature07272

Wakabayashi, T., Hamana, M., Mori, A., Akiyama, R., Ueno, K., Osakabe, K., et al. (2019). Direct conversion of carlactonoic acid to orobanchol by cytochrome P450 CYP722C in strigolactone biosynthesis. Sci. Adv. 5, eaax9067. doi: 10.1126/ sciadv.aax9067

Wakabayashi, T., Shida, K., Kitano, Y., Takikawa, H., Mizutani, M., and Sugimoto, Y. (2020). CYP722C from Gossypium arboreum catalyzes the conversion of carlactonoic acid to 5-deoxystrigol. Planta 251, 97. doi: 10.1007/s00425-020-03390-6

Wakabayashi, T., Ueno, K., and Sugimoto, Y. (2022). Structure elucidation and biosynthesis of orobanchol. Front. Plant Sci. 0. doi: 10.3389/fpls.2022.835160

Wang, Y., Durairaj, J., Suá rez Duran, H. G., van Velzen, R., Flokova, K., Liao, C.- Y., et al. (2022). The tomato cytochrome P450 CYP712G1 catalyses the double oxidation of orobanchol en route to the rhizosphere signalling strigolactone, solanacol. New Phytol. 235, 1884–1899. doi: 10.1111/nph.18272

Xie, X., Kusumoto, D., Takeuchi, Y., Yoneyama, K., Yamada, Y., and Yoneyama, K. (2007). 2′-epi-orobanchol and solanacol, two unique strigolactones, germination stimulants for root parasitic weeds, produced by tobacco. J. Agric. Food Chem. 55, 8067–8072. doi: 10.1021/jf0715121

Yoneyama, K., Mori, N., Sato, T., Yoda, A., Xie, X., Okamoto, M., et al. (2018). Conversion of carlactone to carlactonoic acid is a conserved function of MAX1 homologs in strigolactone biosynthesis. New Phytol. 218, 1522–1533. doi: 10.1111/ nph.15055

Zhang, Y., Cheng, X., Wang, Y., Dıez-Simo ́ ́n, C., Flokova, K., Bimbo, A., et al. (2018). The tomato MAX1 homolog, SlMAX1, is involved in the biosynthesis of tomato strigolactones from carlactone. New Phytol. 219, 297–309. doi: 10.1111/ nph.15131

Zhang, Y., van Dijk, A. D. J., Scaffidi, A., Flematti, G. R., Hofmann, M., Charnikhova, T., et al. (2014). Rice cytochrome P450 MAX1 homologs catalyze distinct steps in strigolactone biosynthesis. Nat. Chem. Biol. 10, 1028–1033. doi: 10.1038/nchembio.1660

参考文献をもっと見る