PRMT1通过对MyoD精氨酸甲基化修饰调控成肌细胞分化
详细信息    本馆镜像全文|  推荐本文 |  |   获取CNKI官网全文
摘要
表观遗传学是指在基因的DNA序列不发生改变的情况下,基因的表达水平与功能发生改变,并产生可遗传的表型组,主要包括DNA甲基化、组蛋白修饰、染色质重塑、RNA介导的调控。组蛋白修饰是表观遗传学中很重要的调控方式。包括组蛋白的甲基化、乙酰化、磷酸化、泛素化等,它对基因的激活和抑制起着重要作用。
     MyoD是碱性螺旋-环-螺旋(basic helix-loop-helix, bHLH)家族的一个重要蛋白,在肌肉分化过程中起关键作用。我们以低血清诱导小鼠C2C12细胞作为肌肉分化的模型,并对其分化机制以及分化过程中早期分化标志—myogenin基因启动子区转录相关因子的结合进行研究,特别是精氨酸甲基转移酶修饰对myogenin基因表达调控机制。
     C2C12细胞,主要呈三角形和纺锤形,少数多边形;而低血清诱导处理后,C2C12细胞逐渐呈现典型的分化形态,如胞体伸长,出现多核肌纤维类似结构并整齐排列等。我们利用甲基转移酶抑制剂AdOx处理C2C12细胞后发现,C2C12细胞的分化受到抑制。包括细胞形态学上不能出现多核肌纤维类似结构,分化标志myogenin的蛋白水平及mRNA水平的抑制。而组蛋白修饰酶PRMT1、PCAF及CARM1的蛋白质水平在药物处理前后变化不明显。表明甲基化修饰可能是肌肉分化过程中的必需过程。
     双荧光素酶报告基因分析发现,在转染的各PRMT家族成员中,PRMT1和CARM1显著地增强myogenin基因启动子活性。我们进一步共转染组蛋白乙酰基转移酶PCAF、当PRMT1与CARM1及PCAF共转染时,比单独转染三者任何一个,或二二组合,激活效应更明显。说明PRMT1可与CARM1及PCAF协同激活MyoD介导的myogenin基因启动子活性。PRMT1是肌肉分化主导开关(master switch) MyoD的转录共激活子。
     染色质免疫沉淀(Chromatin Immunoprecipitation, ChIP)分析是近些年发展起来检测蛋白因子在体内与染色质的结合,进而为蛋白因子参与染色质水平基因转录调控提供证据的方法。本研究运用该方法,检测了低血清诱导的C2C12细胞分化过程中,肌细胞分化决定因子MyoD、乙酰转移酶PCAF与精氨酸甲基转移酶CARM1及PRMT1在myogenin基因启动子区染色质的结合情况。这些转录因子在诱导3h后即可结合到染色质。
     GST-pulldown实验显示,MyoD与PRMT1共存于一个复合物中。我们进一步确定了MyoD的氨基端及bHLH结构域与PRMT1相互作用,而MyoD的羧基端(MyoD162-318)则不能与PRMT1相互作用。将MyoD亚克隆至pEGFP-C1表达绿色荧光蛋白标签的融合蛋白,将PRMT1克隆至pmCherry-C1载体表达带红色荧光蛋白标签的融合蛋白。将GFP融合的MyoD与Cherry融合的PRMT1共转染C2C12细胞,48h后用激光扫描共聚焦荧光显微镜观察。我们发现,GFP-MyoD主要定位于核内,而Cherry-PRMT1则在整个细胞广泛分布,将二者荧光融合(merge)发现,绿色荧光的MyoD与红色荧光的PRMT1在细胞核内很好的融合为黄色,说明二者在细胞核内定位于相同位置。
     我们通过体外GST-pulldown实验以及体内免疫共沉淀(CoIP)实验证实CARM1和PCAF共存于一个复合物中,并且这种相互作用不受AdOx抑制剂的影响的。通过分子生物学手段,我们进一步找到两者的具体结合部位。其中CARM1的催化结构域及C端结构域可与PCAF相互作用,而PCAF主要是通过HAT结构域(PCAF352-658)与CARM1相互作用。
     利用体外甲基化反应体系及放射性自显影技术,我们首次发现PRMT家族酶的各个成员,只有PRMT1可甲基化修饰MyoD的精氨酸残基。为了确定MyoD甲基化修饰在细胞内是不是也发生,我们将带Flag标签的MyoD转染HEK 293T细胞,标记3H-甲硫氨酸(3H-Met)的同时加入蛋白质合成抑制剂。免疫亲和沉淀Flag-MyoD,放射自显影。结果显示,MyoD在细胞内存在甲基化修饰,并且AdOx可以抑制MyoD的甲基化修饰。为了确定MyoD的甲基化修饰位点,我们在细菌中表达并纯化了MyoD的N端及C端的删切突变体,加入GST-PRMT1后进行体外甲基化反应并放射性自显影,结果显示MyoD的甲基化修饰位点位于氨基酸162-318之间,即bHLH及C端结构域。
     为了确定MyoD甲基化修饰的精确位点,我们在体外用PRMT1甲基化修饰MyoD,ESI质谱鉴定得到带两个电荷且质荷比为910.36的MyoD片段,即RQNGYDTAYYSEAVR(氨基酸221-235),二级质谱鉴定得到上述片段的第一个精氨酸(全长MyoD的221位精氨酸)存在双甲基化修饰。
     综上结果,我们认为PRMT1对C2C12细胞的诱导分化过程中具有正调节作用。通过该类酶促的蛋白质精氨酸甲基化,以及其他调节因子对组蛋白的修饰或转录因子的活化,可有效地促进myogenin基因表达,并导致C2C12细胞的分化。我们首次发现PRMT1可甲基化修饰MyoD的精氨酸残基,深入研究PRMT1的调控机制有助于对肌肉细胞诱导分化机制的理解。
Epigenetics is the study of stable alternation in gene expression and function without changing DNA sequence and leading to inheritable phenotype, mainly including DNA methylation, histone modification, chromatin remodeling and non-coding RNA-mediated regulation. Histone modifications including histone methylation, acetylation, phosphorylation and ubiquitination play crucial roles in gene activation and inactivation.
     MyoD, one of bHLH protein, play a crucial role in muscle differentiation. In our experiments, Low serum medium was used to induce C2C12 cells to differentiate, and the differentiation mechanism and the binding of associated transcription factors and chromatin modifying enzymes, especially protein arginine methyltransferase 1 to the myogenin promoter were investigated.
     The morphology of C2C12 cells are triangular or spindly, and fewer are polygonal. After transferred to low serum medium, the C2C12 cells align regularly and acquire typical myogenic modality gradually, such as a more elongated shape, and multinucleated myofiber-like structure. when the C2C12 cells are treated with adenosine dialdehyde (AdOx) to block the PRMT, Low serum medium induced C2C12 cells differentiation was inhibit. The morphology of C2C12 cells are more big and don't present a multinucleated myofiber-like structure. Moreover, the increase of protein level and mRNA of myogenin was prevented by AdOx during C2C12 cells differentiation. In contrast, the protein level of PRMT1, PCAF and CARM1 was not changed during C2C12 cells differentiation in the presence of AdOx. These data indicate that arginie methylation is involved in muscle differentiation.
     Dual-luciferase reporter assay showed that PRMT1 and CARM1 can enhanced the myogenin gene reporter activity when we transfected the C2C12 cells with each member of PRMTs. We further co-transfected C2C12 cells with PRMT1, CARM1 and PCAF, the myogenin gene reporter activity assays showed that PRMT1 synergistically coactivates together with CARM1 and PCAF MyoD-mediated transcription at the myogenin promoter.
     Chromatin Immunoprecipitation (ChIP) assay was developed in recent years to detect protein factors that bind to chromatin. The binding of MyoD, PRMT1, PCAF and CARM1 to the myogenin promoter in control or differentiated C2C12 cells induced by low serum medium was studied by this method. We found that these factors were recruited on the myogenin gene promoter after induced for 3 hours in the low serum medium.
     In vitro GST-pulldown assays showed that MyoD and PRMT1 exist in a complex, And the binding domain required for the interaction between MyoD and PRMT1 was determined, while the C-terminal domain of MyoD don't interact with PRMTl. We further subcloned MyoD into pEGFP-C1 and PRMT1 into pmCherry-C1, which express GFP-MyoD and Cherry-PRMT1 respectively. When C2C12 were co-transfected with these fusion protein plasmids, MyoD and PRMT1 were visualized by autofluorescence (green) of GFP and (red) of Cherry with a confocal microscopy, respectively. Green fluorescence of GFP-MyoD and red fluorescence of Cherry-PRMT1 merge well (yellow) in nucleus.
     To demonstrate the direct interaction between CARM1 with PCAF, the in vitro and in vivo binding assays were done. And the binding domain required for the interaction between CARM1 and PCAF was determined. C-terminal domain of CARM1 (CARM11-141) can not interacts with PCAF, and PCAF HAT (histone acetyltransferase) domain interacts with CARM1.
     In vitro methylation reaction assay and autoradiography indicated that, among the members of protein arginine methyltransferase family, only PRMT1 methylates MyoD. To investigate the occurrence of in vivo methylation of MyoD, we performed metabolical labeling of HEK293T cells with L-[methy-3H]-methionine. Cells were transfected with Flag-tagged MyoD in the absence and presence of AdOx and simultaneously metabolical labeling was performed for 1h in the presence of translational inhibitors. Flag-MyoD protein was immunopurified from the cell extracts, resolved by SDS-PAGE and methylation was detected by autoradiography. The data indicate that in vivo methylation of MyoD occurs and was inhibited in the presence of AdOx. In order to map the methylation site(s) in MyoD, we expressed several N-and C-terminal deletion constructs of MyoD as GST-tagged fusions in bacteria. In vitro methylation in the presence of GST-PRMT1 revealed that the methylation site(s) of PRMT1 in MyoD are located between amino acid 102 and 318, and which encompass the bHLH domain and the C-terminal domain.
     To identify the exact site(s) of PRMT1 methylation in MyoD we methylated MyoD in the presence of GST-PRMT1 in vitro. The ESI mass spectrum revealed a doubly charged peak at m/z 910.36 corresponding to the dimethylated sequence of the tryptic MyoD fragment RQNGYDTAYYSEAVR (amino acids 221-235). Fragment ion (MS/MS) spectrum resulting from this doubly charged precursor clearly indicated that the first arginine within the peptide (R221 in the full-length context of MyoD) is dimethylated.
     In conclusion, PRMTs play an positive role in regulating the low serum medium induced C2C12 cells differentiation. Histone arginine methylation by PRMTs along with other histone modifications and the activation of transcription factors that jointly exert an increased expression of myogenin gene and result in the C2C12 cells to differentiation. We found for the first time that PRMT1 methylate MyoD in vitro and in vivo. Further studies on the regulatory functions of PRMT1 in C2C12 cell differentiation shed lights on our understanding of muscle differentiation.
引文
1. Troffer-Charlier, N., et al., Functional insights from structures of coactivator-associated arginine methyltransferase 1 domains. EMBO J,2007.26(20):p.4391-401.
    2. Valentna Migliori, S.P., Marco Bezzi, Ernesto Guccione, Arginine/lysine-methyl/methyl switches:biochemical role of histone arginine methylation in transcriptional regulation. epigenomics,2010.2(1):p.119-137.
    3. Tapscott, S.J., The circuitry of a master switch:Myod and the regulation of skeletal muscle gene transcription. Development,2005.132(12):p.2685-95.
    4. Borycki, A.G. and C.P. Emerson, Jr., Study of skeletal myogenesis in cultures of unsegmented paraxial mesoderm. Methods Mol Biol,2000.137:p.351-7.
    5. Cossu, G. and U. Borello, Wnt signaling and the activation of myogenesis in mammals. Embo J,1999.18(24):p.6867-72.
    6. Hauschka, S.D., The embryonic origin of skeletal muscle, in'The scientific basis of Myology".1994:Academic Press.3-72.
    7. Christ, B., and Ordhal, C. P., Early stages of chick somite development. Anat Embryol, 1994.191:p.381-396.
    8. Tajbakhsh S, B.M., The birth of muscle progenitor cells in the mouse:spatiotemporal considerations, in Current topics in developmental biology:Somitogenesis, O. CP, Editor. 2000, Academic Press, p.225-268.
    9. Cossu, G., S. Tajbakhsh, and M. Buckingham, How is myogenesis initiated in the embryo? Trends Genet,1996.12(6):p.218-23.
    10. Tajbakhsh, S., et al., Redefining the genetic hierarchies controlling skeletal myogenesis: Pax-3 and Myf-5 act upstream of MyoD. Cell,1997.89(1):p.127-38.
    11. Braun, T., et al., Targeted inactivation of the muscle regulatory gene Myf-5 results in abnormal rib development and perinatal death. Cell,1992.71(3):p.369-82.
    12. Kablar, B., et al., MyoD and Myf-5 differentially regulate the development of limb versus trunk skeletal muscle. Development,1997.124(23):p.4729-38.
    13. Kablar, B., et al., MyoD and Myf-5 define the specification of musculature of distinct embryonic origin. Biochem Cell Biol,1998.76(6):p.1079-91.
    14. Rudnicki, M.A., et al., MyoD or Myf-5 is required for the formation of skeletal muscle. Cell,1993.75(7):p.1351-9.
    15. Munsterberg, A.E., et al., Combinatorial signaling by Sonic hedgehog and Wnt family members induces myogenic bHLH gene expression in the somite. Genes Dev,1995.9(23): p.2911-22.
    16. Pourquie, O., et al., Lateral and axial signals involved in avian somite patterning:a role for BMP4. Cell,1996.84(3):p.461-71.
    17. Blau, H.M., C.P. Chiu, and C. Webster, Cytoplasmic activation of human nuclear genes in stable heterocaryons. Cell,1983.32(4):p.1171-80.
    18. Wright, W.E., Expression of differentiated functions in heterokaryons between skeletal myocytes, adrenal cells, fibroblasts and glial cells. Exp Cell Res,1984.151(1):p.55-69.
    19. Davis, R.L., H. Weintraub, and A.B. Lassar, Expression of a single transfected cDNA converts fibroblasts to myoblasts. Cell,1987.51(6):p.987-1000.
    20. Wright, W.E., D.A. Sassoon, and V.K. Lin, Myogenin, a factor regulating myogenesis, has a domain homologous to MyoD. Cell,1989.56(4):p.607-17.
    21. Buckingham M, T.S., Myogenic cell specification during somitogenesis, in Cell lineage and fate determination, M. SA, Editor.1999, Academic Press.
    22. Choi, J., et al., MyoD converts primary dermal fibroblasts, chondroblasts, smooth muscle, and retinal pigmented epithelial cells into striated mononucleated myoblasts and multinucleated myotubes. Proc Natl Acad Sci U S A,1990.87(20):p.7988-92.
    23. Weintraub, H., et al., The myoD gene family:nodal point during specification of the muscle cell lineage. Science,1991.251(4995):p.761-6.
    24. Ferre-D'Amare, A.R., et al., Recognition by Max of its cognate DNA through a dimeric b/HLH/Z domain. Nature,1993.363(6424):p.38-45.
    25. Ma, P.C., et al., Crystal structure of MyoD bHLH domain-DNA complex:perspectives on DNA recognition and implications for transcriptional activation. Cell,1994.77(3):p. 451-9.
    26. Lassar, A.B., et al., Functional activity of myogenic HLH proteins requires hetero-oligomerization with E12/E47-like proteins in vivo. Cell,1991.66(2):p.305-15.
    27. Murre, C., P.S. McCaw, and D. Baltimore, A new DNA binding and dimerization motif in immunoglobulin enhancer binding, daughterless, MyoD, and myc proteins. Cell,1989. 56(5):p.777-83.
    28. Benezra, R., et al., The protein Id:a negative regulator of helix-loop-helix DNA binding proteins. Cell,1990.61(1):p.49-59.
    29. Langlands, K., et al., Differential interactions of Id proteins with basic-helix-loop-helix transcription factors. J Biol Chem,1997.272(32):p.19785-93.
    30. Lu, J., et al., MyoR:a muscle-restricted basic helix-loop-helix transcription factor that antagonizes the actions of MyoD. Proc Natl Acad Sci U S A,1999.96(2):p.552-7.
    31. Maleki, S.J., C.A. Royer, and B.K. Hurlburt, MyoD-E12 heterodimers and MyoD-MyoD homodimers are equally stable. Biochemistry,1997.36(22):p.6762-7.
    32. Lenormand, J.L., et al., Mos activates myogenic differentiation by promoting heterodimerization of MyoD and E12 proteins. Mol Cell Biol,1997.17(2):p.584-93.
    33. Pelpel, K., et al., Mutation of MyoD-Ser237 abolishes its up-regulation by c-Mos. FEBS Lett,2000.474(2-3):p.233-7.
    34. Brennan, T.J., T. Chakraborty, and E.N. Olson, Mutagenesis of the myogenin basic region identifies an ancient protein motif critical for activation of myogenesis. Proc Natl Acad Sci US A,1991.88(13):p.5675-9.
    35. Davis, R.L., et al., The MyoD DNA binding domain contains a recognition code for muscle-specific gene activation. Cell,1990.60(5):p.733-46.
    36. Huang, J., H. Weintraub, and L. Kedes, Intramolecular regulation of MyoD activation domain conformation and function. Mol Cell Biol,1998.18(9):p.5478-84.
    37. Davis, R.L. and H. Weintraub, Acquisition of myogenic specificity by replacement of three amino acid residues from MyoD into E12. Science,1992.256(5059):p.1027-30.
    38. Dezan, C., et al., Acquisition of myogenic specificity through replacement of one amino acid of MASH-1 and introduction of an additional alpha-helical turn. Biol Chem,1999. 380(6):p.705-10.
    39. Tapscott, S.J., et al., MyoDl:a nuclear phosphoprotein requiring a Myc homology region to convert fibroblasts to myoblasts. Science,1988.242(4877):p.405-11.
    40. Braun, T., et al., Myf-6, a new member of the human gene family of myogenic determination factors:evidence for a gene cluster on chromosome 12. EMBO J,1990. 9(3):p.821-31.
    41. Mak, K.L., et al., The MRF4 activation domain is required to induce muscle-specific gene expression. Mol Cell Biol,1992.12(10):p.4334-46.
    42. Olson, E., Activation of muscle-specific transcription by myogenic helix-loop-helix proteins. Symp Soc Exp Biol,1992.46:p.331-41.
    43. Gossett, L.A., et al., A new myocyte-specific enhancer-binding factor that recognizes a conserved element associated with multiple muscle-specific genes. Mol Cell Biol,1989. 9(11):p.5022-33.
    44. Black, B.L. and E.N. Olson, Transcriptional control of muscle development by myocyte enhancer factor-2 (MEF2) proteins. Annu Rev Cell Dev Biol,1998.14:p.167-96.
    45. Molkentin, J.D., et al., Mutational analysis of the DNA binding, dimerization, and transcriptional activation domains of MEF2C. Mol Cell Biol,1996.16(6):p.2627-36.
    46. Pollock, R. and R. Treisman, Human SRF-related proteins:DNA-binding properties and potential regulatory targets. Genes Dev,1991.5(12A):p.2327-41.
    47. Yu, Y.T., et al., Human myocyte-specific enhancer factor 2 comprises a group of tissue-restricted MADS box transcription factors. Genes Dev,1992.6(9):p.1783-98.
    48. Hasty, P., et al., Muscle deficiency and neonatal death in mice with a targeted mutation in the myogenin gene. Nature,1993.364(6437):p.501-6.
    49. Kaushal, S., et al., Activation of the myogenic lineage by MEF2A, a factor that induces and cooperates with MyoD. Science,1994.266(5188):p.1236-40.
    50. Nabeshima, Y., et al., Myogenin gene disruption results in perinatal lethality because of severe muscle defect. Nature,1993.364(6437):p.532-5.
    51. Ornatsky, O.I., J.J. Andreucci, and J.C. McDermott, A dominant-negative form of transcription factor MEF2 inhibits myogenesis. J Biol Chem,1997.272(52):p.33271-8.
    52. Molkentin, J.D., et al., Cooperative activation of muscle gene expression by MEF2 and myogenic bHLHproteins. Cell,1995.83(7):p.1125-36.
    53. Bi, W., C.J. Drake, and J.J. Schwarz, The transcription factor MEF2C-null mouse exhibits complex vascular malformations and reduced cardiac expression of angiopoietin I and VEGF. Dev Biol,1999.211(2):p.255-67.
    54. Lin, Q., et al., Control of mouse cardiac morphogenesis and myogenesis by transcription factor MEF2C. Science,1997.276(5317):p.1404-7.
    55. Kolodziejczyk, S.M., et al., MEF2 is upregulated during cardiac hypertrophy and is required for normal post-natal growth of the myocardium. Curr Biol,1999.9(20):p. 1203-6.
    56. Sif, S., ATP-dependent nucleosome remodeling complexes:enzymes tailored to deal with chromatin. J Cell Biochem,2004.91(6):p.1087-98.
    57. Dilworth, F.J., et al., In vitro transcription system delineates the distinct roles of the coactivators pCAF and p300 during MyoD/E47-dependent transactivation. Proc Natl Acad Sci U S A,2004.101(32):p.11593-8.
    58. Eckner, R., et al., Interaction and functional collaboration of p300/CBP and bHLH proteins in muscle and B-cell differentiation. Genes Dev,1996.10(19):p.2478-90.
    59. Puri, P.L., et al., Differential roles of p300 and PCAF acetyltransferases in muscle differentiation. Mol Cell,1997.1(1):p.35-45.
    60. Sartorelli, V., et al., Acetylation of MyoD directed by PCAF is necessary for the execution of the muscle program. Mol Cell,1999.4(5):p.725-34.
    61. Lemercier, C., et al., mHDA1/HDAC5 histone deacetylase interacts with and represses MEF2A transcriptional activity. J Biol Chem,2000.275(20):p.15594-9.
    62. Lu, J., et al., Signal-dependent activation of the MEF2 transcription factor by dissociation from histone deacetylases. Proc Natl Acad Sci U S A,2000.97(8):p. 4070-5.
    63. Miska, E.A., et al., HDAC4 deacetylase associates with and represses the MEF2 transcription factor. EMBO J,1999.18(18):p.5099-107.
    64. Sparrow, D.B., et al., MEF-2 function is modified by a novel co-repressor, MITR. EMBO J,1999.18(18):p.5085-98.
    65. Wang, A.H., et al., HDAC4, a human histone deacetylase related to yeast HDA1, is a transcriptional corepressor. Mol Cell Biol,1999.19(11):p.7816-27.
    66. Bailey, P., et al., The nuclear receptor corepressor N-CoR regulates differentiation: N-CoR directly interacts with MyoD. Mol Endocrinol,1999.13(7):p.1155-68.
    67. 沈珝琲,染色质与表现遗传调控.1 ed.2006,北京:高等教育出版社.
    68. de la Serna, I.L., K.A. Carlson, and A.N. Imbalzano, Mammalian SWI/SNF complexes promote MyoD-mediated muscle differentiation. Nat Genet,2001.27(2):p.187-90.
    69. Najbauer, J., et al., Peptides with sequences similar to glycine, arginine-rich motifs in proteins interacting with RNA are efficiently recognized by methyltransferase(s) modifying arginine in numerous proteins. J Biol Chem,1993.268(14):p.10501-9.
    70. Bedford, M.T. and S.G. Clarke, Protein arginine methylation in mammals:who, what, and why. Mol Cell,2009.33(1):p.1-13.
    71. Hughes, R.M. and M.L. Waters, Arginine methylation in a beta-hairpin peptide: implications for Arg-pi interactions, DeltaCp(o), and the cold denatured state. J Am Chem Soc,2006.128(39):p.12735-42.
    72. Stetler, A., et al., Identification and characterization of the methyl arginines in the fragile X mental retardation protein Fmrp. Hum Mol Genet,2006.15(1):p.87-96.
    73. Krause, C.D., et al., Protein arginine methyltransferases:evolution and assessment of their pharmacological and therapeutic potential. Pharmacol Ther,2007.113(1):p.50-87.
    74. Cheng, X., R.E. Collins, and X. Zhang, Structural and sequence motifs of protein (histone) methylation enzymes. Annu Rev Biophys Biomol Struct,2005.34:p.267-94.
    75. Katz, J.E., M. Dlakic, and S. Clarke, Automated identification of putative methyltransferases from genomic open reading frames. Mol Cell Proteomics,2003.2(8): p.525-40.
    76. Cook, J.R., et al., FBXO11/PRMT9, a new protein arginine methyltransferase, symmetrically dimethylates arginine residues. Biochem Biophys Res Commun,2006. 342(2):p.472-81.
    77. Lin, W.J., et al., The mammalian immediate-early TIS21 protein and the leukemia-associated BTGl protein interact with a protein-arginine N-methyltransferase. J Biol Chem,1996.271(25):p.15034-44.
    78. Pawlak, M.R., et al., Arginine N-methyltransferase 1 is required for early postimplantation mouse development, but cells deficient in the enzyme are viable. Mol Cell Biol,2000.20(13):p.4859-69.
    79. Tang, J., et al., PRMT1 is the predominant type I protein arginine methyltransferase in mammalian cells. J Biol Chem,2000.275(11):p.7723-30.
    80. Katsanis, N., M.L. Yaspo, and E.M. Fisher, Identification and mapping of a novel human gene, HRMT1L1, homologous to the rat protein arginine N-methyltransferase I (PRMT1) gene. Mamm Genome,1997.8(7):p.526-9.
    81. Tang, J., et al., PRMT 3, a type I protein arginine N-methyltransferase that differs from PRMT1 in its oligomerization, subcellular localization, substrate specificity, and regulation. J Biol Chem,1998.273(27):p.16935-45.
    82. Chen, D., et al., Regulation of transcription by a protein methyltransferase. Science, 1999.284(5423):p.2174-7.
    83. Branscombe, T.L., et al., PRMT5 (Janus kinase-binding protein I) catalyzes the formation of symmetric dimethylarginine residues in proteins. J Biol Chem,2001. 276(35):p.32971-6.
    84. Pollack, B.P., et al., The human homologue of the yeast proteins Skbl and Hsl7p interacts with Jak kinases and contains protein methyltransferase activity. J Biol Chem,1999. 274(44):p.31531-42.
    85. Frankel, A., et al., The novel human protein arginine N-methyltransferase PRMT6 is a nuclear enzyme displaying unique substrate specificity. J Biol Chem,2002.277(5):p. 3537-43.
    86. Gros, L., et al., Identification of new drug sensitivity genes using genetic suppressor elements:protein arginine N-methyltransferase mediates cell sensitivity to DNA-damaging agents. Cancer Res,2003.63(1):p.164-71.
    87. Lee, J., et al., PRMT8, a new membrane-bound tissue-specific member of the protein arginine methyltransferase family. J Biol Chem,2005.280(38):p.32890-6.
    88. Yadav, N., et al., Specific protein methylation defects and gene expression perturbations in coactivator-associated arginine methyltransferase 1-deficient mice. Proc Natl Acad Sci U S A,2003.100(11):p.6464-8.
    89. Scott, H.S., et al., Identification and characterization of two putative human arginine methyltransferases (HRMT1L1 and HRMT1L2). Genomics,1998.48(3):p.330-40.
    90. Herrmann, F., et al., Dynamics of human protein arginine methyltransferase 1(PRMT1) in vivo. J Biol Chem,2005.280(45):p.38005-10.
    91. Lim, I.K., et al., Enzymatic methylation of recombinant TIS21 protein-arginine residues. Biochem Mol Biol Int,1998.45(5):p.871-8.
    92. Gary, J.D. and S. Clarke, RNA and protein interactions modulated by protein arginine methylation. Prog Nucleic Acid Res Mol Biol,1998.61:p.65-131.
    93. Smith, J.J., et al., Unusual sites of arginine methylation in Poly(A)-binding protein II and in vitro methylation by protein arginine methyltransferases PRMTI and PRMT3. J Biol Chem,1999.274(19):p.13229-34.
    94. Klein, S., et al., Biochemical analysis of the arginine methylation of high molecular weight fibroblast growth factor-2. J Biol Chem,2000.275(5):p.3150-7.
    95. Tang, J., P.N. Kao, and H.R. Herschman, Protein-arginine methyltransferase I, the predominant protein-arginine methyltransferase in cells, interacts with and is regulated by interleukin enhancer-binding factor 3. J Biol Chem,2000.275(26):p.19866-76.
    96. Mowen, K.A., et al., Arginine methylation of STAT1 modulates IFNalpha/beta-induced transcription. Cell,2001.104(5):p.731-41.
    97. Strahl, B.D., et al., Methylation of histone H4 at arginine 3 occurs in vivo and is mediated by the nuclear receptor coactivator PRMTI. Curr Biol,2001.11(12):p. 996-1000.
    98. Wang, H., et al., Methylation of histone H4 at arginine 3 facilitating transcriptional activation by nuclear hormone receptor. Science,2001.293(5531):p.853-7.
    99. Lee, J. and M.T. Bedford, PABP1 identified as an arginine methyltransferase substrate using high-density protein arrays. EMBO Rep,2002.3(3):p.268-73.
    100. Batonnet-Pichon, S., et al., MyoD undergoes a distinct G2/M-specific regulation in muscle cells. Exp Cell Res,2006.312(20):p.3999-4010.
    101. Kitzmann, M., et al., cdkl-and cdk2-mediated phosphorylation of MyoD Ser200 in growing C2 myoblasts:role in modulating MyoD half-life and myogenic activity. Mol Cell Biol,1999.19(4):p.3167-76.
    102. Song, A., et al., Phosphorylation of nuclear MyoD is required for its rapid degradation. Mol Cell Biol,1998.18(9):p.4994-9.
    103. Tintignac, L.A., et al., Cyclin E-cdk2 phosphorylation promotes late G1-phase degradation of MyoD in muscle cells. Exp Cell Res,2000.259(1):p.300-7.
    104. Tintignac, L.A., et al., Mutant MyoD lacking Cdc2 phosphorylation sites delays M-phase entry. Mol Cell Biol,2004.24(4):p.1809-21.
    105. Puri, P.L., et al., A myogenic differentiation checkpoint activated by genotoxic stress. Nat Genet,2002.32(4):p.585-93.
    106. Duquet, A., et al., Acetylation is important for MyoD function in adult mice. EMBO Rep, 2006.7(11):p.1140-6.
    107. Polesskaya, A., et al., CREB-binding protein/p300 activates MyoD by acetylation. J Biol Chem,2000.275(44):p.34359-64.
    108. Polesskaya, A. and A. Harel-Bellan, Acetylation of MyoD by p300 requires more than its histone acetyltransferase domain. J Biol Chem,2001.276(48):p.44502-3.
    109. Liu, L.N., P. Dias, and P.J. Houghton, Mutation of Thr115 in MyoD positively regulates function in murine fibroblasts and human rhabdomyosarcoma cells. Cell Growth Differ, 1998.9(9):p.699-711.
    110. Lagirand-Cantaloube, J., et al., Inhibition of atrogin-1/MAFbx mediated MyoD proteolysis prevents skeletal muscle atrophy in vivo. PLoS One,2009.4(3):p. e4973.
    111. McMahon, D.K., et al., C2C12 cells:biophysical, biochemical, and immunocytochemical properties. Am J Physiol,1994.266(6 Pt 1):p. C1795-802.
    112. Melo, F., D.J. Carey, and E. Brandan, Extracellular matrix is required for skeletal muscle differentiation but not myogenin expression. J Cell Biochem,1996.62(2):p.227-39.
    113. Thinakaran, G., J. Ojala, and J. Bag, Expression of c-jun/AP-1 during myogenic differentiation in mouse C2C12 myoblasts. FEBS Lett,1993.319(3):p.271-6.
    114. Xiao, L. and W. Lang, A dominant role for the c-Jun NH2-terminal kinase in oncogenic ras-induced morphologic transformation of human lung carcinoma cells. Cancer Res, 2000.60(2):p.400-8.
    115. Liu, R., et al., Regulation of CSF1 promoter by the SWI/SNF-like BAF complex. Cell, 2001.106(3):p.309-18.
    116. Bedford, M.T., et al., Arginine methylation inhibits the binding of proline-rich ligands to Src homology 3, but not WW, domains. J Biol Chem,2000.275(21):p.16030-6.
    117. Brunet, A., et al., Stress-dependent regulation of FOXO transcription factors by the SIRT1 deacetylase. Science,2004.303(5666):p.2011-5.
    118. Zhang, X. and X. Cheng, Structure of the predominant protein arginine methyltransferase PRMT1 and analysis of its binding to substrate peptides. Structure,2003.11(5):p. 509-20.
    119. Wada, K., K. Inoue, and M. Hagiwara, Identification of methylated proteins by protein arginine N-methyltransferase 1, PRMT1, with a new expression cloning strategy. Biochim Biophys Acta,2002.1591(1-3):p.1-10.
    120. Yamagata, K., et al., Arginine methylation of FOXO transcription factors inhibits their phosphorylation byAkt. Mol Cell,2008.32(2):p.221-31.
    121. Arnold, H.H. and B. Winter, Muscle differentiation:more complexity to the network of myogenic regulators. Curr Opin Genet Dev,1998.8(5):p.539-44.
    122. Lassar, A.B., S.X. Skapek, and B. Novitch, Regulatory mechanisms that coordinate skeletal muscle differentiation and cell cycle withdrawal. Curr Opin Cell Biol,1994.6(6): p.788-94.
    123. Molkentin, J.D. and E.N. Olson, Defining the regulatory networks for muscle development. Curr Opin Genet Dev,1996.6(4):p.445-53.
    124. Yun, K. and B. Wold, Skeletal muscle determination and differentiation:story of a core regulatory network and its context. Curr Opin Cell Biol,1996.8(6):p.877-89.
    125. Xu, Q. and Z. Wu, The insulin-like growth factor-phosphatidylinositol 3-kinase-Akt signaling pathway regulates myogenin expression in normal myogenic cells but not in rhabdomyosarcoma-derived RD cells. J Biol Chem,2000.275(47):p.36750-7.
    126. Puri, P.L., et al., p300 is required for MyoD-dependent cell cycle arrest and muscle-specific gene transcription. EMBO J,1997.16(2):p.369-83.
    127. Sartorelli, V., et al., Molecular mechanisms of myogenic coactivation by p300:direct interaction with the activation domain of MyoD and with the MADS box of MEF2C. Mol Cell Biol,1997.17(2):p.1010-26.
    128. Yuan, W., et al., Human p300 protein is a coactivator for the transcription factor MyoD. J Biol Chem,1996.271(15):p.9009-13.
    129. Polesskaya, A., et al., CBP/p300 and muscle differentiation:no HAT, no muscle. EMBO J, 2001.20(23):p.6816-25.
    130. Solanes, G, et al., Functional relationship between MyoD and peroxisome proliferator-activated receptor-dependent regulatory pathways in the control of the human uncoupling protein-3 gene transcription. Mol Endocrinol,2003.17(10):p. 1944-58.
    131. Bachand, F., Protein arginine methyltransferases:from unicellular eukaryotes to humans. Eukaryot Cell,2007.6(6):p.889-98.
    132. Bedford, M.T., Arginine methylation at a glance. J Cell Sci,2007.120(Pt 24):p.4243-6.
    133. Bedford, M.T. and S. Richard, Arginine methylation an emerging regulator of protein function. Mol Cell,2005.18(3):p.263-72.
    134. Lee, D.Y., et al., Role of protein methylation in regulation of transcription. Endocr Rev, 2005.26(2):p.147-70.
    135. McBride, A.E. and P.A. Silver, State of the arg:protein methylation at arginine comes of age. Cell,2001.106(1):p.5-8.
    136. Wysocka, J., C.D. Allis, and S. Coonrod, Histone arginine methylation and its dynamic regulation. Front Biosci,2006.11:p.344-55.
    137. Shi, Y., et al., Histone demethylation mediated by the nuclear amine oxidase homolog LSD1. Cell,2004.119(7):p.941-53.
    138. Christensen, J., et al., RBP2 belongs to a family of demethylases, specific for tri-and dimethylated lysine 4 on histone 3. Cell,2007.128(6):p.1063-76.
    139. Cloos, P.A., et al., The putative oncogene GASC1 demethylates tri-and dimethylated lysine 9 on histone H3. Nature,2006.442(7100):p.307-11.
    140. Iwase, S., et al., The X-linked mental retardation gene SMCX/JAR1D1C defines a family of histone H3 lysine 4 demethylases. Cell,2007.128(6):p.1077-88.
    141. Klose, R.J., et al., The retinoblastoma binding protein RBP2 is an H3K4 demethylase. Cell,2007.128(5):p.889-900.
    142. Lee, M.G., et al., Demethylation of H3K27 regulates polycomb recruitment and H2A ubiquitination. Science,2007.318(5849):p.447-50.
    143. Whetstine, J.R., et al., Reversal of histone lysine trimethylation by the JMJD2 family of histone demethylases. Cell,2006.125(3):p.467-81.
    144. Yamane, K., et al., PLU-1 is an H3K4 demethylase involved in transcriptional repression and breast cancer cell proliferation. Mol Cell,2007.25(6):p.801-12.
    145. Yamane, K., et al., JHDM2A, a JmjC-containing H3K9 demethylase, facilitates transcription activation by androgen receptor. Cell,2006.125(3):p.483-95.
    146. Gu, W., et al., Interaction of myogenic factors and the retinoblastoma protein mediates muscle cell commitment and differentiation. Cell,1993.72(3):p.309-24.
    147. Cenciarelli, C., et al., Critical role played by cyclin D3 in the MyoD-mediated arrest of cell cycle during myoblast differentiation. Mol Cell Biol,1999.19(7):p.5203-17.
    148. Halevy, O., et al., Correlation of terminal cell cycle arrest of skeletal muscle with induction of p21 by MyoD. Science,1995.267(5200):p.1018-21.
    149. Franklin, D.S. and Y. Xiong, Induction of p18INK4c and its predominant association with CDK4 and CDK6 during myogenic differentiation. Mol Biol Cell,1996.7(10):p. 1587-99.
    150. Guo, K., et al., MyoD-induced expression of p21 inhibits cyclin-dependent kinase activity upon myocyte terminal differentiation. Mol Cell Biol,1995.15(7):p.3823-9.
    151. Missero, C., et al., The absence of p21Cipl/WAFl alters keratinocyte growth and differentiation and promotes ras-tumor progression. Genes Dev,1996.10(23):p. 3065-75.
    152. Parker, S.B., et al., p53-independent expression of p21Cipl in muscle and other terminally differentiating cells. Science,1995.267(5200):p.1024-7.
    153. Zhang, J.M., et al., Direct inhibition of G(1) cdk kinase activity by MyoD promotes myoblast cell cycle withdrawal and terminal differentiation. EMBO J,1999.18(24):p. 6983-93.
    154. Cheng, T.C., et al., Mapping of myogenin transcription during embryogenesis using transgenes linked to the myogenin control region. J Cell Biol,1992.119(6):p.1649-56.
    155. Yee, S.P. and P.W. Rigby, The regulation of myogenin gene expression during the embryonic development of the mouse. Genes Dev,1993.7(7A):p.1277-89.
    156. Buchberger, A., K. Ragge, and H.H. Arnold, The myogenin gene is activated during myocyte differentiation by pre-existing, not newly synthesized transcription factor MEF-2. J Biol Chem,1994.269(25):p.17289-96.
    157. Edmondson, D.G., et al., Analysis of the myogenin promoter reveals an indirect pathway for positive autoregulation mediated by the muscle-specific enhancer factor MEF-2. Mol Cell Biol,1992.12(9):p.3665-77.
    158. Johanson, M., et al., Transcriptional activation of the myogenin gene by MEF2-mediated recruitment of myf5 is inhibited by adenovirus E1A protein. Biochem Biophys Res Commun,1999.265(1):p.222-32.
    159. Lazazzera, B.A., D.M. Bates, and P.J. Kiley, The activity of the Escherichia coli transcription factor FNR is regulated by a change in oligomeric state. Genes Dev,1993. 7(10):p.1993-2005.
    160. Malik, S., C.F. Huang, and J. Schmidt, The role of the CANNTG promoter element (E box) and the myocyte-enhancer-binding-factor-2 (MEF-2) site in the transcriptional regulation of the chick myogenin gene. Eur J Biochem,1995.230(1):p.88-96.
    161. Spitz, F., et al., Expression of myogenin during embryogenesis is controlled by Six/sine oculis homeoproteins through a conserved MEF3 binding site. Proc Natl Acad Sci U S A, 1998.95(24):p.14220-5.
    162. Gerber, A.N., et al., Two domains of MyoD mediate transcriptional activation of genes in repressive chromatin:a mechanism for lineage determination in myogenesis. Genes Dev, 1997.11(4):p.436-50.
    163. Koh, S.S., et al., Synergistic enhancement of nuclear receptor function by p160 coactivators and two coactivators with protein methyltransferase activities. J Biol Chem, 2001.276(2):p.1089-98.
    164. Kleinschmidt, M.A., et al., The protein arginine methyltransferases CARM1 and PRMT1 cooperate in gene regulation. Nucleic Acids Res,2008.36(10):p.3202-13.
    165. Hassa, P.O., et al., Protein arginine methyltransferase 1 coactivates NF-kappaB-dependent gene expression synergistically with CARM1 and PARP1. J Mol Biol,2008.377(3):p.668-78.
    166. An, W., J. Kim, and R.G. Roeder, Ordered cooperative functions of PRMT1, p300, and CARM1 in transcriptional activation byp53. Cell,2004.117(6):p.735-48.
    167. Rezai-Zadeh, N., et al., Targeted recruitment of a histone H4-specific methyltransferase by the transcription factor YY1. Genes Dev,2003.17(8):p.1019-29.
    168. Zhao, X., et al., Methylation of RUNX1 by PRMT1 abrogates SIN3A binding and potentiates its transcriptional activity. Genes Dev,2008.22(5):p.640-53.
    169. Xie, Y., et al., Epigenetic regulation of transcriptional activity of pregnane X receptor by protein arginine methyltransferase 1. J Biol Chem,2009.284(14):p.9199-205.
    170. Li, Z.Y., et al., Sequential recruitment of PCAF and BRG1 contributes to myogenin activation in 12-O-tetradecanoylphorbol-l3-acetate-induced early differentiation of rhabdomyosarcoma-derived cells. J Biol Chem,2007.282(26):p.18872-8.
    171. Gao, X., et al., CARM1 activates myogenin gene via PCAF in the early differentiation of TPA-induced rhabdomyosarcoma-derived cells. J Cell Biochem.110(1):p.162-70.
    172. Pawlak, M.R., et al., Protein arginine methyltransferase I:substrate specificity and role in hnRNP assembly. J Cell Biochem,2002.87(4):p.394-407.
    173. Weintraub, H., et al., Activation of muscle-specific genes in pigment, nerve, fat, liver, and fibroblast cell lines by forced expression of MyoD. Proc Natl Acad Sci U S A,1989. 86(14):p.5434-8.
    174. Venuti, J.M. and P. Cserjesi, Molecular embryology of skeletal myogenesis. Curr Top Dev Biol,1996.34:p.169-206.
    175. Weintraub, H., et al., Muscle-specific transcriptional activation by MyoD. Genes Dev, 1991.5(8):p.1377-86.
    176. Weintraub, H., et al., MyoD binds cooperatively to two sites in a target enhancer sequence:occupancy of two sites is required for activation. Proc Natl Acad Sci U S A, 1990.87(15):p.5623-7.
    177. Knoepfler, P.S., et al., A conserved motif N-terminal to the DNA-binding domains of myogenic bHLH transcription factors mediates cooperative DNA binding with pbx-Meisl/Prepl. Nucleic Acids Res,1999.27(18):p.3752-61.
    178. Bergstrom, D.A. and S.J. Tapscott, Molecular distinction between specification and differentiation in the myogenic basic helix-loop-helix transcription factor family. Mol Cell Biol,2001.21(7):p.2404-12.
    1. Holtzer, H., et al., Effect of oncogenic virus on muscle differentiation. Proc Natl Acad Sci U S A,1975.72(10):p.4051-5.
    2. Holtzer, H., et al., Lineages, quantal cell cycles, and the generation of cell diversity. Q Rev Biophys,1975.8(4):p.523-57.
    3. Weintraub, H., G.L. Campbell, and H. Holtzer, Differentiation in the presence of bromodeoxyuridine id "all-or-none". Nat New Biol,1973.244(135):p.140-2.
    4. Lassar, A.B., B.M. Paterson, and H. Weintraub, Transfection of a DNA locus that mediates the conversion of 10T1/2 fibroblasts to myoblasts. Cell,1986.47(5):p.649-56.
    5. Taylor, S.M. and P.A. Jones, Multiple new phenotypes induced in 10T1/2 and 3T3 cells treated with 5-azacytidine. Cell,1979.17(4):p.771-9.
    6. Konieczny, S.F. and C.P. Emerson, Jr.,5-Azacytidine induction of stable mesodermal stem cell lineages from 10T1/2 cells:evidence for regulatory genes controlling determination. Cell,1984.38(3):p.791-800.
    7. Davis, R.L., H. Weintraub, and A.B. Lassar, Expression of a single transfected cDNA converts fibroblasts to myoblasts. Cell,1987.51(6):p.987-1000.
    8. Weintraub, H., et al., Activation of muscle-specific genes in pigment, nerve, fat, liver, and fibroblast cell lines by forced expression of MyoD. Proc Natl Acad Sci U S A,1989. 86(14):p.5434-8.
    9. Venuti, J.M. and P. Cserjesi, Molecular embryology of skeletal myogenesis. Curr Top Dev Biol,1996.34:p.169-206.
    10. Edmondson, D.G. and E.N. Olson, A gene with homology to the myc similarity region of MyoDl is expressed during myogenesis and is sufficient to activate the muscle differentiation program. Genes Dev,1989.3(5):p.628-40.
    11. Wright, W.E., D.A. Sassoon, and V.K. Lin, Myogenin, a factor regulating myogenesis, has a domain homologous to MyoD. Cell,1989.56(4):p.607-17.
    12. Braun, T., et al., A novel human muscle factor related to but distinct from MyoDl induces myogenic conversion in 10T1/2 fibroblasts. EMBO J,1989.8(3):p.701-9.
    13. Braun, T., et al., Myf-6, a new member of the human gene family of myogenic determination factors:evidence for a gene cluster on chromosome 12. EMBO J,1990. 9(3):p.821-31.
    14. Miner, J.H. and B. Wold, Herculin, a fourth member of the MyoD family of myogenic regulatory genes. Proc Natl Acad Sci U S A,1990.87(3):p.1089-93.
    15. Rhodes, S.J. and S.F. Konieczny, Identification of MRF4:a new member of the muscle regulatory factor gene family. Genes Dev,1989.3(12B):p.2050-61.
    16. Choi, J., et al., MyoD converts primary dermal fibroblasts, chondroblasts, smooth muscle, and retinal pigmented epithelial cells into striated mononucleated myoblasts and multinucleated myotubes. Proc Natl Acad Sci U S A,1990.87(20):p.7988-92.
    17. Ferre-D'Amare, A.R., et al., Recognition by Max of its cognate DNA through a dimeric b/HLH/Z domain. Nature,1993.363(6424):p.38-45.
    18. Ma, P.C., et al., Crystal structure of MyoD bHLH domain-DNA complex:perspectives on DNA recognition and implications for transcriptional activation. Cell,1994.77(3):p. 451-9.
    19. Lassar, A.B., et al., Functional activity of myogenic HLH proteins requires hetero-oligomerization with E12/E47-like proteins in vivo. Cell,1991.66(2):p.305-15.
    20. Murre, C., P.S. McCaw, and D. Baltimore, A new DNA binding and dimerization motif in immunoglobulin enhancer binding, daughterless, MyoD, and myc proteins. Cell,1989. 56(5):p.777-83.
    21. Brennan, T.J., T. Chakraborty, and E.N. Olson, Mutagenesis of the myogenin basic region identifies an ancient protein motif critical for activation of myogenesis. Proc Natl Acad Sci U S A,1991.88(13):p.5675-9.
    22. Davis, R.L., et al., The MyoD DNA binding domain contains a recognition code for muscle-specific gene activation. Cell,1990.60(5):p.733-46.
    23. Huang, J., H. Weintraub, and L. Kedes, Intramolecular regulation of MyoD activation domain conformation and function. Mol Cell Biol,1998.18(9):p.5478-84.
    24. Davis, R.L. and H. Weintraub, Acquisition of myogenic specificity by replacement of three amino acid residues from MyoD into E12. Science,1992.256(5059):p.1027-30.
    25. Weintraub, H., et al., Muscle-specific transcriptional activation by MyoD. Genes Dev, 1991.5(8):p.1377-86.
    26. Dezan, C., et al., Acquisition of myogenic specificity through replacement of one amino acid of MASH-1 and introduction of an additional alpha-helical turn. Biol Chem,1999. 380(6):p.705-10.
    27. Weintraub, H., et al., MyoD binds cooperatively to two sites in a target enhancer sequence:occupancy of two sites is required for activation. Proc Natl Acad Sci U S A, 1990.87(15):p.5623-7.
    28. Knoepfler, P.S., et al., A conserved motif N-terminal to the DNA-binding domains of myogenic bHLH transcription factors mediates cooperative DNA binding with pbx-Meisl/Prepl. Nucleic Acids Res,1999.27(18):p.3752-61.
    29. Gerber, A.N., et al., Two domains of MyoD mediate transcriptional activation of genes in repressive chromatin:a mechanism for lineage determination in myogenesis. Genes Dev, 1997.11(4):p.436-50.
    30. Bergstrom, D.A. and S.J. Tapscott, Molecular distinction between specification and differentiation in the myogenic basic helix-loop-helix transcription factor family. Mol Cell Biol,2001.21(7):p.2404-12.
    31. Zhang, J.M., et al., Direct inhibition of G(1) cdk kinase activity by MyoD promotes myoblast cell cycle withdrawal and terminal differentiation. EMBO J,1999.18(24):p. 6983-93.
    32. Chen, J.C. and D.J. Goldhamer, Transcriptional mechanisms regulating MyoD expression in the mouse. Cell Tissue Res,1999.296(1):p.213-9.
    33. Chen, J.C., C.M. Love, and D.J. Goldhamer, Two upstream enhancers collaborate to regulate the spatial patterning and timing of MyoD transcription during mouse development. Dev Dyn,2001.221(3):p.274-88.
    34. Tapscott, S.J., A.B. Lassar, and H. Weintraub, A novel myoblast enhancer element mediates MyoD transcription. Mol Cell Biol,1992.12(11):p.4994-5003.
    35. Goldhamer, D.J., et al., Regulatory elements that control the lineage-specific expression of myoD. Science,1992.256(5056):p.538-42.
    36. Kucharczuk, K.L., et al., Fine-scale transgenic mapping of the MyoD core enhancer: MyoD is regulated by distinct but overlapping mechanisms in myotomal and non-myotomal muscle lineages. Development,1999.126(9):p.1957-65.
    37. Faerman, A., et al., The distal human myoD enhancer sequences direct unique muscle-specific patterns of lacZ expression during mouse development. Dev Biol,1995. 171(1):p.27-38.
    38. Goldhamer, D.J., et al., Embryonic activation of the myoD gene is regulated by a highly conserved distal control element. Development,1995.121(3):p.637-49.
    39. Kablar, B., et al., MyoD and Myf-5 define the specification of musculature of distinct embryonic origin. Biochem Cell Biol,1998.76(6):p.1079-91.
    40. Asakura, A., G.E. Lyons, and S.J. Tapscott, The regulation of MyoD gene expression: conserved elements mediate expression in embryonic axial muscle. Dev Biol,1995. 171(2):p.386-98.
    41. Kablar, B., et al., Myogenic determination occurs independently in somites and limb buds. Dev Biol,1999.206(2):p.219-31.
    42. Kablar, B., et al., MyoD and Myf-5 differentially regulate the development of limb versus trunk skeletal muscle. Development,1997.124(23):p.4729-38.
    43. Vandromme, M., et al., Serum response factor p67SRF is expressed and required during myogenic differentiation of both mouse C2 and rat L6 muscle cell lines. J Cell Biol,1992. 118(6):p.1489-500.
    44. Gauthier-Rouviere, C., et al., Expression and activity of serum response factor is required for expression of the muscle-determining factor MyoD in both dividing and differentiating mouse C2C12 myoblasts. Mol Biol Cell,1996.7(5):p.719-29.
    45. Soulez, M., et al., Growth and differentiation of C2 myogenic cells are dependent on serum response factor. Mol Cell Biol,1996.16(11):p.6065-74.
    46. L'Honore, A., et al., MyoD distal regulatory region contains an SRF binding CArG element required for MyoD expression in skeletal myoblasts and during muscle regeneration. Mol Biol Cell,2003.14(5):p.2151-62.
    47. L'Honore, A., et al., Identification of a new hybrid serum response factor and myocyte enhancer factor 2-binding element in MyoD enhancer required for MyoD expression during myogenesis. Mol Biol Cell,2007.18(6):p.1992-2001.
    48. Iwasaki, K., et al., Rho/Rho-associated kinase signal regulates myogenic differentiation via myocardin-related transcription factor-A/Smad-dependent transcription of the Id3 gene. J Biol Chem,2008.283(30):p.21230-41.
    49. Carnac, G., et al., RhoA GTPase and serum response factor control selectively the expression of MyoD without affecting Myf5 in mouse myoblasts. Mol Biol Cell,1998. 9(7):p.1891-902.
    50. Takano, H., et al., The Rho family G proteins play a critical role in muscle differentiation. Mol Cell Biol,1998.18(3):p.1580-9.
    51. Wei, L., et al., RhoA signaling via serum response factor plays an obligatory role in myogenic differentiation. J Biol Chem,1998.273(46):p.30287-94.
    52. Lang, D., et al., PAX genes:roles in development, pathophysiology, and cancer. Biochem Pharmacol,2007.73(1):p.1-14.
    53. Buckingham, M. and F. Relaix, The role of Pax genes in the development of tissues and organs:Pax3 and Pax7 regulate muscle progenitor cell functions. Annu Rev Cell Dev Biol,2007.23:p.645-73.
    54. Kassar-Duchossoy, L., et al., Pax3/Pax7 mark a novel population of primitive myogenic cells during development. Genes Dev,2005.19(12):p.1426-31.
    55. Bober, E., et al., Pax-3 is required for the development of limb muscles:a possible role for the migration of dermomyotomal muscle progenitor cells. Development,1994.120(3): p.603-12.
    56. Goulding, M., A. Lumsden, and A.J. Paquette, Regulation of Pax-3 expression in the dermomyotome and its role in muscle development. Development,1994.120(4):p. 957-71.
    57. Maroto, M., et al., Ectopic Pax-3 activates MyoD and Myf-5 expression in embryonic mesoderm and neural tissue. Cell,1997.89(1):p.139-48.
    58. McKinnell, I.W., et al., Pax7 activates myogenic genes by recruitment of a histone methyltransferase complex. Nat Cell Biol,2008.10(1):p.77-84.
    59. Relaix, F., et al., A Pax3/Pax7-dependent population of skeletal muscle progenitor cells. Nature,2005.435(7044):p.948-53.
    60. Seale, P., et al., Pax7 is required for the specification of myogenic satellite cells. Cell, 2000.102(6):p.777-86.
    61. Accili, D. and K.C. Arden, FoxOs at the crossroads of cellular metabolism, differentiation, and transformation. Cell,2004.117(4):p.421-6.
    62. Bois, P.R. and G.C. Grosveld, FKHR (FOXO1a) is required for myotube fusion of primary mouse myoblasts. EMBO J,2003.22(5):p.1147-57.
    63. Hribal, M.L., et al., Regulation of insulin-like growth factor-dependent myoblast differentiation by Foxo forkhead transcription factors. J Cell Biol,2003.162(4):p. 535-41.
    64. Kitamura, T., et al., A Foxo/Notch pathway controls myogenic differentiation and fiber type specification. J Clin Invest,2007.117(9):p.2477-85.
    65. Machida, S., E.E. Spangenburg, and F.W. Booth, Forkhead transcription factor FoxO1 transduces insulin-like growth factor's signal to p27Kipl in primary skeletal muscle satellite cells. J Cell Physiol,2003.196(3):p.523-31.
    66. Mammucari, C., et al., FoxO3 controls autophagy in skeletal muscle in vivo. Cell Metab, 2007.6(6):p.458-71.
    67. Sandri, M., et al., Foxo transcription factors induce the atrophy-related ubiquitin ligase atrogin-1 and cause skeletal muscle atrophy. Cell,2004.117(3):p.399-412.
    68. Galili, N., et al., Fusion of a fork head domain gene to PAX3 in the solid tumour alveolar rhabdomyosarcoma. Nat Genet,1993.5(3):p.230-5.
    69. Mercado, G.E. and F.G. Barr, Fusions involving PAX and FOX genes in the molecular pathogenesis of alveolar rhabdomyosarcoma:recent advances. Curr Mol Med,2007. 7(1):p.47-61.
    70. Shapiro, D.N., et al., Fusion of PAX3 to a member of the forkhead family of transcription factors in human alveolar rhabdomyosarcoma. Cancer Res,1993.53(21):p.5108-12.
    71. Khan, J., et al., cDNA microarrays detect activation of a myogenic transcription program by the PAX3-FKHR fusion oncogene. Proc Natl Acad Sci U S A,1999.96(23):p. 13264-9.
    72. Hu, P., et al., Codependent activators direct myoblast-specific MyoD transcription. Dev Cell,2008.15(4):p.534-46.
    73. Roth, J.F., et al., Differential role of p300 and CBP acetyltransferase during myogenesis: p300 acts upstream of MyoD and Myf5. EMBO J,2003.22(19):p.5186-96.
    74. Eckner, R., et al., Interaction and functional collaboration of p300/CBP and bHLH proteins in muscle and B-cell differentiation. Genes Dev,1996.10(19):p.2478-90.
    75. Puri, P.L., et al., p300 is required for MyoD-dependent cell cycle arrest and muscle-specific gene transcription. EMBO J,1997.16(2):p.369-83.
    76. Sartorelli, V., et al., Molecular mechanisms of myogenic coactivation by p300:direct interaction with the activation domain of MyoD and with the MADS box of MEF2C. Mol Cell Biol,1997.17(2):p.1010-26.
    77. Yuan, W., et al., Human p300 protein is a coactivator for the transcription factor MyoD. J Biol Chem,1996.271(15):p.9009-13.
    78. Dilworth, F.J., et al., In vitro transcription system delineates the distinct roles of the coactivators pCAF and p300 during MyoD/E47-dependent transactivation. Proc Natl Acad Sci U S A,2004.101(32):p.11593-8.
    79. Polesskaya, A., et al., CREB-binding protein/p300 activates MyoD by acetylation. J Biol Chem,2000.275(44):p.34359-64.
    80. Polesskaya, A., et al., CBP/p300 and muscle differentiation:no HAT, no muscle. EMBO J, 2001.20(23):p.6816-25.
    81. Sartorelli, V., et al., Acetylation of MyoD directed by PCAF is necessary for the execution of the muscle program. Mol Cell,1999.4(5):p.725-34.
    82. Solanes, G., et al., Functional relationship between MyoD and peroxisome proliferator-activated receptor-dependent regulatory pathways in the control of the human uncoupling protein-3 gene transcription. Mol Endocrinol,2003.17(10):p. 1944-58.
    83. Zhang, C.L., T.A. McKinsey, and E.N. Olson, Association of class Ⅱ histone deacetylases with heterochromatin protein 1:potential role for histone methylation in control of muscle differentiation. Mol Cell Biol,2002.22(20):p.7302-12.
    84. Fulco, M., et al., Sir2 regulates skeletal muscle differentiation as a potential sensor of the redox state. Mol Cell,2003.12(1):p.51-62.
    85. Chen, S.L., et al., The coactivator-associated arginine methyltransferase is necessary for muscle differentiation:CARM1 coactivates myocyte enhancer factor-2. J Biol Chem, 2002.277(6):p.4324-33.
    86. Dacwag, C.S., et al., The protein arginine methyltransferase Prmt5 is required for myogenesis because it facilitates ATP-dependent chromatin remodeling. Mol Cell Biol, 2007.27(1):p.384-94.
    87. Iwasaki, H. and T. Yada, Protein arginine methylation regulates insulin signaling in L6 skeletal muscle cells. Biochem Biophys Res Commun,2007.364(4):p.1015-21.

© 2004-2018 中国地质图书馆版权所有 京ICP备05064691号 京公网安备11010802017129号

地址:北京市海淀区学院路29号 邮编:100083

电话:办公室:(+86 10)66554848;文献借阅、咨询服务、科技查新:66554700