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 「Structurally-discovered KLF4 variants accelerate and stabilize reprogramming to pluripotency」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

Structurally-discovered KLF4 variants accelerate and stabilize reprogramming to pluripotency

Borisova, Evgeniia 筑波大学 DOI:10.15068/0002006128

2023.01.17

概要

Cellular reprogramming technology enables to overturn a particular somatic cell type into another desired cell type through the man-made introduction of defined transcription factors that regulate cell differentiation patterns [68, 69]. Pluripotent reprogramming sets up a platform for bioengineering mammalian tissues and organs [98]. Man-made induced pluripotent stem cells (iPSCs) retain the distinguished features of embryonic stem cells, that is to retain one’s own pluripotency and self-regeneration ability in a limitless way in prospect becoming over 200 mammalian adult cell types [101]. The limitless self-regeneration plus the retainment of an undifferentiated, pluripotent state marks both embryonic stem and induced pluripotent stem cells as primary building blocks. Creation of iPSCs suggests an opportunity to replace malfunctioning tissues or/and organs in clinics as well as to establish malfunctioning or physiological tissues and organs for pharmaceutical research or other bio- medical research fields. The theoretical broad-scale applicability is constrained by current low and time-consuming yield of the induced pluripotent cells from primary cell sources.

Given the prospect of iPSCs-based bioengineering and regenerative medicine, industrially compatible generation of invariably high-quality iPSCs is sought after by many research groups. Among cellular reprogramming techniques, the direct lineage reprogramming, performed by activating pluripotency-related genes via viral vector mediated overexpression of selected transcription factors (OCT4, SOX2, C-MYC, and KLF4) is a canon strategy for iPSCs generation [98]. Induced reprogramming by co-expression of transcription factors displays spontaneous potentials and low reprogramming success rates. Estimated causes for faulty reprogramming may be attributed to transcription factors capacities suitable for carrying out physiological tasks only [94, 103].

Transcription factors are generally short-lived proteins governing cellular identity and integrity. By altering gene expression patterns they guide cell state transitions enabling cellular reprogramming. The well-established iPSCs induction protocols use natural transcription factors that fail to meet high-throughput production requirements for cell therapies with the induction efficiency ranging from 0.001 to 0.01 percent [1]. While the addition of certain chemicals or the change of delivery system can increase iPSC colony- forming efficiency, approaches to enforce and adapt reprogramming toolkit through mutation analysis of TFs is majorly unexplored [100, 102]. Limited capacities of natural reprogramming factors might restrain scaling up the successful and synchronous reprogramming into iPSCs.

Among pluripotency reprogramming factors, a zinc-finger transcription factor KLF4 displays a two-faced behaviour attributed to its tilt of being either an oncogene or a tumor suppressor in a context-dependent manner [26]. Consistent with other family members, KLF4 shows a high conservation in carboxyl-terminal residues (that comprise a three zinc-fingers mini- array), as opposed to minimal conservation in amino-terminal residues (that constitute a transactivation domain). The KLF4, being both transcriptional activator and/or repressor, balance multiplex cellular states through modulating cell proliferation, self-renewal, differentiation, pluripotency state, cell survival, apoptosis. Conforming to tumor suppressor activities in many instances, KLF4 is validated to modulate a cell growth arrest reactions’ cascade in a response to DNA damage (observed in intestinal adenomas, colorectal, bladder cancers) [22-23]. Conversely, KLF4 in an assembly with OCT4, SOX2 and C-MYC, transforms terminally differentiated cell phenotype into pluripotent state, eventually resulting in an unlimited proliferation potency [68]. These physiologic effects of KLF4 point out its key role in cell homeostasis and cell cycle integration. Although the scope of carried by KLF4 tasks is well-validated, only a rough estimate of how biochemical and structural properties of KLF4 make out (rule out) its mechanisms is described. The ins and outs of KLF4 functionality remain poorly understood in overall [21].

KLF proteins are necessary intermediate in processes of prevention malignant transformation and maintenance of stem cells viability. The characteristic unit of KLFs is a carboxy-terminal array of zinc fingers, an evolutionary conserved DNA-binding domain that facilitates binding to closed chromatin and recruitment of protein-protein interactions, co- activator enzymes. Compared to other DNA-binding proteins, KLFs seem to use a single discrete zinc finger domain for their three-dimensional allocation and sequence specific DNA-binding rather than a full-length stretch of a protein [28, 29].

Despite binding to identical cis-regulatory DNA sequences (elements), KLF family members display contrasting transcriptional activities in the context of cell cycle regulation. For instance, anti-proliferative activity of Klf4 antagonizes Klf5, which stimulates proliferation, in intestinal and skin epithelial cells. On the other side, Klf4 shares an overlapping (mutually exclusive) activities with Klf5, Klf2 and Klf17 in maintaining pluripotency and self-renewal of mouse embryonic stem cells [20]. Congruent with its role in sustaining pluripotency state, Klf4 plays a central role in iPSCs technology by inducing generation of pluripotent stem cells, an ability attributed only to a specific representatives of transcription factors, namely pioneer transcription factors [104].

Transcription factors exert gene regulation functions through the direct DNA binding, therefore residues constituting a DNA-binding domain are believed to underline functional specificity. One-to-one hydrogen bonds in amino acid-base interactions are viewed as major constitutors of transcription factor-DNA complex stability and reciprocal affinity, especially in the factors of sequence-specific DNA readout, such as pioneer transcription factors that regulate milestone events of cell state transition gene networks. Compared to the unspecific readout mechanism, the indiscriminate specificity towards DNA bases that is preferred by histone proteins involved in broad-base DNA interaction, a highly specific DNA readout is essential for the correct functioning of pioneer TFs that drive remarkable phenotype- changing transitions and maintain a particular cell state.

Here, determination of the several deoxyribonucleic (DNA) bases bonded-amino acid residues of a zinc finger structural class transcription factor, namely KLF4, followed by biophysical, biochemical, computational and bio functional in vitro analyses, suggested differential spatial-temporal conformation and genome-wide binding pattern of a particular residue substituted KLF4 variant that displayed remarkable iPSCs generation activities, exemplified by cell yield acceleration and yield purification for the high-quality homogenous population. Using cell culture experiments functionally non-redundant residues of a zinc finger domain were sorted out and were defined by complete loss or significant impairment of pluripotent function; they represented a major portion of the profiled residues. In vitro screening has also identified a single variant with an advanced reprogramming activity.

Using the methods of cell culturing, conducted in both mouse and human cell lines, successfully reprogrammed colony numbers were calculated (GFP Nanog Fluorescence, immunocytochemistry for Nanog) in a time-course manner indicating that L507A variant urged both faster and greater induced pluripotent stem cells compared to the WT KLF4. To further weight the importance of protein L507 site in pluripotency induction, all the natural amino acid residues were probed, uncovering a stronger favorability of smaller molecular weight (roughly proportional to the size) amino acids to better reprogramming outcomes.

Integrating ChIP-Seq and RNA-Seq data from the same cell samples revealed that KLF4 L507A variant featured enhanced transcription rates of a pluripotency-related genes, including Dppa5a (Esg1), Dsg2 (Desmoglein-2), and Klf5, that were deemed specific for the higher reprogramming efficiency. To introduce the structural reference to biological experiments’ calculations, the atomic level details of DNA-protein complex that show a formation of new direct linear bonds and a presence of a new protein conformation with nearly 33% retention rate has been provided by Molecular Dynamics simulations.

With the provided empirical evidence this study exemplifies that a particular amino acid alteration turns on a prominent practicality in an otherwise unsatisfactory natural TF. Thus, TF-based iPSC conversions utilizing remodeled TFs enable high fold change cost-efficient raise in mouse and human iPSCs generation efficiency, having only minimal alteration in the natural TF analogues.

この論文で使われている画像

参考文献

1. Huangfu, D., et al., Induction of pluripotent stem cells from primary human fibroblasts with only Oct4 and Sox2. Nat Biotechnol, 2008. 26(11): p. 1269-75.

2. Hirai, H., T. Tani, and N. Kikyo, Structure and functions of powerful transactivators: VP16, MyoD and FoxA. Int J Dev Biol, 2010. 54(11-12): p. 1589-96.

3. Boija, A., et al., Transcription Factors Activate Genes through the Phase-Separation Capacity of Their Activation Domains. Cell, 2018. 175(7): p. 1842-1855.e16.

4. Seeman, N.C., J.M. Rosenberg, and A. Rich, Sequence-specific recognition of double helical nucleic acids by proteins. Proc Natl Acad Sci U S A, 1976. 73(3): p. 804-8.

5. Zaret, K.S., Genetic programming of liver and pancreas progenitors: lessons for stem-cell differentiation. Nat Rev Genet, 2008. 9(5): p. 329-40.

6. Hirai, H., et al., Radical acceleration of nuclear reprogramming by chromatin remodeling with the transactivation domain of MyoD. Stem Cells, 2011. 29(9): p. 1349-61.

7. Soufi, A., et al., Pioneer transcription factors target partial DNA motifs on nucleosomes to initiate reprogramming. Cell, 2015. 161(3): p. 555-568.

8. Tupler, R., G. Perini, and M.R. Green, Expressing the human genome. Nature, 2001. 409(6822): p. 832-3.

9. Jantz, D., et al., The design of functional DNA-binding proteins based on zinc finger domains. Chem Rev, 2004. 104(2): p. 789-99.

10. Michael, S.F., et al., Metal binding and folding properties of a minimalist Cys2His2 zinc finger peptide. Proc Natl Acad Sci U S A, 1992. 89(11): p. 4796-800.

11. Tan, S., et al., Zinc-finger protein-targeted gene regulation: genomewide single-gene specificity. Proc Natl Acad Sci U S A, 2003. 100(21): p. 11997-2002.

12. Giesecke, A.V., R. Fang, and J.K. Joung, Synthetic protein-protein interaction domains created by shuffling Cys2His2 zinc-fingers. Mol Syst Biol, 2006. 2: p. 2006.2011.

13. Urnov, F.D., et al., Highly efficient endogenous human gene correction using designed zinc-finger nucleases. Nature, 2005. 435(7042): p. 646-51.

14. Chen, C.D., et al., A method to specifically activate the Klotho promoter by using zinc finger proteins constructed from modular building blocks and from naturally engineered Egr1 transcription factor backbone. Faseb j, 2020. 34(6): p. 7234-7246.

15. Minczuk, M., et al., Sequence-specific modification of mitochondrial DNA using a chimeric zinc finger methylase. Proc Natl Acad Sci U S A, 2006. 103(52): p. 19689- 94.

16. Jo, Y.I., H. Kim, and S. Ramakrishna, Recent developments and clinical studies utilizing engineered zinc finger nuclease technology. Cell Mol Life Sci, 2015. 72(20): p. 3819-30.

17. Miller, J.C. and C.O. Pabo, Rearrangement of side-chains in a Zif268 mutant highlights the complexities of zinc finger-DNA recognition. J Mol Biol, 2001. 313(2): p. 309-15.

18. Paillard, G., C. Deremble, and R. Lavery, Looking into DNA recognition: zinc finger binding specificity. Nucleic Acids Res, 2004. 32(22): p. 6673-82.

19. Philipsen, S. and G. Suske, A tale of three fingers: the family of mammalian Sp/XKLF transcription factors. Nucleic Acids Res, 1999. 27(15): p. 2991-3000.

20. Yamane, M., et al., Overlapping functions of Krüppel-like factor family members: targeting multiple transcription factors to maintain the naïve pluripotency of mouse embryonic stem cells. Development, 2018. 145(10).

21. Evans, P.M. and C. Liu, Roles of Krüpel-like factor 4 in normal homeostasis, cancer and stem cells. Acta Biochim Biophys Sin (Shanghai), 2008. 40(7): p. 554-64.

22. Wei, D., et al., Drastic down-regulation of Krüppel-like factor 4 expression is critical in human gastric cancer development and progression. Cancer Res, 2005. 65(7): p. 2746-54.

23. Foster, K.W., et al., Induction of KLF4 in basal keratinocytes blocks the proliferation- differentiation switch and initiates squamous epithelial dysplasia. Oncogene, 2005. 24(9): p. 1491-500.

24. Foster, K.W., et al., Increase of GKLF messenger RNA and protein expression during progression of breast cancer. Cancer Res, 2000. 60(22): p. 6488-95.

25. Hu, D., et al., Novel insight into KLF4 proteolytic regulation in estrogen receptor signaling and breast carcinogenesis. J Biol Chem, 2012. 287(17): p. 13584-97.

26. Rowland, B.D., R. Bernards, and D.S. Peeper, The KLF4 tumour suppressor is a transcriptional repressor of p53 that acts as a context-dependent oncogene. Nat Cell Biol, 2005. 7(11): p. 1074-82.

27. Chen, X., et al., Integration of external signaling pathways with the core transcriptional network in embryonic stem cells. Cell, 2008. 133(6): p. 1106-17.

28. Shields, J.M. and V.W. Yang, Identification of the DNA sequence that interacts with the gut-enriched Krüppel-like factor. Nucleic Acids Res, 1998. 26(3): p. 796-802.

29. Liu, Y., et al., Structural basis for Klf4 recognition of methylated DNA. Nucleic Acids Res, 2014. 42(8): p. 4859-67.

30. Geiman, D.E., et al., Transactivation and growth suppression by the gut-enriched Krüppel-like factor (Krüppel-like factor 4) are dependent on acidic amino acid residues and protein-protein interaction. Nucleic Acids Res, 2000. 28(5): p. 1106-13.

31. Chen, Z.Y., et al., Destabilization of Krüppel-like factor 4 protein in response to serum stimulation involves the ubiquitin-proteasome pathway. Cancer Res, 2005. 65(22): p. 10394-400.

32. Lim, K.H., et al., Critical lysine residues of Klf4 required for protein stabilization and degradation. Biochem Biophys Res Commun, 2014. 443(4): p. 1206-10.

33. Evans, P.M., et al., Kruppel-like factor 4 is acetylated by p300 and regulates gene transcription via modulation of histone acetylation. J Biol Chem, 2007. 282(47): p. 33994-4002.

34. Kawai-Kowase, K., et al., PIAS1 mediates TGFbeta-induced SM alpha-actin gene expression through inhibition of KLF4 function-expression by protein sumoylation. Arterioscler Thromb Vasc Biol, 2009. 29(1): p. 99-106.

35. Tahmasebi, S., et al., Sumoylation of Krüppel-like factor 4 inhibits pluripotency induction but promotes adipocyte differentiation. J Biol Chem, 2013. 288(18): p. 12791-804.

36. Du, J.X., B.B. McConnell, and V.W. Yang, A small ubiquitin-related modifier- interacting motif functions as the transcriptional activation domain of Krüppel-like factor 4. J Biol Chem, 2010. 285(36): p. 28298-308.

37. Ye, B., et al., Klf4 glutamylation is required for cell reprogramming and early embryonic development in mice. Nat Commun, 2018. 9(1): p. 1261.

38. Pitzschke, A., Modes of MAPK substrate recognition and control. Trends Plant Sci, 2015. 20(1): p. 49-55.

39. Kim, M.O., et al., ERK1 and ERK2 regulate embryonic stem cell self-renewal through phosphorylation of Klf4. Nat Struct Mol Biol, 2012. 19(3): p. 283-90.

40. Okita, K., T. Ichisaka, and S. Yamanaka, Generation of germline-competent induced pluripotent stem cells. Nature, 2007. 448(7151): p. 313-7.

41. Nakagawa, M., et al., Promotion of direct reprogramming by transformation-deficient Myc. Proc Natl Acad Sci U S A, 2010. 107(32): p. 14152-7.

42. Nishimura, K., et al., Development of defective and persistent Sendai virus vector: a unique gene delivery/expression system ideal for cell reprogramming. J Biol Chem, 2011. 286(6): p. 4760-71.

43. Nishimura, K., et al., Manipulation of KLF4 expression generates iPSCs paused at successive stages of reprogramming. Stem Cell Reports, 2014. 3(5): p. 915-29.

44. Li, H. and R. Durbin, Fast and accurate long-read alignment with Burrows-Wheeler transform. Bioinformatics, 2010. 26(5): p. 589-95.

45. Zhang, Y., et al., Model-based analysis of ChIP-Seq (MACS). Genome Biol, 2008. 9(9): p. R137.

46. Robinson, J.T., et al., Integrative genomics viewer. Nat Biotechnol, 2011. 29(1): p. 24-6.

47. Zhu, L.J., et al., ChIPpeakAnno: a Bioconductor package to annotate ChIP-seq and ChIP-chip data. BMC Bioinformatics, 2010. 11: p. 237.

48. Heinz, S., et al., Simple combinations of lineage-determining transcription factors prime cis-regulatory elements required for macrophage and B cell identities. Mol Cell, 2010. 38(4): p. 576-89.

49. Chen, E.Y., et al., Enrichr: interactive and collaborative HTML5 gene list enrichment analysis tool. BMC Bioinformatics, 2013. 14: p. 128.

50. Maier, J.A., et al., ff14SB: Improving the Accuracy of Protein Side Chain and Backbone Parameters from ff99SB. J Chem Theory Comput, 2015. 11(8): p. 3696- 713.

51. Krepl, M., et al., Reference simulations of noncanonical nucleic acids with different chi variants of the AMBER force field: quadruplex DNA, quadruplex RNA and Z-DNA. J Chem Theory Comput, 2012. 8(7): p. 2506-2520.

52. Zgarbova, M., et al., Toward Improved Description of DNA Backbone: Revisiting Epsilon and Zeta Torsion Force Field Parameters. J Chem Theory Comput, 2013. 9(5): p. 2339-2354.

53. Zgarbova, M., et al., Refinement of the Sugar-Phosphate Backbone Torsion Beta for AMBER Force Fields Improves the Description of Z- and B-DNA. J Chem Theory Comput, 2015. 11(12): p. 5723-36.

54. Jorgensen, W.L., et al., Comparison of Simple Potential Functions for Simulating Liquid Water. Journal of Chemical Physics, 1983. 79(2): p. 926-935.

55. Bussi, G., D. Donadio, and M. Parrinello, Canonical sampling through velocity rescaling. J Chem Phys, 2007. 126(1): p. 014101.

56. Berendsen, H.J.C., et al., Molecular-Dynamics with Coupling to an External Bath. Journal of Chemical Physics, 1984. 81(8): p. 3684-3690.

57. Hess, B., et al., LINCS: A linear constraint solver for molecular simulations. Journal of Computational Chemistry, 1997. 18(12): p. 1463-1472.

58. Hess, B., P-LINCS: A Parallel Linear Constraint Solver for Molecular Simulation. J Chem Theory Comput, 2008. 4(1): p. 116-22.

59. Darden, T., D. York, and L. Pedersen, Particle Mesh Ewald - an N.Log(N) Method for Ewald Sums in Large Systems. Journal of Chemical Physics, 1993. 98(12): p. 10089- 10092.

60. Essmann, U., et al., A Smooth Particle Mesh Ewald Method. Journal of Chemical Physics, 1995. 103(19): p. 8577-8593.

61. Hess, B., et al., GROMACS 4: Algorithms for Highly Efficient, Load-Balanced, and Scalable Molecular Simulation. J Chem Theory Comput, 2008. 4(3): p. 435-47.

62. Terada, T. and A. Kidera, Comparative molecular dynamics simulation study of crystal environment effect on protein structure. J Phys Chem B, 2012. 116(23): p. 6810-8.

63. Terada, T., et al., Understanding the roles of amino acid residues in tertiary structure formation of chignolin by using molecular dynamics simulation. Proteins, 2008. 73(3): p. 621-31.

64. Krissinel, E. and K. Henrick, Inference of macromolecular assemblies from crystalline state. J Mol Biol, 2007. 372(3): p. 774-97.

65. Schuetz, A., et al., The structure of the Klf4 DNA-binding domain links to self-renewal and macrophage differentiation. Cell Mol Life Sci, 2011. 68(18): p. 3121-31.

66. Jiang, J., et al., A core Klf circuitry regulates self-renewal of embryonic stem cells. Nat Cell Biol, 2008. 10(3): p. 353-60.

67. Jeon, H., et al., Comprehensive Identification of Krüppel-Like Factor Family Members Contributing to the Self-Renewal of Mouse Embryonic Stem Cells and Cellular Reprogramming. PLoS One, 2016. 11(3): p. e0150715.

68. Nakagawa, M., et al., Generation of induced pluripotent stem cells without Myc from mouse and human fibroblasts. Nat Biotechnol, 2008. 26(1): p. 101-6.

69. Takahashi, K. and S. Yamanaka, Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell, 2006. 126(4): p. 663- 76.

70. Chan, E.M., et al., Live cell imaging distinguishes bona fide human iPS cells from partially reprogrammed cells. Nat Biotechnol, 2009. 27(11): p. 1033-7.

71. Nishimura, K., et al., A Role for KLF4 in Promoting the Metabolic Shift via TCL1 during Induced Pluripotent Stem Cell Generation. Stem Cell Reports, 2017. 8(3): p. 787-801.

72. Kim, S.I., et al., KLF4 N-terminal variance modulates induced reprogramming to pluripotency. Stem Cell Reports, 2015. 4(4): p. 727-43.

73. Reinhardt, A., H. Kagawa, and K. Woltjen, N-Terminal Amino Acids Determine KLF4 Protein Stability in 2A Peptide-Linked Polycistronic Reprogramming Constructs. Stem Cell Reports, 2020. 14(3): p. 520-527.

74. Dhaliwal, N.K., L.E. Abatti, and J.A. Mitchell, KLF4 protein stability regulated by interaction with pluripotency transcription factors overrides transcriptional control. Genes Dev, 2019. 33(15-16): p. 1069-1082.

75. Martinez Molina, D., et al., Monitoring drug target engagement in cells and tissues using the cellular thermal shift assay. Science, 2013. 341(6141): p. 84-7.

76. Takahashi, K., et al., Induction of pluripotency in human somatic cells via a transient state resembling primitive streak-like mesendoderm. Nat Commun, 2014. 5: p. 3678.

77. Maekawa, M., et al., Direct reprogramming of somatic cells is promoted by maternal transcription factor Glis1. Nature, 2011. 474(7350): p. 225-9.

78. Liao, B., et al., MicroRNA cluster 302-367 enhances somatic cell reprogramming by accelerating a mesenchymal-to-epithelial transition. J Biol Chem, 2011. 286(19): p. 17359-64.

79. Yu, J., et al., Induced pluripotent stem cell lines derived from human somatic cells. Science, 2007. 318(5858): p. 1917-20.

80. Hong, H., et al., Suppression of induced pluripotent stem cell generation by the p53- p21 pathway. Nature, 2009. 460(7259): p. 1132-5.

81. Hotta, A., et al., Isolation of human iPS cells using EOS lentiviral vectors to select for pluripotency. Nat Methods, 2009. 6(5): p. 370-6.

82. Chronis, C., et al., Cooperative Binding of Transcription Factors Orchestrates Reprogramming. Cell, 2017. 168(3): p. 442-459 e20.

83. Pico, A.R., et al., WikiPathways: pathway editing for the people. PLoS Biol, 2008. 6(7): p. e184.

84. Kelder, T., et al., Mining biological pathways using WikiPathways web services. PLoS One, 2009. 4(7): p. e6447.

85. Slenter, D.N., et al., WikiPathways: a multifaceted pathway database bridging metabolomics to other omics research. Nucleic Acids Res, 2018. 46(D1): p. D661- D667.

86. Wienken, M., et al., MDM2 Associates with Polycomb Repressor Complex 2 and Enhances Stemness-Promoting Chromatin Modifications Independent of p53. Mol Cell, 2016. 61(1): p. 68-83.

87. Azami, T., et al., Klf5 suppresses ERK signaling in mouse pluripotent stem cells. PLoS One, 2018. 13(11): p. e0207321.

88. Aksoy, I., et al., Klf4 and Klf5 differentially inhibit mesoderm and endoderm differentiation in embryonic stem cells. Nat Commun, 2014. 5: p. 3719.

89. Parisi, S., et al., Direct targets of Klf5 transcription factor contribute to the maintenance of mouse embryonic stem cell undifferentiated state. BMC Biol, 2010. 8: p. 128.

90. Yoshioka, N., et al., Efficient generation of human iPSCs by a synthetic self- replicative RNA. Cell Stem Cell, 2013. 13(2): p. 246-54.

91. Lee, S.Y., et al., Glis family proteins are differentially implicated in the cellular reprogramming of human somatic cells. Oncotarget, 2017. 8(44): p. 77041-77049.

92. Wang, L., et al., NANOG and LIN28 dramatically improve human cell reprogramming by modulating LIN41 and canonical WNT activities. Biol Open, 2019. 8(12).

93. Kim, Y.S., et al., Identification of Glis1, a novel Gli-related, Kruppel-like zinc finger protein containing transactivation and repressor functions. J Biol Chem, 2002. 277(34): p. 30901-13.

94. Hayashi, Y., et al., Structure-based discovery of NANOG variant with enhanced properties to promote self-renewal and reprogramming of pluripotent stem cells. Proc Natl Acad Sci U S A, 2015. 112(15): p. 4666-71.

95. Hashimoto, H., et al., Distinctive Klf4 mutants determine preference for DNA methylation status. Nucleic Acids Res, 2016. 44(21): p. 10177-10185.

96. Pabo, C.O., E. Peisach, and R.A. Grant, Design and selection of novel Cys2His2 zinc finger proteins. Annu Rev Biochem, 2001. 70: p. 313-40.

97. Pettersen, E.F., et al., UCSF Chimera--a visualization system for exploratory research and analysis. J Comput Chem, 2004. 25(13): p. 1605-12.

98. Yamanaka, S., Blau H.M., Nuclear reprogramming to a pluripotent state by three approaches. Nature, 2010. 10;465(7299): p. 704-12.

99. Hochedlinger, K., Jaenisch, R., Nuclear reprogramming and pluripotency. Nature, 2006. 29;441(7097): p. 1061-7.

100. Morris, S.A., Direct lineage reprogramming via pioneer factors; a detour through developmental gene regulatory networks. Development, 2016. 1;143(15): p. 2696-705.

101. Tapia, N., Han, D.W., Scholer, H.R., Restoring stem cell function in aged tissues by direct reprogramming? Cell Stem Cell, 2012. 14;10(6): p. 653–656.

102. Jauch, R., Cell fate reprogramming through engineering of native transcription factors. Curr Opin Genet Dev, 2018. 52, p. 109-116.

103. Tan, D.S., Scholer, H.R., Jauch, R., et al., Directed Evolution of an Enhanced POU Reprogramming Factor for Cell Fate Engineering. Molecular Biology and Evolution, 2021. 38(7): p. 2854-2868.

104. Di Giammartino D.C., et al., KLF4 is involved in the organization and regulation of pluripotency-associated three-dimensional enhancer networks. Nat Cell Biol, 2019. 21(10): p. 1179-1190.

105. Liu, Y., Beyer, A., Aebersold, R., On the Dependency of Cellular Protein Levels on mRNA Abundance. Cell, 2016. 21;165(3): p. 535-50.

106. Perl, K., Ushakov, K., Pozniak, Y. et al., Reduced changes in protein compared to mRNA levels across non-proliferating tissues. BMC Genomics, 2017. 18(1): p. 305

107. Gedeon, T., Bokes, P., Delayed protein synthesis reduces the correlation between mRNA and protein fluctuations. Biophys J, 2012. 8;103(3): p. 377-385.

108. Li, W., Zhou, H., Abujarour, R., Zhu, S., Young Joo, J., Lin, T., Hao, E., Schöler, H. R., Hayek, A., Ding, S., Generation of human-induced pluripotent stem cells in the absence of exogenous Sox2. Stem cells, 2009. 27(12): p. 2992–3000.

参考文献をもっと見る