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 「MoSET1-dependent transcription factors regulate different stages of infection-related morphogenesis in Pyricularia oryzae」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

MoSET1-dependent transcription factors regulate different stages of infection-related morphogenesis in Pyricularia oryzae

Minh, Dang Ngoc Tsukahara, Yusaku Thach, Dang An Ikeda, Ken-ichi Nakayashiki, Hitoshi 神戸大学

2023.03

概要

MoSET1, an H3K4 histone methyltransferase in Pyricularia oryzae plays a key role in infection-related morphogenesis of the fungus. We previously identified approximately 400 genes for which expression might be regulated directly by MoSET1 during appressorium formation. Here, we focused on five MoSET1-dependent transcription factor (TF) genes that were induced during infection, using a gene deletion approach for three of the five genes (MGG_04699, MGG_06898, MGG_07450) and, for the remaining two genes (MGG_00472 and MGG_07386), a gene silencing technique due to difficulty in constructing a gene deletion mutant. Phenotypic characterization of the knockout and -down mutants revealed that MGG_06898 was crucial for sporulation, MGG_04699 was involved in appressorium formation, and MGG_00472 and MGG_04699 were required for full virulence of the fungus. These results demonstrated that MoSET1 contributes to the pathogenicity of the fungus by controlling transcription factors that further regulate different steps in the infection process of P. oryzae.

この論文で使われている画像

参考文献

276 Aristizabal MJ, Anreiter I, Halldorsdottird T, Odgersc CL, McDadec TW, Goldenberg A,

277 Mostafavi S, Kobor MS, Binder EB, Sokolowski MB, O’Donnell KJ (2019) Biological

278 embedding of experience: A primer on epigenetics. Proc Natl Acad Sci USA 117:23261-

279 23269

280 Boni AC, Ambrosio DL, Cupertino FB, Montenegro-Montero A, Freitas FZ, Corrocher FA,

281 Goncalves RD, Yang A, Weirauch MT, Hughes TR, Larrondo LF, Bertolini MC (2018)

282 Neurospora crassa developmental control mediated by the FLB-3 transcription factor.

283 Fungal Biol 122:570-582

284 Cao H, Huang P, Zhang L, Shi Y, Sun D, Yan Y, Liu X, Dong B, Chen G, Snyder JH, Lin F, Lu

285 J (2016) Characterization of 47 Cys2-His2 zinc finger proteins required for the

286 development and pathogenicity of the rice blast fungus Magnaporthe oryzae. New

287 Phytol 211:1035-1051

288 289 290 291 Catlett NL, Lee B, Yoder OC, Turgeon BG (2003) Split-marker recombination for efficient

targeted deletion of fungal genes. Fungal Genetics Reports 50:9-11

Chen H, Crabb JW, Kinsey JA (1998) The Neurospora aab-1 gene encodes a CCAAT binding

protein homologous to yeast HAP5. Genetics148:123-130

292 Dong Y, Zhao Q, Liu X, Zhang X, Qi Z, Zhang H, Zheng X, Zhang Z (2015) MoMyb1 is

293 required for asexual development and tissue-specific infection in the rice blast fungus

294 Magnaporthe oryzae. BMC Microbiol 15: 37

295 Gomi K, Akeno T, Minetoki T, Ozeki K, Kunagai C, Okazaki N, Iimura Y (2000) Molecular

296 cloning and characterization of a transcriptional cctivator gene, amyR, involved in the

297 amylolytic gene expression in Aspergillus oryzae. Biosci Biotechnol Biochem 64:816-

298 827

299 Hyon GS, Nga NTT, Chuma I, Inoue Y, Asano H, Murata N, Kusaba M, Tosa Y (2012)

300 Characterization of interactions between barley and various host-specific subgroups

301 of Magnaporthe oryzae and M. grisea. J Gen Plant Pathol 78:237-246

302 Jeon J, Lee GW, Kim KT, Park SY, Kim S, Kwon S, Huh A, Chung H, Lee DY, Kim CY, Lee

303 YH (2020) Transcriptome profiling of the Rice blast fungus Magnaporthe oryzae and

304 its host Oryza sativa during infection. Mol Plant Microbe Interact 33:141-144

305 Jones DAB, John E, Rybak K, Phan HTT, Singh KB, Lin SY, Solomon PS, Oliver RP, Tan

306 KC (2019) A specific fungal transcription factor controls effector gene expression and

14

Dang et al. p. 15 307 orchestrates the establishment of the necrotrophic pathogen lifestyle on wheat. Sci

308 Rep 9:15884

309 Kato H, Yamamoto M, Yamaguchi-Ozaki T, Kadouchi H, Iwamoto Y, Nakayashiki H, Tosa Y,

310 Mayama S, Mori N (2000) Pathogenicity, mating ability and DNA restriction fragment

311 length polymorphisms of Pyricularia populations isolated from Gramineae,

312 Bambusideae and Zingiberaceae plants. J Gen Plant Pathol 66:30-47 313 Kim HJ, Han JH, Kim KS, Lee YH (2014) Comparative functional analysis of the velvet gene

314 family reveals unique roles in fungal development and pathogenicity in Magnaporthe

315 oryzae. Fungal Genet Biol 66:33-43

316 Kwon NJ, Garzia A, Espeso EA, Ugalde U, Yu JH (2010) FlbC is a putative nuclear C2H2

317 transcription factor regulating development in Aspergillus nidulans. Mol Microbiol

318 77:1203-1219

319 Lu J, Cao H, Zhang L, Huang P, Lin F (2014) Systematic analysis of Zn2Cys6 transcription

320 factors required for development and pathogenicity by high-throughput gene knockout

321 in the rice blast fungus. PLoS Pathog 10:e1004432

322 Malapi-Wight M, Kim JE, Shim WB (2014) The N-terminus region of the putative C2H2

323 transcription factor Ada1 harbors a specie-specific activation motif that regulates

324 asexual reproduction in Fusarium verticillioides. Fungal Genet Biol 62:25-33

325 Morita Y, Hyon G, Hosogi N, Miyata N, Nakayashiki H, Muranaka Y, Inada N, Park P, Ikeda

326 K (2013) Appressorium-localized NADPH oxidase B is essential for aggressiveness and

327 pathogenicity in the host-specific, toxin-producing fungus Alternaria alternata

328 Japanese pear pathotype. Mol Plant Pathol 14:365-378

329 330 Nakayashiki H, Kiyotomi K, Tosa Y, Mayama S (1999) Transposition of the retrotransposon

MAGGY in heterologous species of filamentous fungi. Genetics 153:693-703

331 Pham KTM, Inoue Y, Vu BV, Nguyen HH, Nakayashiki T, Ikeda K, Nakayashiki H (2015a)

332 MoSET1 (histone H3K4 methyltransferase in Magnaporthe oryzae) regulates global

333 gene expression during infection-related morphogenesis. PLoS Genet 11:e1005385

334 Pham KTM, Nguyen HH, Murai T, Chuma I, Tosa Y, Nakayashiki H (2015b) Histone H3K4

335 methyltransferase globally regulates substrate-dependent activation of cell-wall-

336 degrading enzymes in Magnaporthe oryzae. J Gen Plant Pathol 81: 127-130

337 Shelest E (2008) Transcription factors in fungi. FEMS Microbiology Letters 286:145-151

15

Dang et al. p. 16 338 Talbot NJ (2003) On the trail of a cereal killer: exploring the biology of Magnaporthe

grisea. Ann Rev in Microbiol 57:177-202

339 340 Urashima AS, Hashimoto Y, Don LD, Kusaba M, Tosa Y, Nakayashiki H, Mayama S (1999)

341 Molecular analysis of the wheat blast population in Brazil with a homolog of

342 retrotransposon MGR583. Ann Phytopathol Soc Jpn 65:429-436

343 Vu BV, Takino M, Murata T, Nakayashiki H (2011) Novel vectors for retrotransposon induced

gene silencing in Magnaporthe oryzae . J Gen Plant Pathol 77:147-151

344 345 Vu BV, Pham KT, Nakayashiki H (2013) Substrate-induced transcriptional activation of the

346 MoCel7C cellulase gene is associated with methylation of histone H3 at lysine 4 in the

347 rice blast fungus Magnaporthe oryzae. Appl Environ Microbiol 79:6823-6832

348 349 16

Dang et al. p. 17 350 Figure legends

351 352 Fig. 1 Quantitative RT-PCR analysis of five MoSET1-dependent transcription factors

353 at hyphal, conidial and infectious stages in P. oryzae. Actin was used to normalize

354 mRNA expression level. Data show fold change (relative to mRNA quantity in hyphae)

355 ± standard error (n = 3). Asterisks are given to indicate significant difference at p < 0.05

356 (*) and p < 0.01 (**) (two-tailed t-test).

357 358 Fig. 2 Quantitative RT-PCR analysis of MGG_00472 and MGG_07386 mRNA in

359 candidates of their knock-down mutants. Actin was used to normalize mRNA

360 expression level. Data show fold change (relative to mRNA quantity in the wild-type

361 strain) ± standard error (n = 3). Different characters indicate significant differences by

362 Tukey's HSD (p < 0.05).

363 364 Fig. 3 Phenotypic characterization of knock-out and -down mutants of MoSET1-

365 dependent TFs in P. oryzae. a Diameters of fungal colonies were measured at 9 days

366 after inoculation on rich agar medium. b Conidiation was evaluated by counting the

367 number of conidia under a light microscopy as described in details in Materials and

368 methods. c-d The rates of conidial germination (c) and appressorium formation (d) were

369 measured by observing conidial suspension on hydrophobic surface under a light

370 microscope after 5 h (conidial germination) and 24 h (appressorium formation)

371 incubation at 25̊C.

372 Black bars indicate the wild-type strain Br48 (WT) and grey bars represent knock-out

373 and -down mutants of MoSET1-dependent TFs and their gene complemented strains

374 (c∆mgg_04699 and c∆mgg_06898). Data show fold change (relative to the wild-type

375 strain) ± standard error (n = 3). Different characters in the graphs indicate significant

376 differences by Tukey's HSD (p < 0.05) ND, not determined.

377 378 Fig. 4 Inoculation test of knock-out and -down mutants of MoSET1-dependent TFs in

379 P. oryzae. a Infection assay was performed on the wheat cultivar Norin 4 at 23̊C. Four

380 to five days after inoculation, symptoms on the inoculated plants were evaluated. Letters

17

Dang et al. p. 18 381 under pictures of infected leaves indicate disease index values by a grading method

382 (Hyon et al., 2012). This experiment was repeated at least three times, and

383 representative samples are presented. b The rates of infection hyphae formation in

384 infected leaves. The black bar indicates the wild-type strain Br48 (WT) and grey bars

385 represent knock-out and –down mutants of MoSET1-dependent TFs and a gene

386 complemented strain (c∆mgg_04699). Error bars represent standard errors of the mean

387 (n = 10). Different characters in the graph indicate significant differences by Tukey's

388 HSD (p < 0.05).

389 390 Fig. 5 Schematic diagram of putative MoSET1 regulatory network during infection-

391 related morphogenesis in P. oryzae.

18

Dang et al. p.19

3.E+01

25

Fold change

(relative to Hyphae)

8.E+00

23

**

**

4.E+00

22

2.E+00

21

Hyphae

Spore

infection (5hpi)

infection (12hpi)

**

2.E+01

24

**

**

1.E+00

5.E‐01

2-1

3.E‐01

2-2

**

2-3

1.E‐01

**

2-4

6.E‐02

MGG_00472

MGG_04699

MGG_06898

MGG_07386

**

MGG_07450

3.E‐02

-5

Fig 1. Quantitative RT‐PCR analysis of five MoSET1‐dependent transcription factors at hyphal, conidial and infectious stages in P. oryzae. Actin was used to normalize mRNA expression level. Data show fold change (relative to mRNA quantity in hyphae) ±standard error (n = 3). Asterisks are given to indicate significant difference at p< 0.05 (*) and p< 0.01 (**) (two‐tailed t‐test). Dang et al. p.20 1.2

Fold change

(relative to WT)

MGG_00472

MGG_07386

0.8

0.6

0.4

0.2

Fig 2. Quantitative RT‐PCR analysis of MGG_00472 and MGG_07386 mRNA in candidates of their knock‐down mutants. Actin was used to normalize mRNA expression level. Data show fold change (relative to mRNA quantity in the wild‐type strain) ±standard error (n = 3). a‐c, Different characters indicate significant differences by Tukey's HSD (P < 0.05).

Dang et al. p.21 a, vegetative growth

ab

bc

ab

ab

bc

0.75

0.5

0.25

Fold change (relative to WT)

b, conidiation

ab

0.75

0.5

0.25

c, germination

0.75

0.5

0.25

d, appressorium formation

0.75

0.5

0.25

N.D.

N.D.

N.D.

N.D.

Fig 3. Phenotypic characterization of knock‐out and ‐down mutants of MoSET1‐

dependent TFs in P. oryzae. a, Diameters of fungal colonies were measured at 9 days after inoculation on rich agar medium. b, Conidiation was evaluated by counting the number of conidia under a light microscopy as described in details in Materials and method. c‐d, The rates of conidial germination (c) and appressorium formation (d) were measured by observing conidial suspension on hydrophobic surface under a light microscope after 5 h (conidial germination) and 24 h (appressorium formation) incubation at 25̊C. Black bars indicate the wild‐type strain Br48 (WT) and grey bars represent knock‐

out and –down mutants of MoSET1‐dependent TFs and their gene complemented strains (c∆mgg_04699 and c∆mgg_06898). Data show fold change (relative to the wild‐type strain) ±standard error (n = 3). a‐d, Different characters in the graphs indicate significant differences by Tukey's HSD (P < 0.05). ND, not determined.

Dang et al. p.22 a

5G

2-3BG

5G

3-4BG

5B

5G

ab

(%)

100

Rates of infection

hyphae formation

80

60

40

20

Fig. 4. Inoculation test of knock‐out and ‐down mutants of MoSET1‐

dependent TFs in P. oryzae. a, Infection assay was performed on the wheat cultivar Norin 4 at 23̊C. Four to five days after inoculation, symptoms on the inoculated plants were evaluated. Letters under pictures of infected leaves indicate disease index values by a grading method (Hyon et al., 2012). This experiment was repeated at least three times, and representative samples are presented. b, The rates of infection hyphae formation in infected leaves. The black bar indicates the wild‐type strain Br48 (WT) and grey bars represent knock‐out and –

down mutants of MoSET1‐dependent TFs and a gene complemented strain (c∆mgg_04699). Error bars represent standard errors of the mean (n = 10). a‐d, Different characters in the graph indicate significant differences by Tukey's HSD (P < 0.05). Dang et al. p.23 MoSET1

TFs

vegetative growth

sporulation

appressorium formation

Fig. 5. Schematic diagram of putative MoSET1 regulatory network during infection‐related morphogenesis in P. oryzae.

virulence

...

参考文献をもっと見る

全国の大学の
卒論・修論・学位論文

一発検索!

この論文の関連論文を見る