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 「Minimal upstream open reading frame of Per2 mediates phase fitness of the circadian clock to day/night physiological body temperature rhythm」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

Minimal upstream open reading frame of Per2 mediates phase fitness of the circadian clock to day/night physiological body temperature rhythm

Miyake, Takahito Inoue, Yuichi Shao, Xinyan Seta, Takehito Aoki, Yuto Nguyen Pham, Khanh Tien Shichino, Yuichi Sasaki, Junko Sasaki, Takehiko Ikawa, Masahito Yamaguchi, Yoshiaki Okamura, Hitoshi Iwasaki, Shintaro Doi, Masao 京都大学 DOI:10.1016/j.celrep.2023.112157

2023.03.28

概要

Body temperature in homeothermic animals does not remain constant but displays a regular circadian fluctuation within a physiological range (e.g., 35°C–38.5°C in mice), constituting a fundamental systemic signal to harmonize circadian clock-regulated physiology. Here, we find the minimal upstream open reading frame (uORF) encoded by the 5′ UTR of the mammalian core clock gene Per2 and reveal its role as a regulatory module for temperature-dependent circadian clock entrainment. A temperature shift within the physiological range does not affect transcription but instead increases translation of Per2 through its minimal uORF. Genetic ablation of the Per2 minimal uORF and inhibition of phosphoinositide-3-kinase, lying upstream of temperature-dependent Per2 protein synthesis, perturb the entrainment of cells to simulated body temperature cycles. At the organismal level, Per2 minimal uORF mutant skin shows delayed wound healing, indicating that uORF-mediated Per2 modulation is crucial for optimal tissue homeostasis. Combined with transcriptional regulation, Per2 minimal uORF-mediated translation may enhance the fitness of circadian physiology.

この論文で使われている画像

参考文献

1. Brown, S.A., Zumbrunn, G., Fleury-Olela, F., Preitner, N., and Schibler, U.

(2002). Rhythms of mammalian body temperature can sustain peripheral

circadian clocks. Curr. Biol. 12, 1574–1583. https://doi.org/10.1016/

s0960-9822(02)01145-4.

2. Buhr, E.D., Yoo, S.H., and Takahashi, J.S. (2010). Temperature as a universal resetting cue for mammalian circadian oscillators. Science 330,

379–385. https://doi.org/10.1126/science.1195262.

3. Patke, A., Murphy, P.J., Onat, O.E., Krieger, A.C., O¨zc¸elik, T., Campbell,

S.S., and Young, M.W. (2017). Mutation of the human circadian clock

gene CRY1 in familial delayed sleep phase disorder. Cell 169, 203–

215.e13. https://doi.org/10.1016/j.cell.2017.03.027.

4. Asher, G., and Schibler, U. (2011). Crosstalk between components of

circadian and metabolic cycles in mammals. Cell Metab. 13, 125–137.

https://doi.org/10.1016/j.cmet.2011.01.006.

5. Schibler, U., Gotic, I., Saini, C., Gos, P., Curie, T., Emmenegger, Y., Sinturel, F., Gosselin, P., Gerber, A., Fleury-Olela, F., et al. (2015). Clocktalk: interactions between central and peripheral circadian oscillators in

mammals. Cold Spring Harb. Symp. Quant. Biol. 80, 223–232. https://

doi.org/10.1101/sqb.2015.80.027490.

6. Welz, P.S., Zinna, V.M., Symeonidi, A., Koronowski, K.B., Kinouchi, K.,

Smith, J.G., Guille´n, I.M., Castellanos, A., Furrow, S., Arago´n, F., et al.

(2019). BMAL1-driven tissue clocks respond independently to light to

14. Zhang, H., Wang, Y., and Lu, J. (2019). Function and evolution of upstream

ORFs in eukaryotes. Trends Biochem. Sci. 44, 782–794. https://doi.org/

10.1016/j.tibs.2019.03.002.

16. Ingolia, N.T., Ghaemmaghami, S., Newman, J.R.S., and Weissman, J.S.

(2009). Genome-wide analysis in vivo of translation with nucleotide resolution using ribosome profiling. Science 324, 218–223. https://doi.org/10.

1126/science.1168978.

17. Chothani, S.P., Adami, E., Widjaja, A.A., Langley, S.R., Viswanathan, S.,

Pua, C.J., Zhihao, N.T., Harmston, N., D’Agostino, G., Whiffin, N., et al.

(2022). A high-resolution map of human RNA translation. Mol. Cell 82,

2885–2899.e8. https://doi.org/10.1016/j.molcel.2022.06.023.

18. Zheng, B., Larkin, D.W., Albrecht, U., Sun, Z.S., Sage, M., Eichele, G., Lee,

C.C., and Bradley, A. (1999). The mPer2 gene encodes a functional

component of the mammalian circadian clock. Nature 400, 169–173.

https://doi.org/10.1038/22118.

19. Lee, C., Etchegaray, J.P., Cagampang, F.R., Loudon, A.S., and Reppert,

S.M. (2001). Posttranslational mechanisms regulate the mammalian circadian clock. Cell 107, 855–867. https://doi.org/10.1016/s0092-8674(01)

00610-9.

20. Zheng, B., Albrecht, U., Kaasik, K., Sage, M., Lu, W., Vaishnav, S., Li, Q.,

Sun, Z.S., Eichele, G., Bradley, A., and Lee, C.C. (2001). Nonredundant

roles of the mPer1 and mPer2 genes in the mammalian circadian clock.

Cell 105, 683–694.

21. Doi, M., Shimatani, H., Atobe, Y., Murai, I., Hayashi, H., Takahashi, Y., Fustin, J.M., Yamaguchi, Y., Kiyonari, H., Koike, N., et al. (2019). Non-coding

cis-element of Period2 is essential for maintaining organismal circadian

behaviour and body temperature rhythmicity. Nat. Commun. 10, 2563.

https://doi.org/10.1038/s41467-019-10532-2.

22. Tainaka, M., Doi, M., Inoue, Y., Murai, I., and Okamura, H. (2018). Circadian PER2 protein oscillations do not persist in cycloheximide-treated

Cell Reports 42, 112157, March 28, 2023 9

ll

OPEN ACCESS

mouse embryonic fibroblasts in culture. Chronobiol. Int. 35, 132–136.

https://doi.org/10.1080/07420528.2017.1316731.

23. Tamaru, T., Hattori, M., Honda, K., Benjamin, I., Ozawa, T., and Takamatsu, K. (2011). Synchronization of circadian Per2 rhythms and HSF1BMAL1:CLOCK interaction in mouse fibroblasts after short-term heat

shock pulse. PLoS One 6, e24521. https://doi.org/10.1371/journal.pone.

0024521.

24. Koyanagi, S., Hamdan, A.M., Horiguchi, M., Kusunose, N., Okamoto, A.,

Matsunaga, N., and Ohdo, S. (2011). cAMP-response element (CRE)mediated transcription by activating transcription factor-4 (ATF4) is essential for circadian expression of the Period2 gene. J. Biol. Chem. 286,

32416–32423. https://doi.org/10.1074/jbc.M111.258970.

25. Gerber, A., Esnault, C., Aubert, G., Treisman, R., Pralong, F., and Schibler,

U. (2013). Blood-borne circadian signal stimulates daily oscillations in

actin dynamics and SRF activity. Cell 152, 492–503. https://doi.org/10.

1016/j.cell.2012.12.027.

26. Tisi, L.C., White, P.J., Squirrell, D.J., Murphy, M.J., Lowe, C.R., and

Murray, J.A.H. (2002). Development of a thermostable firefly luciferase.

Anal. Chim. Acta 457, 115–123. https://doi.org/10.1016/s0003-2670(01)

01496-9.

27. Lin, R.Z., Hu, Z.W., Chin, J.H., and Hoffman, B.B. (1997). Heat shock activates c-Src tyrosine kinases and phosphatidylinositol 3-kinase in NIH3T3

fibroblasts. J. Biol. Chem. 272, 31196–31202. https://doi.org/10.1074/jbc.

272.49.31196.

28. Sato, S., Fujita, N., and Tsuruo, T. (2000). Modulation of Akt kinase activity

by binding to Hsp90. Proc. Natl. Acad. Sci. USA 97, 10832–10837. https://

doi.org/10.1073/pnas.170276797.

29. Yoshihara, T., Naito, H., Kakigi, R., Ichinoseki-Sekine, N., Ogura, Y., Sugiura, T., and Katamoto, S. (2013). Heat stress activates the Akt/mTOR

signalling pathway in rat skeletal muscle. Acta Physiol. 207, 416–426.

https://doi.org/10.1111/apha.12040.

30. Kitani, S., Teshima, R., Morita, Y., Ito, K., Matsuda, Y., and Nonomura, Y.

(1992). Inhibition of IgE-mediated histamine release by myosin light chain

kinase inhibitors. Biochem. Biophys. Res. Commun. 183, 48–54. https://

doi.org/10.1016/0006-291x(92)91607-r.

31. Hausdorff, S.F., Fingar, D.C., Morioka, K., Garza, L.A., Whiteman, E.L.,

Summers, S.A., and Birnbaum, M.J. (1999). Identification of wortmannin-sensitive targets in 3T3-L1 adipocytes. Dissociation of insulinstimulated glucose uptake and glut4 translocation. J. Biol. Chem. 274,

24677–24684. https://doi.org/10.1074/jbc.274.35.24677.

32. Hennessy, B.T., Smith, D.L., Ram, P.T., Lu, Y., and Mills, G.B. (2005).

Exploiting the PI3K/AKT pathway for cancer drug discovery. Nat. Rev.

Drug Discov. 4, 988–1004. https://doi.org/10.1038/nrd1902.

33. Crosby, P., Hamnett, R., Putker, M., Hoyle, N.P., Reed, M., Karam, C.J.,

Maywood, E.S., Stangherlin, A., Chesham, J.E., Hayter, E.A., et al.

(2019). Insulin/IGF-1 drives PERIOD synthesis to entrain circadian rhythms

with feeding time. Cell 177, 896–909.e20. https://doi.org/10.1016/j.cell.

2019.02.017.

34. Zhou, J., Wan, J., Gao, X., Zhang, X., Jaffrey, S.R., and Qian, S.B. (2015).

Dynamic m(6)A mRNA methylation directs translational control of

heat shock response. Nature 526, 591–594. https://doi.org/10.1038/

nature15377.

35. Fustin, J.M., Doi, M., Yamaguchi, Y., Hida, H., Nishimura, S., Yoshida, M.,

Isagawa, T., Morioka, M.S., Kakeya, H., Manabe, I., and Okamura, H.

(2013). RNA-methylation-dependent RNA processing controls the speed

of the circadian clock. Cell 155, 793–806. https://doi.org/10.1016/j.cell.

2013.10.026.

36. Hirota, T., Lee, J.W., Lewis, W.G., Zhang, E.E., Breton, G., Liu, X., Garcia,

M., Peters, E.C., Etchegaray, J.P., Traver, D., et al. (2010). Highthroughput chemical screen identifies a novel potent modulator of cellular

circadian rhythms and reveals CKIalpha as a clock regulatory kinase.

PLoS Biol. 8, e1000559. https://doi.org/10.1371/journal.pbio.1000559.

10 Cell Reports 42, 112157, March 28, 2023

Article

37. Narasimamurthy, R., Hunt, S.R., Lu, Y., Fustin, J.M., Okamura, H., Partch,

C.L., Forger, D.B., Kim, J.K., and Virshup, D.M. (2018). CK1delta/

epsilon protein kinase primes the PER2 circadian phosphoswitch. Proc.

Natl. Acad. Sci. USA 115, 5986–5991. https://doi.org/10.1073/pnas.

1721076115.

38. Goda, T., Doi, M., Umezaki, Y., Murai, I., Shimatani, H., Chu, M.L., Nguyen,

V.H., Okamura, H., and Hamada, F.N. (2018). Calcitonin receptors are

ancient modulators for rhythms of preferential temperature in insects

and body temperature in mammals. Genes Dev. 32, 140–155. https://

doi.org/10.1101/gad.307884.117.

39. Gotic, I., Omidi, S., Fleury-Olela, F., Molina, N., Naef, F., and Schibler, U.

(2016). Temperature regulates splicing efficiency of the cold-inducible

RNA-binding protein gene Cirbp. Genes Dev. 30, 2005–2017. https://

doi.org/10.1101/gad.287094.116.

40. Spo¨rl, F., Schellenberg, K., Blatt, T., Wenck, H., Wittern, K.P., Schrader,

A., and Kramer, A. (2011). A circadian clock in HaCaT keratinocytes.

J. Invest. Dermatol. 131, 338–348. https://doi.org/10.1038/jid.2010.315.

41. Hoyle, N.P., Seinkmane, E., Putker, M., Feeney, K.A., Krogager, T.P., Chesham, J.E., Bray, L.K., Thomas, J.M., Dunn, K., Blaikley, J., and O’Neill,

J.S. (2017). Circadian actin dynamics drive rhythmic fibroblast mobilization during wound healing. Sci. Transl. Med. 9, eaal2774. https://doi.org/

10.1126/scitranslmed.aal2774.

42. Cao, R., Gkogkas, C.G., de Zavalia, N., Blum, I.D., Yanagiya, A., Tsukumo,

Y., Xu, H., Lee, C., Storch, K.F., Liu, A.C., et al. (2015). Light-regulated

translational control of circadian behavior by eIF4E phosphorylation.

Nat. Neurosci. 18, 855–862. https://doi.org/10.1038/nn.4010.

43. Cao, R., Robinson, B., Xu, H., Gkogkas, C., Khoutorsky, A., Alain, T., Yanagiya, A., Nevarko, T., Liu, A.C., Amir, S., and Sonenberg, N. (2013).

Translational control of entrainment and synchrony of the suprachiasmatic

circadian clock by mTOR/4E-BP1 signaling. Neuron 79, 712–724. https://

doi.org/10.1016/j.neuron.2013.06.026.

44. Bohlen, J., Harbrecht, L., Blanco, S., Clemm von Hohenberg, K., Fenzl, K.,

Kramer, G., Bukau, B., and Teleman, A.A. (2020). DENR promotes translation reinitiation via ribosome recycling to drive expression of oncogenes

including ATF4. Nat. Commun. 11, 4676. https://doi.org/10.1038/s41467020-18452-2.

45. Park, H.G., Han, S.I., Oh, S.Y., and Kang, H.S. (2005). Cellular responses

to mild heat stress. Cell. Mol. Life Sci. 62, 10–23. https://doi.org/10.1007/

s00018-004-4208-7.

46. Murata, N., and Los, D.A. (1997). Membrane fluidity and temperature

perception. Plant Physiol. 115, 875–879. https://doi.org/10.1104/pp.115.

3.875.

47. Kowalska, E., Ripperger, J.A., Hoegger, D.C., Bruegger, P., Buch, T.,

Birchler, T., Mueller, A., Albrecht, U., Contaldo, C., and Brown, S.A.

(2013). NONO couples the circadian clock to the cell cycle. Proc.

Natl. Acad. Sci. USA 110, 1592–1599. https://doi.org/10.1073/pnas.

1213317110.

48. Dekoninck, S., and Blanpain, C. (2019). Stem cell dynamics, migration and

plasticity during wound healing. Nat. Cell Biol. 21, 18–24. https://doi.org/

10.1038/s41556-018-0237-6.

49. Nagoshi, E., Saini, C., Bauer, C., Laroche, T., Naef, F., and Schibler, U.

(2004). Circadian gene expression in individual fibroblasts: cell-autonomous and self-sustained oscillators pass time to daughter cells. Cell

119, 693–705. https://doi.org/10.1016/j.cell.2004.11.015.

50. O’Neill, J.S., and Hastings, M.H. (2008). Increased coherence of circadian

rhythms in mature fibroblast cultures. J. Biol. Rhythms 23, 483–488.

https://doi.org/10.1177/0748730408326682.

51. Sun, B.K., Siprashvili, Z., and Khavari, P.A. (2014). Advances in skin grafting and treatment of cutaneous wounds. Science 346, 941–945. https://

doi.org/10.1126/science.1253836.

52. VanInsberghe, M., van den Berg, J., Andersson-Rolf, A., Clevers, H., and

van Oudenaarden, A. (2021). Single-cell Ribo-seq reveals cell

ll

Article

cycle-dependent translational pausing. Nature 597, 561–565. https://doi.

org/10.1038/s41586-021-03887-4.

53. Balsalobre, A., Brown, S.A., Marcacci, L., Tronche, F., Kellendonk, C.,

€tz, G., and Schibler, U. (2000). Resetting of circadian

Reichardt, H.M., Schu

time in peripheral tissues by glucocorticoid signaling. Science 289, 2344–

2347. https://doi.org/10.1126/science.289.5488.2344.

€tz, G., and Schibler, U. (2001).

54. Le Minh, N., Damiola, F., Tronche, F., Schu

Glucocorticoid hormones inhibit food-induced phase-shifting of peripheral circadian oscillators. EMBO J. 20, 7128–7136. https://doi.org/10.

1093/emboj/20.24.7128.

55. Sato, S., Dyar, K.A., Treebak, J.T., Jepsen, S.L., Ehrlich, A.M., Ashcroft,

S.P., Trost, K., Kunzke, T., Prade, V.M., Small, L., et al. (2022). Atlas of exercise metabolism reveals time-dependent signatures of metabolic homeostasis. Cell Metab. 34, 329–345.e8. https://doi.org/10.1016/j.cmet.

2021.12.016.

56. Asher, G., Reinke, H., Altmeyer, M., Gutierrez-Arcelus, M., Hottiger, M.O.,

and Schibler, U. (2010). Poly(ADP-ribose) polymerase 1 participates in the

phase entrainment of circadian clocks to feeding. Cell 142, 943–953.

https://doi.org/10.1016/j.cell.2010.08.016.

57. Acosta-Rodrı´guez, V., Rijo-Ferreira, F., Izumo, M., Xu, P., Wight-Carter,

M., Green, C.B., and Takahashi, J.S. (2022). Circadian alignment of early

onset caloric restriction promotes longevity in male C57BL/6J mice. Science 376, 1192–1202. https://doi.org/10.1126/science.abk0297.

58. Grimm, D., Lee, J.S., Wang, L., Desai, T., Akache, B., Storm, T.A., and Kay,

M.A. (2008). In vitro and in vivo gene therapy vector evolution via multispe-

OPEN ACCESS

cies interbreeding and retargeting of adeno-associated viruses. J. Virol.

82, 5887–5911. https://doi.org/10.1128/JVI.00254-08.

59. Yamaguchi, Y., Murai, I., Goto, K., Doi, S., Zhou, H., Setsu, G., Shimatani,

H., Okamura, H., Miyake, T., and Doi, M. (2021). Gpr19 is a circadian

clock-controlled orphan GPCR with a role in modulating free-running

period and light resetting capacity of the circadian clock. Sci. Rep. 11,

22406. https://doi.org/10.1038/s41598-021-01764-8.

60. Seluanov, A., Vaidya, A., and Gorbunova, V. (2010). Establishing primary

adult fibroblast cultures from rodents. J. Vis. Exp. 44, e2033. https://doi.

org/10.3791/2033.

61. Mito, M., Mishima, Y., and Iwasaki, S. (2020). Protocol for disome profiling

to survey ribosome collision in humans and zebrafish. STAR Protoc. 1,

100168. https://doi.org/10.1016/j.xpro.2020.100168.

62. Sasaki, L., Hamada, Y., Yarimizu, D., Suzuki, T., Nakamura, H., Shimada,

A., Pham, K.T.N., Shao, X., Yamamura, K., Inatomi, T., et al. (2022). Intracrine activity involving NAD-dependent circadian steroidogenic activity

governs age-associated meibomian gland dysfunction. Nat. Aging 2,

105–114.

63. Morioka, S., Nakanishi, H., Yamamoto, T., Hasegawa, J., Tokuda, E., Hikita, T., Sakihara, T., Kugii, Y., Oneyama, C., Yamazaki, M., et al. (2022).

A mass spectrometric method for in-depth profiling of phosphoinositide

regioisomers and their disease-associated regulation. Nat. Commun. 13,

83. https://doi.org/10.1038/s41467-021-27648-z.

Cell Reports 42, 112157, March 28, 2023 11

ll

OPEN ACCESS

Article

STAR+METHODS

KEY RESOURCES TABLE

REAGENT or RESOURCE

SOURCE

IDENTIFIER

anti-Per2

Doi et al., 201921

N/A

anti-Cry2

MBL

PM082

anti-Clock

MBL

D349-3

anti-Bmal1

MBL

D335-3

anti-HSF1

Abcam

Cat# Ab61382; RRID: AB_942016

Antibodies

anti-p-CREB

Cell Signaling

Cat# 9198; RRID: AB_2561044

anti-FosB

Cell Signaling

Cat# 2251; RRID: AB_2106903

anti-tubulin

Sigma

Cat# T6199; RRID: AB_477583

anti-b-actin

Sigma

Cat# A5441; RRID: AB_476744

AAV-DJ

Grimm et al., 200858

N/A

DH5a Escherichia coli

Takara Bio

#9057

Bacterial and virus strains

Chemicals, peptides, and recombinant proteins

17b-hydroxy-wortmannin

Cayman

#13812

PI3-Kinase a Inhibitor 2

Cayman

#21197

GSK1059615

Cayman

#11569

EGF

Peprotech

14321-54

Dexamethasone

Sigma

D8893

Cycloheximide

Nacalai Tesque

06741-91

Biotin-PEG4-alkyne

Sigma

#764213

KNK437

Sigma

SML0964

Rapamycin

Cayman

#13346

MK2206

Selleck chemicals

S1078

U0126

Cayman

#70970

DZnepA

Sigma

SML0305

Ruthenium Red

Wako

189-03181

Thermo Fisher

C10276

Ribo-seq and RNA-seq data (for Figures 1, 3, and S1)

This paper

GEO: GSE188529

Ribo-seq data (for Figures 3 and S1)

This paper

GEO: GSE211532

Unprocessed original immunoblot data

(for Figures 2, 3, 4, and S2–S6)

This paper

Mendeley Data: https://doi.org/10.17632/k4wygddpkd.1

Critical commercial assays

Click-iT Protein Reaction Buffer Kit

Deposited data

Experimental models: Cell lines

Per2::LucTS knock-in MEFs

This paper

N/A

Tet-On 3G NIH3T3

Clontech

#631197; RRID: CVCL_V360

This paper

RIKEN BRC No. 11461

Quantitative RT-PCR forward primer for Per2:

50 -CTC ACT GAG ATT CGG GAT ATG-30

Doi et al., 201921

N/A

Quantitative RT-PCR reverse primer for Per2:

50 -CTC CCA CCT TGT CTC CAG TC-30

Doi et al., 201921

N/A

Quantitative RT-PCR forward primer for Cirp:

50 -GCA GAT CTC CGA AGT GGT G-30

Morf et al., 20128

N/A

Experimental models: Organisms/strains

Per2 minimal uORF mutant mice

Oligonucleotides

(Continued on next page)

12 Cell Reports 42, 112157, March 28, 2023

ll

OPEN ACCESS

Article

Continued

REAGENT or RESOURCE

SOURCE

IDENTIFIER

Quantitative RT-PCR reverse primer for Cirp:

50 -CAG CCT GGT CAA CTC TGA T-30

Morf et al., 2012

N/A

Quantitative RT-PCR forward primer for Rplp0:

50 -CTC ACT GAG ATT CGG GAT ATG-30

Doi et al., 201921

N/A

Quantitative RT-PCR reverse primer for Rplp0:

50 -CTC CCA CCT TGT CTC CAG TC-30

Doi et al., 201921

N/A

Probes and primers for Taqman qPCR, see Table S2

Doi et al., 2019,21

Yamaguchi et al., 202159

N/A

shRNA targeting sequence for control:

50 -ATA ACA TGG CCA TCA TCA AGG AGT TCA TG-30

This paper

N/A

shRNA targeting sequence for Cirp #1:

50 -CAG AGA CAG CTA TGA CAG TTA-30

Thermo Fisher

https://www.thermofisher.com/jp/ja/home/

life-science/rnai/synthetic-rnai-analysis.html

shRNA targeting sequence for Cirp #2:

50 -CGT CCT TCC ATG GCT GTA ATT-30

Thermo Fisher

https://www.thermofisher.com/jp/ja/home/

life-science/rnai/synthetic-rnai-analysis.html

Recombinant DNA

pHSE-Luc

Clontech

#K2049-1

pCRE-Luc

Clontech

#K2049-1

pSRE-Luc

Clontech

#K2049-1

pTRE3G-Luc

Clontech

#631168

STAR v2.7.0

N/A

https://github.com/alexdobin/STAR

Samtools v1.10

N/A

https://github.com/samptools/samtools

fp-framing, fp-transcript

N/A

https://github.com/ingolia-lab/RiboSeq

ImageJ

NIH

https://imagej.nih.gov/ij/

Prism 8

GraphPad

N/A

Software and algorithms

RESOURCE AVAILABILITY

Lead contact

Further information and requests for resources and reagents should be directed to and will be fulfilled by the lead contact, Masao Doi

(doimasao@pharm.kyoto-u.ac.jp).

Materials availability

Per2 m-uORF mutant mouse line has been deposited to RIKEN BioResource Research Center (RIKEN BRC No. 11461, https://

knowledge.brc.riken.jp/resource/animal/). All other reagents generated in this study are available from the lead contact with a

completed materials transfer agreement.

Data and code availability

d Ribo-seq and RNA-seq data have been deposited at GEO and are publicly available as of the date of publication. Accession

numbers are GSE188529 and GSE211532. Unprocessed original immunoblot data have been deposited to Mendely Data are

available at: https://doi.org/10.17632/k4wygddpkd.1. All other data reported in this paper will be shared by the lead contact

upon request.

d This paper does not report original code.

d Any additional information required to reanalyze the data reported in this paper is available from the lead contact upon request.

EXPERIMENTAL MODEL AND SUBJECT DETAILS

Cell lines

The Tet-On 3G NIH3T3 cell line was purchased from Clontech (Mountain View, CA) and maintained in Dulbecco’s modified Eagle’s

medium (DMEM) supplemented with 10% fetal bovine serum (FBS), 100 mg/mL G418 at 37 C in a humidified 5% CO2 atmosphere. To

generate Per2::LucTS knockin MEF cells, we used a firefly luciferase containing the mutations T214A, I232A, F295L, and E354K

(LucTS).26 The LucTS coding sequence was inserted in frame before the endogenous stop codon of Per2 using CRISPR/Cas9

Cell Reports 42, 112157, March 28, 2023 13

ll

OPEN ACCESS

Article

technology. The gRNA sequence used was 50 -TGA GGT ATC ACA GAT TCC CG-30 . The edited genome sequence was verified by

PCR-Sanger sequencing. Cells were maintained in DMEM/Ham’s F-12 (1:1) supplemented with 10% FBS.

Mice

The C57BL/6J male mice 6–8 weeks old were purchased from Japan SLC (Shizuoka, Japan). Per2 m-uORF mutant mice were established at the Animal Resource Center for Infectious Diseases, Research Institute for Microbial Diseases, Osaka University, by

CRISPR/Cas9-based genome editing using C57BL/6J mouse zygotes. The sequences of gRNA and the template oligonucleotide

were 50 -TTT CCA CTA TGT GAC AGC GGA GG-30 and 50 -CAA TGG CGC GCG CAG GGG CGG GCT CAG CGC GCG CGG TCA

CGT TTT CCA CTT AAC AGC GGA GGG CGA CGC GGC GGC AGC GGC GCT ACT GGG ACT AGC GGC TCC G-30 , respectively.

The edited sequence was verified by Sanger DNA sequencing. All procedures for animal experiments in this study were conducted in

compliance with the Ethical Regulations of Kyoto University and performed under protocols approved by the Animal Care and Experimentation Committee of Kyoto University and the Institutional Animal Care and Use Committees of Osaka university.

Primary fibroblasts

Primary fibroblasts were isolated from the lung or skin of adult Per2 m-uORF mutant mice and control wild-type siblings, according

to a protocol previously established.60 Briefly, the tissue fragments were cut into 1 mm pieces and incubated in Blendzyme 3

(0.14 Wunsch units/mL, Roche, 11814176001)-containing DMEM/F12 medium at 37 C for 90 min. After centrifugation for

5 min 3 524 g, the pellet was resuspended in DMEM/F12 media supplemented with 15% FBS, and then incubated at 37 C, 5%

CO2. Cells used for experiments were between passages 2 and 4.

METHOD DETAILS

Temperature control

WTS was applied to cells by transferring the cell dish from a 35 C incubator to a 38.5 C-prewarmed incubator. The temperature of the

medium was measured using an electronic thermometer and verified to approach 38.5  C at 7–8 min after being transferred. Application of simulated body temperature was performed using a custom-modified convection-ventilated incubator box, in which the

temperature was controlled by a Peltier device and forced air circulation. This device enabled the programming and recording of temperature profiles in real time with an accuracy of 0.1 C.

Ribo-seq and RNA-seq analysis

We used Dex-synchronized MEF cells treated with or without WTS for 2 h. Cells were lysed with a buffer containing 20 mM Tris-HCl

[pH 7.5], 150 mM NaCl, 5 mM MgCl2, 1 mM DTT, 1% Triton X-100, and 100 mg/mL cycloheximide (CHX). Following the digestion of

genomic DNA with DNase I (25 U/mL, Thermo Fisher Scientific), lysates were centrifuged at 20,000 3 g and the resultant supernatant was processed for library generation.61 For Ribo-seq, samples (10 mg RNA each) were digested with 20 U RNase (Epicentre)

at 25 C for 45 min, and the RNA fragments protected by ribosomes ranging 17–34 nt were gel-excised and ligated with linker oligonucleotides followed by rRNA depletion using Ribo-Zero Gold rRNA Removal Kit (Illumina). The resultant RNA fragments were

reverse transcribed with Photoscript II (NEB) and circularized using CircLigaseII (Epicentre). The circularized cDNA templates

were PCR amplified for 6 to 8 cycles using Phusion polymerase (NEB) and sequenced on an Illumina Hiseq X or HiSeq 4000 instrument. For RNA-seq, samples were subjected to library construction using TruSeq Stranded Total RNA Gold (Illumina) and sequenced

on HiSeq 4000 as described62; single-end reads (150 bp for experiment 1 and 50 bp for the others) were mapped to the mouse

genome (GRCm38/mm10) using STAR (version 2.7.0) and sorted and indexed using samtools (version 1.10). For filtering rRNA

and tRNA, STAR was used with rRNA and tRNA annotations downloaded from the UCSC table browser. Footprints ranging from

20–34 nt were used for further analysis. In Figure 3G, the minimal uORF was identified as NUG-stop (N = A, C, U or G). The A-site

location of the Ribo-seq footprints and 3-nucleotide periodicity, which is supplied to validate our Ribo-seq data quality (Table S1),

were calculated as described61 using the fp-transcript and fp-framing, respectively (https://github.com/ingolia-lab/RiboSeq).

Polysome profiling

Linear 15%–50% sucrose gradients were prepared using a BioComp Gradient Master (Biocomp Instruments). Sucrose was dissolved in 20 mM Tris-HCl [pH 7.5], 10 mM MgCl2, 100 mM KCl, 2 mM DTT, 100 U/mL human recombinant RNase Inhibitor

(TOYOBO), and 100 mg/mL CHX. Cells were lysed in ice-cold buffer containing 20 mM Tris-HCl [pH 7.5], 150 mM NaCl, 5 mM

MgCl2, 1 mM DTT, 1% Triton X-100, and 100 mg/mL CHX. Following the digestion of genomic DNA with DNase I (25 U/mL, Thermo

Fisher Scientific), lysates were centrifuged at 20,000 3 g. The supernatants were loaded on top of the gradients and sedimented at

35,000 rpm in an SW41 rotor at 4 C for 2 h. After centrifugation, gradients were fractionated using the Gradient Master instrument

with continuous monitoring at A260. RNA in each fraction was extracted using TRIzol LS reagent (Thermo Fisher Scientific).

Immunoblotting

To minimize proteolysis of the endogenous Per2 protein, cells and tissues (dorsal ear skin and lung) were lysed immediately in

Laemmli buffer containing 1x cOmplete Protease Inhibitor cocktail (Roche Diagnostics), as described previously.21 Immunoblots

14 Cell Reports 42, 112157, March 28, 2023

ll

OPEN ACCESS

Article

were performed using our standard method with affinity-purified anti-mPer2 rabbit polyclonal antibody (final concentration,

2 mg/mL)21 or commercially available antibodies against a-Tubulin (Sigma Aldrich, T6199, 1:1,000), b-actin (Sigma Aldrich, A5441,

1:1,000), Clock (MBL, D349-3, 1:1,200), Bmal1 (MBL, D335-3, 1:200), Cry2 (MBL, PM082, 1:200), HSF1 (abcam, ab61382, 1:200),

p-CREB (Cell Signaling, #9198, 1:1,000), and FosB (Cell Signaling, #2251, 1:1,000). The CHX chase assay was performed by adding

CHX (100 mg/mL) to the medium at Time 24. Where specified, cells were treated with 17b-hydroxy Wortmannin (Cayman chemical,

10 mM), PI3-Kinase a Inhibitor 2 (Cayman chemical, 10 mM), GSK1059615 (Cayman chemical 10 mM), or EGF (Peprotech, 10 ng/mL).

Labeling of de novo protein synthesis

Dex-synchronized cells were preincubated with a Met/Cys-free DMEM medium (Thermo Fisher Scientific) for 3 h before experiments.

[35S]Methionine (14.8 MBq/sample, Muromachi Kikai) was added to the medium at Time 0 and incubated for 2 h with or without WTS.

Cells were lysed in RIPA buffer (50 mM Tris-HCl [pH 8.0], 150 mM NaCl, 0.1% SDS, 1% Nonidet P40, 1% sodium deoxycholate)

containing 1x cOmplete Protease Inhibitor cocktail (Roche Diagnostics) and subjected to immunoprecipitation with anti-mPer2 polyclonal antibody21 or anti-Cry2 antibody (MBL, PM082). The immunoprecipitate was separated by SDS-PAGE and visualized with an

X-ray film (Kodak). For click-labeling experiments, azidohomoalanine (AHA) was added to the medium at Time 0 (final concentration,

50 mM) instead of [35S]Methionine. Cells were lysed in RIPA buffer containing 1x cOmplete Protease Inhibitor cocktail. The lysates

were conjugated with biotin-PEG4-alkyne (Sigma Aldrich) using the Click-iT Protein Reaction Buffer Kit (Thermo Fisher Scientific).

After removal of excess free biotin-PEG4-alkyne using PD SpinTrap G-25 columns (Cytiva), AHA-biotin-alkyne labeled proteins

were affinity-purified with streptavidin beads (Thermo Fisher Scientific). The precipitates were analyzed by immunoblotting using

the anti-mPer2 polyclonal antibody,21 anti-Cry2 (MBL, PM082), or anti-b-actin antibodies (Sigma Aldrich, A5441). Biotin-labeled proteins were detected using Vectastain Elite ABC Standard Kit (Vector Laboratories).

LucTS reporter assay

pHSE-LucTS, pCRE-LucTS, and pSRE-LucTS were generated by replacing the Luc coding sequence of pHSE-Luc, pCRE-Luc, and

pSRE-Luc (all vectors from Clontech) with the LucTS. One day after transfection, MEFs were either treated with WTS, heat shock

(42 C), 100 mM forskolin, or 50% bovine serum. The mean values of the 4 h pre-stimulation baseline of lumi-nescence of pHSE-,

pCRE-, and pSRE-LucTS were set to 1. The Tet-inducible expression vector pTRE3G (Clontech) was used to construct the Per2

50 UTR-LucTS vectors. The 50 UTR region of the pTRE3G vector was substituted with the full-length Per2 50 UTR sequence of mouse

(NM_011066) or human (XM_006712824) origin. Mutagenesis of the Per2 m-uORF was performed using a standard sequential PCR

method.21 Tet-On 3G NIH3T3cells were transfected with a plasmid using the Viofectin reagent (Viogene). To reduce technical variation between transfections, the transfection was performed in a single dish and the cells were split into test dishes. D-luciferin

(Promega) was added to the medium before experiments.21 Cells were treated with doxycycline (1 mg/mL). Luminescence was

continuously monitored using a custom-modified dish-type luminometer AB-2550 Kronos Dio (ATTO), which enables the programming and recording of temperature profiles in real-time with an accuracy of 0.1 C. Luminescence was measured for 2 min at 30-min

intervals. Values were normalized to the luminescence intensity that was recorded 30 min before WTS. The fold increase was calculated by dividing the intensity of WTS-treated cells by that of non-treated controls.

Per2::LucTS cell-based kinase inhibitor assay

Per2::LucTS cells were treated with a compound in the Kinase Screening Library (Cayman Chemical, #10505) at a concentration of

10 mM/0.1% DMSO immediately before WTS. The score was calculated using the following formula;

Score =  log2

Td of tested compound

Td of vehicle

where Td represents the difference in the half-maximum time between WTS and no-WTS cells. The results of compounds that profoundly affect basal LucTS activity were omitted. To evaluate dose dependency, Wortm was administrated at a concentration of 0.3,

1, 10, 30, 100, and 1000 nM. Where indicated, KNK437 (Sigma Aldrich, 100 mM), Rapamycin (Cayman chemical, 20 nM), MK2206

(Selleck chemicals, 10 mM), U0126 (Cayman chemical, 10 mM), DZnepA (Sigma Aldrich, 100 mM), D4476 (Cayman chemical,

5 mM), and Ruthenium Red (Wako, 10 mM) were used.

Quantification of PIP3 species

Cells were incubated in a serum-free medium 1 day before experiments and sampled afte ...

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