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 「Uniform transgene activation in Tet-On systems depends on sustained rtTA expression」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

Uniform transgene activation in Tet-On systems depends on sustained rtTA expression

Otomo, Jun Woltjen, Knut Sakurai, Hidetoshi 京都大学 DOI:10.1016/j.isci.2023.107685

2023.10.20

概要

Application of the tetracycline-inducible gene expression system (Tet-On) in human induced pluripotent stem cells (hiPSCs) has become a fundamental transgenic tool owing to its regulatable gene expression. One of the major hurdles in hiPSC application is non-uniform transgene activation. Here, we report that the supplementation of reverse tetracycline transactivator (rtTA) in polyclonal hiPSCs populations can achieve the uniform transgene activation of Tet-On. Furthermore, the choice of antibiotic selection markers connected by an internal ribosomal entry site (IRES) can influence the expression of upstream transgenes. In particular, expression of the rtTA is more uniform in cell populations when linked to puromycin as compared to neomycin, obviating the need for sub-cloning or supplementation of rtTA. Finally, to expand the range of applications, we adopted our findings to tetracycline-inducible MyoD vector (Tet-MyoD). Our Tet-MyoD promises efficient, robust, and reproducible directed myogenic differentiation of hiPSCs.

この論文で使われている画像

参考文献

1. Baron, U., and Bujard, H. (2000). Tet

repressor-based system for regulated gene

expression in eukaryotic cells: principles and

advances. Methods Enzymol. 327, 401–421.

https://doi.org/10.1016/s0076-6879(00)

27292-3.

2. Gossen, M., Freundlieb, S., Bender, G.,

Mu¨ller, G., Hillen, W., and Bujard, H. (1995).

Transcriptional activation by tetracyclines in

mammalian cells. Science 268, 1766–1769.

https://doi.org/10.1126/science.7792603.

3. Takahashi, K., Tanabe, K., Ohnuki, M., Narita,

M., Ichisaka, T., Tomoda, K., and Yamanaka,

S. (2007). Induction of pluripotent stem cells

from adult human fibroblasts by defined

factors. Cell 131, 861–872. https://doi.org/10.

1016/j.cell.2007.11.019.

4. Grandy, R., Tomaz, R.A., and Vallier, L. (2019).

Modeling Disease with Human Inducible

Pluripotent Stem Cells. Annu. Rev. Pathol. 14,

449–468. https://doi.org/10.1146/annurevpathol-020117-043634.

5. Tanaka, A., Woltjen, K., Miyake, K., Hotta, A.,

Ikeya, M., Yamamoto, T., Nishino, T., Shoji, E.,

Sehara-Fujisawa, A., Manabe, Y., et al. (2013).

Efficient and reproducible myogenic

differentiation from human iPS cells:

prospects for modeling Miyoshi Myopathy

in vitro. PLoS One 8, e61540. https://doi.org/

10.1371/journal.pone.0061540.

6. Pawlowski, M., Ortmann, D., Bertero, A.,

Tavares, J.M., Pedersen, R.A., Vallier, L., and

Kotter, M.R.N. (2017). Inducible and

Deterministic Forward Programming of

Human Pluripotent Stem Cells into Neurons,

Skeletal Myocytes, and Oligodendrocytes.

Stem Cell Rep. 8, 803–812. https://doi.org/

10.1016/j.stemcr.2017.02.016.

7. Sadahiro, T., Isomi, M., Muraoka, N., Kojima,

H., Haginiwa, S., Kurotsu, S., Tamura, F., Tani,

H., Tohyama, S., Fujita, J., et al. (2018). Tbx6

Induces Nascent Mesoderm from Pluripotent

Stem Cells and Temporally Controls Cardiac

versus Somite Lineage Diversification. Cell

Stem Cell 23, 382–395.e5. https://doi.org/10.

1016/j.stem.2018.07.001.

8. Ding, S., Wu, X., Li, G., Han, M., Zhuang, Y.,

and Xu, T. (2005). Efficient transposition of the

piggyBac (PB) transposon in mammalian cells

and mice. Cell 122, 473–483. https://doi.org/

10.1016/j.cell.2005.07.013.

9. Wilson, M.H., Coates, C.J., and George, A.L.

(2007). PiggyBac transposon-mediated gene

transfer in human cells. Mol. Ther. 15,

139–145. https://doi.org/10.1038/sj.mt.

6300028.

10. Li, M.A., Turner, D.J., Ning, Z., Yusa, K., Liang,

Q., Eckert, S., Rad, L., Fitzgerald, T.W., Craig,

N.L., and Bradley, A. (2011). Mobilization of

giant piggyBac transposons in the mouse

11.

12.

13.

14.

15.

genome. Nucleic Acids Res. 39, e148. https://

doi.org/10.1093/nar/gkr764.

Kim, S.I., Oceguera-Yanez, F., Sakurai, C.,

Nakagawa, M., Yamanaka, S., and Woltjen, K.

(2016). Inducible Transgene Expression in

Human iPS Cells Using Versatile All-in-One

piggyBac Transposons. Methods Mol. Biol.

1357, 111–131. https://doi.org/10.1007/

7651_2015_251.

Uchimura, T., Otomo, J., Sato, M., and

Sakurai, H. (2017). A human iPS cell myogenic

differentiation system permitting highthroughput drug screening. Stem Cell Res.

25, 98–106. https://doi.org/10.1016/j.scr.

2017.10.023.

Davis, R.L., Weintraub, H., and Lassar, A.B.

(1987). Expression of a single transfected

cDNA converts fibroblasts to myoblasts. Cell

51, 987–1000. https://doi.org/10.1016/00928674(87)90585-x.

Shoji, E., Sakurai, H., Nishino, T., Nakahata,

T., Heike, T., Awaya, T., Fujii, N., Manabe, Y.,

Matsuo, M., and Sehara-Fujisawa, A. (2015).

Early pathogenesis of Duchenne muscular

dystrophy modelled in patient-derived

human induced pluripotent stem cells. Sci.

Rep. 5, 12831. https://doi.org/10.1038/

srep12831.

Ueki, J., Nakamori, M., Nakamura, M.,

Nishikawa, M., Yoshida, Y., Tanaka, A.,

iScience 26, 107685, October 20, 2023

13

iScience

ll

Article

OPEN ACCESS

16.

17.

18.

19.

20.

21.

22.

23.

24.

25.

14

Morizane, A., Kamon, M., Araki, T., Takahashi,

M.P., et al. (2017). Myotonic dystrophy type 1

patient-derived iPSCs for the investigation of

CTG repeat instability. Sci. Rep. 7, 42522.

https://doi.org/10.1038/srep42522.

Yoshida, T., Awaya, T., Jonouchi, T., Kimura,

R., Kimura, S., Era, T., Heike, T., and Sakurai,

H. (2017). A Skeletal Muscle Model of

Infantile-onset Pompe Disease with Patientspecific iPS Cells. Sci. Rep. 7, 13473. https://

doi.org/10.1038/s41598-017-14063-y.

Sasaki-Honda, M., Jonouchi, T., Arai, M.,

Hotta, A., Mitsuhashi, S., Nishino, I., Matsuda,

R., and Sakurai, H. (2018). A patient-derived

iPSC model revealed oxidative stress

increases facioscapulohumeral muscular

dystrophy-causative DUX4. Hum. Mol. Genet.

27, 4024–4035. https://doi.org/10.1093/hmg/

ddy293.

Kokubu, Y., Nagino, T., Sasa, K., Oikawa, T.,

Miyake, K., Kume, A., Fukuda, M., Fuse, H.,

Tozawa, R., and Sakurai, H. (2019). Phenotypic

Drug Screening for Dysferlinopathy Using

Patient-Derived Induced Pluripotent Stem

Cells. Stem Cells Transl. Med. 8, 1017–1029.

https://doi.org/10.1002/sctm.18-0280.

Duran, A.G., Schwestka, M., Nazari-Shafti,

T.Z., Neuber, S., Stamm, C., and Gossen, M.

(2022). Limiting Transactivator Amounts

Contribute to Transgene Mosaicism in TetOn All-in-One Systems. ACS Synth. Biol. 11,

2623–2635. https://doi.org/10.1021/

acssynbio.2c00036.

Nakatake, Y., Fujii, S., Masui, S., Sugimoto, T.,

Torikai-Nishikawa, S., Adachi, K., and Niwa,

H. (2013). Kinetics of drug selection systems in

mouse embryonic stem cells. BMC

Biotechnol. 13, 64. https://doi.org/10.1186/

1472-6750-13-64.

Ben-Dor, I., Itsykson, P., Goldenberg, D.,

Galun, E., and Reubinoff, B.E. (2006).

Lentiviral vectors harboring a dual-gene

system allow high and homogeneous

transgene expression in selected polyclonal

human embryonic stem cells. Mol. Ther. 14,

255–267. https://doi.org/10.1016/j.ymthe.

2006.02.010.

Bencsik, R., Boto, P., Szabo´, R.N., Toth, B.M.,

Simo, E., Ba´lint, B.L., and Szatmari, I. (2016).

Improved transgene expression in

doxycycline-inducible embryonic stem cells

by repeated chemical selection or cell

sorting. Stem Cell Res. 17, 228–234. https://

doi.org/10.1016/j.scr.2016.08.014.

Alexopoulou, A.N., Couchman, J.R., and

Whiteford, J.R. (2008). The CMV early

enhancer/chicken b actin (CAG) promoter

can be used to drive transgene expression

during the differentiation of murine

embryonic stem cells into vascular

progenitors. BMC Cell Biol. 9, 2. https://doi.

org/10.1186/1471-2121-9-2.

Niwa, H., Yamamura, K., and Miyazaki, J.

(1991). Efficient selection for high-expression

transfectants with a novel eukaryotic vector.

Gene 108, 193–199. https://doi.org/10.1016/

0378-1119(91)90434-d.

Gallagher, D., Norman, A.A., Woodard, C.L.,

Yang, G., Gauthier-Fisher, A., Fujitani, M.,

Vessey, J.P., Cancino, G.I., Sachewsky, N.,

iScience 26, 107685, October 20, 2023

26.

27.

28.

29.

30.

31.

32.

33.

34.

35.

Woltjen, K., et al. (2013). Transient maternal

IL-6 mediates long-lasting changes in neural

stem cell pools by deregulating an

endogenous self-renewal pathway. Cell Stem

Cell 13, 564–576. https://doi.org/10.1016/j.

stem.2013.10.002.

Borman, A.M., Bailly, J.L., Girard, M., and

Kean, K.M. (1995). Picornavirus internal

ribosome entry segments: comparison of

translation efficiency and the requirements

for optimal internal initiation of translation

in vitro. Nucleic Acids Res. 23, 3656–3663.

https://doi.org/10.1093/nar/23.18.3656.

Hennecke, M., Kwissa, M., Metzger, K.,

Oumard, A., Kro¨ger, A., Schirmbeck, R.,

Reimann, J., and Hauser, H. (2001).

Composition and arrangement of genes

define the strength of IRES-driven translation

in bicistronic mRNAs. Nucleic Acids Res. 29,

3327–3334. https://doi.org/10.1093/nar/29.

16.3327.

Bochkov, Y.A., and Palmenberg, A.C. (2006).

Translational efficiency of EMCV IRES in

bicistronic vectors is dependent upon IRES

sequence and gene location. Biotechniques

41, 283–284. 286, 288 passim. https://doi.org/

10.2144/000112243.

Bugaud, O., Barbier, N., Chommy, H.,

Fiszman, N., Le Gall, A., Dulin, D., Saguy, M.,

Westbrook, N., Perronet, K., and Namy, O.

(2017). Kinetics of CrPV and HCV IRESmediated eukaryotic translation using singlemolecule fluorescence microscopy. RNA 23,

1626–1635. https://doi.org/10.1261/rna.

061523.117.

Johnson, A.G., Grosely, R., Petrov, A.N., and

Puglisi, J.D. (2017). Dynamics of IRESmediated translation. Philos. Trans. R. Soc.

Lond. B Biol. Sci. 372, 20160177. https://doi.

org/10.1098/rstb.2016.0177.

Belsham, G.J., and Brangwyn, J.K. (1990). A

region of the 5’ noncoding region of footand-mouth disease virus RNA directs efficient

internal initiation of protein synthesis within

cells: involvement with the role of L protease

in translational control. J. Virol. 64, 5389–

5395. https://doi.org/10.1128/JVI.64.11.53895395.1990.

Gurtu, V., Yan, G., and Zhang, G. (1996). IRES

bicistronic expression vectors for efficient

creation of stable mammalian cell lines.

Biochem. Biophys. Res. Commun. 229,

295–298. https://doi.org/10.1006/bbrc.

1996.1795.

Mizuguchi, H., Xu, Z., Ishii-Watabe, A.,

Uchida, E., and Hayakawa, T. (2000). IRESdependent second gene expression is

significantly lower than cap-dependent first

gene expression in a bicistronic vector. Mol.

Ther. 1, 376–382. https://doi.org/10.1006/

mthe.2000.0050.

Mansha, M., Wasim, M., Ploner, C., Hussain,

A., Latif, A.A., Tariq, M., and Kofler, A. (2012).

Problems encountered in bicistronic IRESGFP expression vectors employed in

functional analyses of GC-induced genes.

Mol. Biol. Rep. 39, 10227–10234. https://doi.

org/10.1007/s11033-012-1898-z.

Stadtfeld, M., Maherali, N., Borkent, M., and

Hochedlinger, K. (2010). A reprogrammable

36.

37.

38.

39.

40.

41.

42.

43.

44.

mouse strain from gene-targeted embryonic

stem cells. Nat. Methods 7, 53–55. https://

doi.org/10.1038/nmeth.1409.

Ordova´s, L., Boon, R., Pistoni, M., Chen, Y.,

Wolfs, E., Guo, W., Sambathkumar, R., BobisWozowicz, S., Helsen, N., Vanhove, J., et al.

(2015). Efficient Recombinase-Mediated

Cassette Exchange in hPSCs to Study the

Hepatocyte Lineage Reveals AAVS1 LocusMediated Transgene Inhibition. Stem Cell

Rep. 5, 918–931. https://doi.org/10.1016/j.

stemcr.2015.09.004.

Bhagwan, J.R., Collins, E., Mosqueira, D.,

Bakar, M., Johnson, B.B., Thompson, A.,

Smith, J.G.W., and Denning, C. (2019).

Variable expression and silencing of CRISPRCas9 targeted transgenes identifies the.

F1000Res. 8, 1911. https://doi.org/10.12688/

f1000research.19894.2.

Wu, F., Zhang, Q., and Wang, X. (2018).

Design of Adjacent Transcriptional Regions

to Tune Gene Expression and Facilitate

Circuit Construction. Cell Syst. 6, 206–215.e6.

https://doi.org/10.1016/j.cels.2018.01.010.

Takahashi, Y., Wu, J., Suzuki, K., MartinezRedondo, P., Li, M., Liao, H.K., Wu, M.Z.,

Herna´ndez-Benı´tez, R., Hishida, T., Shokhirev,

M.N., et al. (2017). Integration of CpG-free

DNA induces de novo methylation of CpG

islands in pluripotent stem cells. Science 356,

503–508. https://doi.org/10.1126/science.

aag3260.

Kabadi, A.M., Thakore, P.I., Vockley, C.M.,

Ousterout, D.G., Gibson, T.M., Guilak, F.,

Reddy, T.E., and Gersbach, C.A. (2015).

Enhanced MyoD-induced transdifferentiation

to a myogenic lineage by fusion to a potent

transactivation domain. ACS Synth. Biol. 4,

689–699. https://doi.org/10.1021/sb500322u.

Go¨decke, N., Zha, L., Spencer, S., Behme, S.,

Riemer, P., Rehli, M., Hauser, H., and Wirth, D.

(2017). Controlled re-activation of

epigenetically silenced Tet promoter-driven

transgene expression by targeted

demethylation. Nucleic Acids Res. 45, e147.

https://doi.org/10.1093/nar/gkx601.

Cullmann, K., Blokland, K.E.C., Sebe, A.,

Schenk, F., Ivics, Z., Heinz, N., and Modlich, U.

(2019). Sustained and regulated gene

expression by Tet-inducible "all-in-one"

retroviral vectors containing the HNRPA2B1CBX3 UCOE. Biomaterials 192, 486–499.

https://doi.org/10.1016/j.biomaterials.2018.

11.006.

Collinson, A., Collier, A.J., Morgan, N.P.,

Sienerth, A.R., Chandra, T., Andrews, S., and

Rugg-Gunn, P.J. (2016). Deletion of the

Polycomb-Group Protein EZH2 Leads to

Compromised Self-Renewal and

Differentiation Defects in Human Embryonic

Stem Cells. Cell Rep. 17, 2700–2714. https://

doi.org/10.1016/j.celrep.2016.11.032.

Kwan, K.M., Fujimoto, E., Grabher, C.,

Mangum, B.D., Hardy, M.E., Campbell, D.S.,

Parant, J.M., Yost, H.J., Kanki, J.P., and Chien,

C.B. (2007). The Tol2kit: A multisite gatewaybased construction kit for Tol2 transposon

transgenesis constructs. Dev. Dan. 236, 3088–

3099. https://doi.org/10.1002/dvdy.21343.

iScience

ll

Article

OPEN ACCESS

STAR+METHODS

KEY RESOURCES TABLE

REAGENT or RESOURCE

SOURCE

IDENTIFIER

Sasaki-Honda et al.17

Graduate School and Faculty

Experimental models: Cell lines

HC #1 hiPS cell lines

of Medicine of Kyoto University

(approval numbers #R0091)

Oligonucleotides

MscI-EagI adapter oligos (Fwd: CCACAACCATGA

This paper

N/A

This paper

N/A

PB-TA-ERN

Laboratory of Knut Woltjen (Kim et al.)11

Addgene Plasmid #80474, KW110

PB-TAC-ERN

Laboratory of Knut Woltjen (Kim et al.)11

Addgene Plasmid #80475, KW111

PB-TA-ERP2

Laboratory of Knut Woltjen (Kim et al.)

11

Addgene Plasmid #80477, KW542

PB-TAC-ERP2

Laboratory of Knut Woltjen (Kim et al.)11

Addgene Plasmid #80478, KW543

PB-TA-ERN2

This paper

KW1406

PB-TAC-ERN2

This paper

TTGAACAAGATGGATTGCACGCAGGTTCTCC)

MscI-EagI adapter oligos (Rev: GGCCGGAGAACCTG

CGTGCAATCCATCTTGTTCAATCATGGTTGTGG)

Recombinant DNA

KW1413

43

PB-CAG-rtTA-ires-Puro

Collinson et al.

pENTR-MyoD

Tanaka et al.5

EMCV-Neo-MYOD1 (PB-TA-ERN2-MyoD)

This paper

KW1410

EMCV-Neo-MYOD1-mCherry

This paper

KW1415

N/A

N/A

(PB-TAC-ERN2-MyoD)

EMCV-Puro-MYOD1-mCherry (PB-TAC-ERP2)

This paper

KW1409

FMDV-Neo-MYOD1-mCherry (PB-TAC-ERN-MyoD)

Tanaka et al.5

KW698

EMCV-Puro-MYOD1 (PB-TA-ERP2-MyoD)

Uchimura et al.12

KW879

CAG-rtTA-EGFP (PB-CAG-rtTA-ires-GFP)

This paper

KW1480

p5E-CAG

Laboratory of Knut Woltjen

KW140

pENTR-rtTA-Adv

Laboratory of Knut Woltjen

KW555

p3E-IRES-EGFPpA

Kwan et al.44

CC389

pDestPB53

Laboratory of Knut Woltjen

KW136

EMCV-Puro-mCherry (PB-TA-ERP2-MyoD)

This paper

KW1561

EMCV-Neo-mCherry (PB-TA-ERN2-mCherry)

This paper

KW1562

CAG-EGFP (PB-GFPa)

Laboratory of Knut Woltjen

KWan091

piggyBac transposase expression

Laboratory of Knut Woltjen,

KW158

vector (pCAG-PBase)

(Kim et al.)11

Software and algorithms

BD FACS Diva software

BD Biosciences

N/A

FlowJo software

Tree Star

N/A

Endmemo

http://www.endmemo.com/index.php

N/A

GraphPad Prism

GraphPad

N/A

BZ-X710

Keyence

N/A

iScience 26, 107685, October 20, 2023

15

ll

OPEN ACCESS

iScience

Article

RESOURCE AVAILABILITY

Lead contact

Further information and requests for resources and reagents should be directed to and will be fulfilled by the Lead Contact, Hidetoshi Sakurai

(hsakurai@cira.kyoto-u.ac.jp).

Materials availability

All unique/stable reagents generated in this study are available from lead contact, Hidetoshi Sakurai, with a completed Material Transfer

Agreement.

Data and code availability

Data: The data generated and analyzed during the current study are available from the lead contact upon reasonable request.

Code: This study did not generate/analyze [dataset/code].

Other items: Any additional information to reanalyze the data reported in this study is available from the lead contact upon reasonable

request.

EXPERIMENTAL MODEL AND STUDY PARTICIPANT ETAILS

Human iPSC lines

HC #1 hiPS cell lines served as healthy controls. HC #1 described in key resources table. HC #1 hiPSCs line was used in following approval by

the Ethics Committee of the Graduate School and Faculty of Medicine of Kyoto University (approval numbers #R0091). Karyotype was evaluated by G-banding method in LSI Medience.

Feeder-free hiPSC culture

Tet-mCherry or Tet-MyoD hiPSCs were cultured on Easy iMatrix-511 silk-coated plates (#892024, Nippi) in StemFit medium (AK02N, Ajinomoto) containing 100 mg/mL G418 (#938044, NacalaiTesque) or 0.5 mg/mL puromycin dihydrochloride (160-23151, Wako Chemicals). Cells

were passaged every 7 days using Accutase (#12679-54, NacalaiTesque) and seeded on Easy iMatrix-511 silk-coated 6-well plates in the presence of 10 mM Y-27632 (NacalaiTesque) at a density of 1.53104 cells/well for the first 2 days after plating. At 48 h after passaging, Y-27632 was

removed and replaced with StemFit medium containing the appropriate antibiotic.

METHOD DETAILS

Plasmid construction

FMDV-Neo piggyBac Gateway Destination vectors PB-TA-ERN (KW110) and PB-TAC-ERN (KW111), and EMCV-Puro piggyBac Gateway

Destination vectors PB-TA-ERP2 (KW542) and PB-TAC-ERP2 (KW543) were described in key resources table. EMCV-Neo piggyBac Gateway

Destination vectors PB-TA-ERN2 (KW1406) and PB-TAC-ERN2 (KW1413) were constructed by replacement of the FMVD IRES sequence with

an FseI-MscI EMCV IRES fragment from PB-CAG-rtTA-ires-Puro (a gift from Dr. Jonathan Draper) 42 and annealed MscI-EagI adapter oligos

(Fwd: CCACAACCATGATTGAACAAGATGGATTGCACGCAGGTTCTCC, Rev: GGCCGGAGAACCTGCGTGCAATCCATCTTGTTCAATCA

TGGTTGTGG). MyoD expression vectors were constructed by Gateway cloning of the MYOD1 cDNA from pENTR-MyoD 5 into the series

of piggyBac Gateway Destination vectors described above, resulting in EMCV-Neo-MYOD1 (PB-TA-ERN2-MyoD, KW1410), EMCV-NeoMYOD1-mCherry (PB-TAC-ERN2-MyoD, KW1415), and EMCV-Puro-MYOD1-mCherry (PB-TAC-ERP2, KW1409). The FMDV-Neo-mCherry

(PB-TAC-ERN-MyoD, KW698) and EMCV-Puro (PB-TA-ERP2-MyoD, KW879) MyoD expression vectors were constructed by Gateway cloning

of the MYOD1 cDNA into KW111 and KW542.

PB-CAG-rtTA-ires-GFP (KW1480) was constructed using Multisite Gateway cloning of p5E-CAG (KW140; Kim et al. in review), pENTR-rtTAAdv (KW555) and p3E-IRES-EGFPpA (CC389; Kwan et al., 200744) into pDestPB53 (KW136; Kim et al. in review).

Generating Tet-mCherry and Tet-MyoD hiPSCs with customized Tetracycline inducible vectors

To generate Tet-mCherry or Tet-MyoD hiPSCs, 5.0 mg of Tet-mCherry (EMCV-Neo/Puro-mCherry) or Tet-MyoD vectors and 5.0 mg of piggyBac transposase expression vector were electroporated into 13106 hiPSCs by using a NEPA21 electroporator (Nepagene). Briefly, hiPSCs

were treated with 10 mM Y-27632 1 day before electroporation. Cells were dissociated into single cells with accutase, and 1.0x106 cells

were resuspended in Opti-MEM. Tet-mCherry or Tet-MyoD vectors with pigguBac transposase expression vector were electroporated under

the conditions described in Table S1. The electroporated cells were plated on Easy iMatrix-511 silk-coated 6-well plates at 2.03105 cells per

well in StemFit medium containing Y-27632. Selection with 100 mg/mL G418 or 0.5 mg/mL puromycin was started 48 h after the electroporation. After 5 days of antibiotic selection, the cells were passaged with continuous antibiotic selection.

Direct skeletal muscle cell differentiation

Skeletal muscle cell differentiation of Tet-MyoD hiPSCs was performed as following. Briefly, 3.03105 cells were seeded on Matrigel-coated

(#356231, BD Biosciences) 6-well plates (1:100) in StemFit medium supplemented with 10 mM Y-27632. At 24 h after seeding, the medium was

16

iScience 26, 107685, October 20, 2023

iScience

Article

ll

OPEN ACCESS

changed to primate embryonic stem cell medium (RCHEMD001, ReproCELL) without Y-27632. After 24 h, 1.5 mg/mL dox (LKT Laboratories)

was added to the culture medium. After an additional 24 h, the medium was changed into differentiation medium composed of a-MEM (with

L-Gln, ribonucleosides, and deoxyribonucleosides, #21444-05, NacalaiTesque) with 5% KSR (#10828028, Invitrogen), 5% Penicillin Streptomycin Mixed solution (#09367-34, NacalaiTesque), 200 mM 2-mercaptoethanol and 1.5 mg/mL dox. Culturing was continued until day 7

with daily medium changes for immunofluorescence microscopy.

Quantification of skeletal muscle cell differentiation

The efficiency of skeletal muscle cell differentiation was analyzed using Keyence software. The total number of nuclei and MHC-positive nuclei

were counted. The differentiation efficiency was then calculated by dividing the number of MHC-positive nuclei by the total number of nuclei.

Immunofluorescence microscopy

To identify differentiated cells, myogenic markers were used for immunostaining. Differentiated cells were fixed with 4% or 2% PFA/DPBS (-) at

4 C for 10 min. After being washed in DPBS (-), the cells were treated with methanol:H2O2 (100:1) at 4 C for 15 min (in the case of d7 differentiated cells only) and subsequently blocked with Blocking One (#03953-95, NacalaiTesque) at 4 C for 45 min. The cells were then incubated

with primary antibody in 10% Blocking One in DPBS (-) with 0.2% Triton X-100 (Santa Cruz Biotechnology) at 4 C overnight. Next, the samples

were washed three times with DPBS (-) containing 0.2% Triton X-100. The cells were then incubated at room temperature for 1 h with secondary antibody in 10% Blocking One in DPBS (-) containing 0.2% Triton X-100. DAPI (1:5000) was used to counter-stain the nuclei. The samples

were observed under a BZ-X710 fluorescence microscope (Keyence).

For pluripotent marker immunostaining, the hiPSCs were fixed with 4% PFA/DPBS (-) at 4 C for 10 min. After being washed in DPBS (-), the

cells were blocked with Blocking One at 4 C for 45 min. The cells were then incubated at 4 C overnight with primary antibody in 10% Blocking

One in DPBS (-) containing 0.2% Triton X-100. Next, the samples were washed three times with DPBS (-) containing 0.2% Triton X-100. The cells

were then incubated at room temperature for 1 h with secondary antibody in 10% Blocking One in DPBS (-) and 0.2% Triton X-100. DAPI

(1:5000) was used to counter-stain the nuclei. The samples were observed under a BZ-X710 microscope at 2003 magnification. The primary

and secondary antibodies used in this study are listed in Table S2.

FACS

To isolate mCherry-positive and mCherry-negative populations of Tet-mCherry or Tet-MyoD hiPSCs, the cells were harvested as a single cell

suspension in HBSS buffer using Accutase and filtered through a cell strainer. Cells were analyzed and sorted on a FACS Aria II (BD

Biosciences).

Co-transfection of Tet-MyoD vector with additional rtTA or EGFP expression vector

For co-transfection, 5.0 mg Tet-MyoD vectors and piggyBac transposase expression vector with CAG-rtTA-EGFP or CAG-EGFP were electroporated simultaneously into hiPSCs using a NEPA21 electroporator. In total, 15 mg of plasmid vectors were electroporated into hiPSCs.

Flow cytometry

To measure mCherry or GFP reporter fluorescence, Tet-mCherry or Tet-MyoD hiPSCs were suspended in HBSS buffer and analyzed using a

BD LSR FortessaTM Cell Analyzer (BD Biosciences) with BD FACS Diva software (BD Biosciences). Hoechst staining (1:2000) was used to

exclude dead cells. Data were analyzed and generated by FlowJo software (Tree Star).

RNA isolation and reverse transcription

Total RNA was isolated using the ReliaPrep RNA cell Miniprep Kit System (Z6012, Promega) according to the manufacturer’s instructions. Residual genomic DNA was digested and removed using DNase I (Promega) treatment. First-strand cDNA was generated from extracted total

RNA using ReverTra Ace qPCR RT Master Mix with gDNA remover (FSQ301, TOYOBO). The qPCR was performed using Power SYBR Green

(#4368708, Applied Biosystems) and the Step One Plus thermal cycler (Applied Biosystems). All samples were normalized to the house-keeping gene Porphobilinogen Deaminase 1 (PBGD) 6. For absolute quantification of transgene expression levels, the plasmid backbone of the

EMCV-Neo/Puro-MYOD1-mCherry vectors were used to construct a standard curve with serial dilutions of 1/10 for qPCR analysis and estimation of the plasmid copy number was calculated using the web server Endmemo (http://www.endmemo.com/index.php). Each transgene’s

copy number was estimated by annotating its cycle threshold (Ct) value and comparing it to the standard curve of the plasmid copy number.

The primer sets used in this study are listed in Table S3.

Genomic DNA isolation and copy number integration analysis

Genomic DNA was extracted from mCherry-positive and mCherry-negative populations of Tet-mCherry or Tet-MyoD hiPSCs after sorting.

Genomic DNA was extracted using the GenElute Mammalian Genomic DNA Miniprep Kit (Sigma-Aldrich) according to the manufacturer’s

instructions. Copy number analysis was performed by qPCR as specified above, using Power SYBR Green and the Step One Plus thermal

cycler. Approximately, 8 ng of extracted genomic DNA was used for each qPCR. rtTA primer was used for detecting the Tet-mCherry or

Tet-MyoD vector and DLX5 was used as internal control. A single knocked-in Tet-MyoD vector cell line served as a single copy number control.

iScience 26, 107685, October 20, 2023

17

ll

OPEN ACCESS

iScience

Article

QUANTIFICATION AND STATISTICAL ANALYSIS

For all experiments, data are reported as the mean G SD. For comparison of two samples, p value were analyzed using either an unpaired

Student’s t-test or Paired t-test. For comparison of multiple samples, p value were analyzed using one-way ANOVA followed by Tukey’s test or

Dunnett’s test. The above p value were analyzed using GraphPad Prism.

18

iScience 26, 107685, October 20, 2023

...

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