Publikationsserver der Universitätsbibliothek Marburg

Titel:Die Testes spezifische Expression von Bromodomänen-Proteinen und Untereinheiten des Polycomb-Repressiven-Komplexes sowie deren mögliche Funktionen während der Spermatogenese in Mammalia
Autor:Gonzalez, Nicola Helena
Weitere Beteiligte: Rathke, Christina (Dr.)
Veröffentlicht:2015
URI:https://archiv.ub.uni-marburg.de/diss/z2015/0081
DOI: https://doi.org/10.17192/z2015.0081
URN: urn:nbn:de:hebis:04-z2015-00815
DDC: Biowissenschaften, Biologie
Titel (trans.):The testis-specific expression of bromodomain proteins and subunits of the Polycomb repressive complex and their possible functions during mammalian spermatogenesis
Publikationsdatum:2015-09-10
Lizenz:https://rightsstatements.org/vocab/InC-NC/1.0/

Dokument

Schlagwörter:
Spermatogenesis, Spermatogenese, Mammalia, Mammalia, Bromodomänen-Proteine, Polycomb-Repressiven-Komplexe ,, Polycomb Repressive Complex, Bromodomain Proteins

Zusammenfassung:
Die Spermatogenese ist ein sehr konservierter Prozess von der Fliege bis zu Mammalia, bei dem aus diploiden Spermatogonien haploide Spermatiden entstehen. Bei der Spermatogenese von Mammalia endet die Transkription in frühen, runden Spermatiden. Die für die Spermiogenese benötigten Transkripte werden translational reprimiert und selektiv bei Bedarf translatiert. In Drosophila melanogaster hingegen findet die Transkription translational reprimierter mRNAs bereits in den frühen Spermatozyten statt. Bezüglich des Drosophila-Testes gilt die Hypothese, dass u. a. ein testisspezifischer Transkriptionsfaktor TFIID existiert, der die Transkription der post-meiotisch relevanten Gene mit Hilfe von TATA-Box bindende Protein assoziierten Faktoren (tTAFs) initiiert. Weiterhin wird postuliert, dass der Polycomb-Repressive-Komplex 1 (PRC1) die Expression der Spermiogenese-relevanten Gene bis hin zu den frühen Spermatozyten reprimiert. Mit Unterstützung der tTAFs soll der PRC1 vom Promotor dieser Gene entfernt und zum Nukleolus der Spermatozyten rekrutiert werden, wobei die tTAFs zusätzlich auch an der Transkriptionsaktivierung dieser Gene beteiligt sein sollen. Zusätzlich konnten testisspezifisch exprimierte Bromodomänen-Proteine (tBRD-1, tBRD-2, tBRD-3) identifiziert werden, die in Drosophila-Testes mit Untereinheiten des PRC1 und den tTAFs in Spermatozyten zu kolokalisieren vermögen. Basierend auf diesen Daten ist die Vermutung aufgestellt worden, dass tBRDs an der Transkriptionsaktivierung in Spermatozyten beteiligt sind. Ob die Reprimierenden Komplexe PRC1 und PRC2 auch in Säuger-Testes vorkommen, ist nicht bekannt. In Mammalia ist BRDT das einzig bekannte BRD-Protein, dem eine Rolle bei der Transkriptionsinitiation als auch bei der Reorganisation des Chromatins im Verlauf der Spermatogenese nachgesagt wird. Im ersten Teil dieser Arbeit wurden klassische BRD-Proteine in Testes von Mammalia evaluiert. Die Resultate verweisen darauf, dass die Transkripte für die BRD-Proteine BRD2, BRD8 und TIF1α in frühen runden Spermatiden signifikant angereichert sind. Zudem überlappt die Expression diese Proteine mit dem Auftreten aktiver RNA-Polymerase II. Die Expression der beiden BRD-Proteine BRD3 und SMARCA2 überlappt mit der des Transitionsproteins 1 sowie mit hyperacetyliertem Histon H4. Die experimentellen Studien legen nahe, dass die genannten BRD-Proteine neben BRDT möglicherweise an der Transkription meiotischer und post-meiotischer Gene (BRD2, BRD8, TIF1α) sowie an der Reorganisation des Chromatins innerhalb des Histon-Protamin-Austausches (BRD3, SMARCA2) beteiligt sind. Im zweiten Teil der Arbeit wurde untersucht, ob die klassischen Untereinheiten des PRC1 und PRC2 in Säuger-Testes (Maus und Mensch) existieren. Es konnte gezeigt werden, dass die Transkripte der PRC1-Untereinheiten (Ring1, Bmi1, Scmh1, Cbx2, Cbx8) und des PRC2 (Ezh2, Eed, Suz12) in murinen Testes synthetisiert werden. Die Daten der Transkriptverteilung in Spermatogonien, Spermatozyten und runden Spermatiden erlaubten zudem den Rückschluss, dass der murine PRC1 u. a. aus den Untereinheiten RING1, BMI1 und SCMH1 sowie der PRC2 u. a. aus EZH2, EED und SUZ12 aufgebaut sein könnte. Zudem führt die korrelierende Transkriptverteilung einiger PRC- und BRD-Transkripte zu der Annahme, dass die Neusynthese der PRC-Untereinheiten erforderlich sind für die Repression aller Spermatogenesegene im Verlauf der Spermatidendifferenzierung.

Bibliographie / References

  1. Laemmli, U. K. (1970). "Cleavage of structural proteins during the assembly of the head of bacteriophage T4." Nature 227(5259): 680-685.
  2. Thomas, M. C. and C.-M. Chiang (2006). "The general transcription machinery and general cofactors." Critical reviews in biochemistry and molecular biology 41(3): 105-178.
  3. White-Cooper, H., M. Schafer, et al. (1998). "Transcriptional and post-transcriptional control mechanisms coordinate the onset of spermatid differentiation with meiosis I in Drosophila." Development 125(1): 125-134.
  4. Stock, J. K., S. Giadrossi, et al. (2007). "Ring1-mediated ubiquitination of H2A restrains poised RNA polymerase II at bivalent genes in mouse ES cells." Nature cell biology 9(12): 1428-1435.
  5. Clermont, Y. (1963). "The cycle of the seminiferous epithelium in man." Am J Anat 112: 35-51.
  6. Zhao, M., C. R. Shirley, et al. (2004a). "Transition nuclear proteins are required for normal chromatin condensation and functional sperm development." Genesis 38(4): 200-213.
  7. Dadoune, J. P. (2003). "Expression of mammalian spermatozoal nucleoproteins." Microscopy research and technique 61(1): 56-75.
  8. Unni, E., A. Mayerhofer, et al. (1995). "Increased accessibility of the N‐terminus of testis‐specific histone TH2B to antibodies in elongating spermatids." Molecular reproduction and development 42(2): 210- 219.
  9. Schafer, M., K. Nayernia, et al. (1995). "Translational control in spermatogenesis." Dev Biol 172(2): 344-352.
  10. Weinbauer, G. F., C. M. Luetjens, et al. (2010). Physiology of testicular function. Andrology, Springer: 11-59.
  11. Meistrich, M. L. and R. A. Hess (2013). "Assessment of spermatogenesis through staging of seminiferous tubules." Methods Mol Biol 927: 299-307.
  12. Mulisch, M. and U. Welsch (2010). Romeis -Mikroskopische Technik, Spektrum Akademischer Verlag.
  13. Meistrich, M. L., B. Mohapatra, et al. (2003). "Roles of transition nuclear proteins in spermiogenesis." Chromosoma 111(8): 483-488.
  14. Steger, K. (2001). "Haploid spermatids exhibit translationally repressed mRNAs." Anat Embryol (Berl) 203(5): 323-334.
  15. Kim, Y.-J., I. Hwang, et al. (1987). "Molecular cloning and differential expression of somatic and testis- specific H2B histone genes during rat spermatogenesis." Developmental biology 124(1): 23-34.
  16. Utakoji, T. (1966). "Chronology of nucleic acid synthesis in meiosis of the male Chinese hamster." Exp Cell Res 42(3): 585-596.
  17. Tanphaichitr, N., P. Sobhon, et al. (1978). "Basic nuclear proteins in testicular cells and ejaculated spermatozoa in man." Experimental Cell Research 117(2): 347-356.
  18. Rathke, C., W. M. Baarends, et al. (2014). "Chromatin dynamics during spermiogenesis." Biochim Biophys Acta 1839(3): 155-168.
  19. Shang, E., G. Salazar, et al. (2004). "Identification of unique, differentiation stage-specific patterns of expression of the bromodomain-containing genes Brd2, Brd3, Brd4, and Brdt in the mouse testis." Gene Expr Patterns 4(5): 513-519.
  20. Krebs, A. R., K. Karmodiya, et al. (2011). "SAGA and ATAC histone acetyl transferase complexes regulate distinct sets of genes and ATAC defines a class of p300-independent enhancers." Mol Cell 44(3): 410- 423.
  21. da Silva, S. M., A. Hacker, et al. (1996). "Sox9 expression during gonadal development implies a conserved role for the gene in testis differentiation in mammals and birds." Nature genetics 14(1): 62-68.
  22. Chi, T. (2004). "A BAF-centred view of the immune system." Nat Rev Immunol 4(12): 965-977.
  23. Talbert, P. B. and S. Henikoff (2010). "Histone variants—ancient wrap artists of the epigenome." Nature reviews Molecular cell biology 11(4): 264-275.
  24. Malik, H. S. and S. Henikoff (2003). "Phylogenomics of the nucleosome." Nature structural & molecular biology 10(11): 882-891.
  25. Turner, B. M. (2005). "Reading signals on the nucleosome with a new nomenclature for modified histones." Nature structural & molecular biology 12(2): 110-112.
  26. Kowalski, A. and J. Pałyga (2012). "Linker histone subtypes and their allelic variants." Cell biology international 36(11): 981-996.
  27. Sonnack, V., K. Failing, et al. (2002). "Expression of hyperacetylated histone H4 during normal and impaired human spermatogenesis." Andrologia 34(6): 384-390.
  28. Steilmann, C., A. Paradowska, et al. (2011). "Presence of histone H3 acetylated at lysine 9 in male germ cells and its distribution pattern in the genome of human spermatozoa." Reproduction, Fertility and Development 23(8): 997-1011.
  29. Larkin, M. A., G. Blackshields, et al. (2007). "Clustal W and Clustal X version 2.0." Bioinformatics 23(21): 2947-2948.
  30. Student (1908). "The probable error of a mean." Biometrika: 1-25.
  31. Magrane, M. and U. Consortium (2011). "UniProt Knowledgebase: a hub of integrated protein data." Database 2011: bar009.
  32. Van Roijen, H., M. P. Ooms, et al. (1998). "Immunoexpression of testis-specific histone 2B in human spermatozoa and testis tissue." Human Reproduction 13(6): 1559-1566.
  33. Renkawitz-Pohl, R., M. Hollmann, et al. (2005). "Spermatogenesis." Comprehensive molecular insect science 1: 157-178.
  34. Kibbe, W. A. (2007). "OligoCalc: an online oligonucleotide properties calculator." Nucleic acids research 35(suppl 2): W43-W46.
  35. Untergasser, A., I. Cutcutache, et al. (2012). "Primer3—new capabilities and interfaces." Nucleic acids research 40(15): e115-e115.
  36. Zhao, M., C. R. Shirley, et al. (2004b). "Nucleoprotein transitions during spermiogenesis in mice with transition nuclear protein Tnp1 and Tnp2 mutations." Biology of reproduction 71(3): 1016-1025.
  37. Leduc, F., V. Maquennehan, et al. (2008). "DNA damage response during chromatin remodeling in elongating spermatids of mice." Biology of reproduction 78(2): 324-332.
  38. Wang, J., H. Gu, et al. (2012). "Essential roles of the chromatin remodeling factor BRG1 in spermatogenesis in mice." Biol Reprod 86(6): 186. 91
  39. Kuzmichev, A., K. Nishioka, et al. (2002). "Histone methyltransferase activity associated with a human multiprotein complex containing the Enhancer of Zeste protein." Genes & development 16(22): 2893- 2905.
  40. Matangkasombut, O., R. M. Buratowski, et al. (2000). "Bromodomain factor 1 corresponds to a missing piece of yeast TFIID." Genes & development 14(8): 951-962.
  41. Deato, M. D. E. and R. Tjian (2007). "Switching of the core transcription machinery during myogenesis." Genes & development 21(17): 2137-2149.
  42. Reina, J. H. and N. Hernandez (2007). "On a roll for new TRF targets." Genes & development 21(22): 2855- 2860.
  43. Min, J., Y. Zhang, et al. (2003). "Structural basis for specific binding of Polycomb chromodomain to histone H3 methylated at Lys 27." Genes & development 17(15): 1823-1828.
  44. Chen, X., M. Hiller, et al. (2005). "Tissue-specific TAFs counteract Polycomb to turn on terminal differentiation." Science 310(5749): 869-872.
  45. Pivot-Pajot, C., C. Caron, et al. (2003). "Acetylation-dependent chromatin reorganization by BRDT, a testis- specific bromodomain-containing protein." Mol Cell Biol 23(15): 5354-5365.
  46. Leser, K., S. Awe, et al. (2012). "The bromodomain-containing protein tBRD-1 is specifically expressed in spermatocytes and is essential for male fertility." Biol Open 1(6): 597-606.
  47. De Vries, M., L. Ramos, et al. (2012). "Chromatin remodelling initiation during human spermiogenesis." Biol Open 1(5): 446-457.
  48. Shang, E., H. D. Nickerson, et al. (2007). "The first bromodomain of Brdt, a testis-specific member of the BET sub-family of double-bromodomain-containing proteins, is essential for male germ cell differentiation." Development 134(19): 3507-3515.
  49. Pointud, J. C., G. Mengus, et al. (2003). "The intracellular localisation of TAF7L, a paralogue of transcription factor TFIID subunit TAF7, is developmentally regulated during male germ-cell differentiation." J Cell Sci 116(Pt 9): 1847-1858.
  50. Lambrot, R., S. Jones, et al. (2012). "Specialized distribution of the histone methyltransferase Ezh2 in the nuclear apical region of round spermatids and its interaction with the histone variant H1t2." J Androl 33(5): 1058-1066.
  51. Russell, L. and B. Frank (1978). "Characterization of rat spermatocytes after plastic embedding." Systems Biology in Reproductive Medicine 1(1): 5-18.
  52. Reyes, J., J. Barra, et al. (1998). "Altered control of cellular proliferation in the absence of mammalian brahma (SNF2α)." The EMBO journal 17(23): 6979-6991.
  53. McPherson, S. and F. Longo (1992). "Chromatin structure-function alterations during mammalian spermatogenesis: DNA nicking and repair in elongating spermatids." European journal of histochemistry: EJH 37(2): 109-128.
  54. Christensen, J. H., B. Elfving, et al. (2012). "The Schizophrenia and Bipolar Disorder associated BRD1 gene is regulated upon chronic restraint stress." Eur Neuropsychopharmacol 22(9): 651-656.
  55. Satijn, D., M. J. Gunster, et al. (1997). "RING1 is associated with the polycomb group protein complex and acts as a transcriptional repressor." Molecular and cellular biology 17(7): 4105-4113.
  56. Scheuermann, J. C., L. Gutiérrez, et al. (2012). "Histone H2A monoubiquitination and Polycomb repression." Fly 6(3): 162-168.
  57. Van der Heijden, G., A. Derijck, et al. (2006). "Transmission of modified nucleosomes from the mouse male germline to the zygote and subsequent remodeling of paternal chromatin." Developmental biology 298(2): 458-469.
  58. Choudhary, C., C. Kumar, et al. (2009). "Lysine acetylation targets protein complexes and co-regulates major cellular functions." Science 325(5942): 834-840.
  59. Chomczynski, P. and N. Sacchi (2006). "The single-step method of RNA isolation by acid guanidinium thiocyanate-phenol-chloroform extraction: twenty-something years on." Nat Protoc 1(2): 581-585.
  60. Clarkson, M. J., J. R. Wells, et al. (1999). "Regions of variant histone His2AvD required for Drosophila development." Nature 399(6737): 694-697.
  61. Sambrook, J., E. F. Fritsch, et al. (1989). Molecular cloning: a laboratory manual, Cold spring harbor laboratory press New York.
  62. Steger, K. (1999). "Transcriptional and translational regulation of gene expression in haploid spermatids." Anatomy and embryology 199(6): 471-487.
  63. Kouzarides, T. (2000). "Acetylation: a regulatory modification to rival phosphorylation?" The EMBO journal 19(6): 1176-1179.
  64. (B) Aktin (1:2000) wurde stets als positiv Kontrolle verwendet und konnte als Signal in der Höhe von ~42 kDa im gesamt Proteinextrakt von Testes aller Alter (7 dpp (7), 16 dpp (16), 25 dpp (25) und Adultem (12 Wochen)) detektiert werden.
  65. Kouzarides, T. (2007). "Chromatin modifications and their function." Cell 128(4): 693-705.
  66. Wochen (Adult) mit spezifischen Antikörper gegen AKTIN Abbildung 27: Coomassie-Färbung und Immunmarkierung eines 10 % SDS-Gels mit spezifischem Antiköper gegen AKTIN. (A) Coomassie-Färbung eines 10% SDS-Gels, mit je 50 µg gesamt Proteinextrakt aus 7 dpp, 16 dpp, 25 dpp und Adultem (12 Wochen) murinem Testis, nach der elektrophoretischen Übertragung auf eine Nitrozellulosemembran. Um die Effizienz des elektrophoretischen Transfers zu überprüfen wurde nach dem Blotten, das SDS-Gel für ca. 2 h in Coomassie-Färbelösung gefärbt und für weitere 2 -3 h mit Coomassie-Entfärber wieder entfärbt um die verbliebenen Proteinbanden auf dem SDS-Gel zu visualisieren. Zum einem kann so überprüft werden, dass wirklich Protein in dem aufgetragenen Proteinextrakt enthalten war, zum anderen kann visuell nachvollzogen werden wie viel des Proteinextraktes nach dem Blotten noch auf dem SDS-Gel verblieben ist.
  67. Czermin, B., R. Melfi, et al. (2002). " Drosophila Enhancer of Zeste/ESC Complexes Have a Histone H3 Methyltransferase Activity that Marks Chromosomal Polycomb Sites." Cell 111(2): 185-196.
  68. Schuettengruber, B., D. Chourrout, et al. (2007). "Genome regulation by polycomb and trithorax proteins." Cell 128(4): 735-745.
  69. Russell, L., R. Ettlin, et al. (1990). "Histopathology of the testis." Histological and histopathological evaluation of the testis: 210-266.
  70. Immunmarkierung an murinen Testes im Alter von 7 dpp, 16 dpp, 25 dpp und 12
  71. Trostle-Weige, P., M. Meistrich, et al. (1982). "Isolation and characterization of TH2A, a germ cell-specific variant of histone 2A in rat testis." Journal of Biological Chemistry 257(10): 5560-5567.
  72. Cooke, H. J. and P. T. Saunders (2002). "Mouse models of male infertility." Nat Rev Genet 3(10): 790-801.
  73. Saiki, R. K., D. H. Gelfand, et al. (1988). "Primer-directed enzymatic amplification of DNA with a thermostable DNA polymerase." Science 239(4839): 487-491.
  74. Mullis, K., F. Faloona, et al. (1986). "Specific enzymatic amplification of DNA in vitro: the polymerase chain reaction." Cold Spring Harb Symp Quant Biol 51 Pt 1: 263-273.
  75. Wu, S.-Y. and C.-M. Chiang (2007). "The double bromodomain-containing chromatin adaptor Brd4 and transcriptional regulation." Journal of Biological Chemistry 282(18): 13141-13145.
  76. Connolly, C. M. and R. E. Braun (2009). The Mammalian Reproductive Genetics Database. Biology of reproduction, Soc Study Reprod.
  77. Chen, H. Y., J.-M. Sun, et al. (1998). "Ubiquitination of histone H3 in elongating spermatids of rat testes." Journal of Biological Chemistry 273(21): 13165-13169.
  78. Monesi, V. (1965). "Differential rate of ribonucleic acid synthesis in the autosomes and sex chromosomes during male meiosis in the mouse." Chromosoma 17(1): 11-21.
  79. Metcalf, C. E. and D. A. Wassarman (2007). "Nucleolar colocalization of TAF1 and testis‐specific TAFs during Drosophila spermatogenesis." Developmental Dynamics 236(10): 2836-2843.
  80. McCullagh, P., T. Chaplin, et al. (1999). "The cloning, mapping and expression of a novel gene, BRL, related to the AF10 leukaemia gene." Oncogene 18(52): 7442-7452.
  81. Luger, K., A. W. Mäder, et al. (1997). "Crystal structure of the nucleosome core particle at 2.8 Å resolution." Nature 389(6648): 251-260.
  82. Theofel, I., M. Bartkuhn, et al. (2014). "tBRD-1 Selectively Controls Gene Activity in the Drosophila Testis and Interacts with Two New Members of the Bromodomain and Extra-Terminal (BET) Family." PLoS One 9(9): e108267.
  83. Tanaka, H., N. Iguchi, et al. (2005). "HANP1/H1T2, a novel histone H1-like protein involved in nuclear formation and sperm fertility." Molecular and cellular biology 25(16): 7107-7119.
  84. Lee, D.-H., N. Gershenzon, et al. (2005). "Functional characterization of core promoter elements: the downstream core element is recognized by TAF1." Molecular and cellular biology 25(21): 9674-9686.
  85. Dey, A., F. Chitsaz, et al. (2003). "The double bromodomain protein Brd4 binds to acetylated chromatin during interphase and mitosis." Proceedings of the National Academy of Sciences 100(15): 8758-8763.
  86. Kim, T. H., L. O. Barrera, et al. (2005). "A high-resolution map of active promoters in the human genome." Nature 436(7052): 876-880.
  87. Cheng, Y., M. G. Buffone, et al. (2007). "Abnormal sperm in mice lacking the Taf7l gene." Mol Cell Biol 27(7): 2582-2589.
  88. Kierszenbaum, A. and L. L. Tres (1975). "Structural and transcriptional features of the mouse spermatid genome." The Journal of cell biology 65(2): 258-270.
  89. Schultz, N., F. K. Hamra, et al. (2003). "A multitude of genes expressed solely in meiotic or postmeiotic spermatogenic cells offers a myriad of contraceptive targets." Proceedings of the National Academy of Sciences 100(21): 12201-12206.
  90. Yang, Z., N. He, et al. (2008). "Brd4 recruits P-TEFb to chromosomes at late mitosis to promote G1 gene expression and cell cycle progression." Molecular and cellular biology 28(3): 967-976.
  91. LeRoy, G., B. Rickards, et al. (2008). "The double bromodomain proteins Brd2 and Brd3 couple histone acetylation to transcription." Mol Cell 30(1): 51-60.
  92. Ullah, M., N. Pelletier, et al. (2008). "Molecular architecture of quartet MOZ/MORF histone acetyltransferase complexes." Mol Cell Biol 28(22): 6828-6843.
  93. Sikorski, T. W. and S. Buratowski (2009). "The basal initiation machinery: beyond the general transcription factors." Curr Opin Cell Biol 21(3): 344-351.
  94. Dey, A., A. Nishiyama, et al. (2009). "Brd4 marks select genes on mitotic chromatin and directs postmitotic transcription." Mol Biol Cell 20(23): 4899-4909.
  95. Poirier, G. L., K. L. Shires, et al. (2008). "Anterior thalamic lesions produce chronic and profuse transcriptional de-regulation in retrosplenial cortex: A model of retrosplenial hypoactivity and covert pathology." Thalamus Relat Syst 4(1): 59-77.
  96. Song, N., J. Liu, et al. (2011). "Immunohistochemical analysis of histone H3 modifications in germ cells during mouse spermatogenesis." Acta histochemica et cytochemica 44(4): 183.
  97. Tan, M., H. Luo, et al. (2011). "Identification of 67 histone marks and histone lysine crotonylation as a new type of histone modification." Cell 146(6): 1016-1028.
  98. Zhao, S., W. Xu, et al. (2010). "Regulation of cellular metabolism by protein lysine acetylation." Science 327(5968): 1000-1004.
  99. Simon, J. A. and R. E. Kingston (2013). "Occupying chromatin: Polycomb mechanisms for getting to genomic targets, stopping transcriptional traffic, and staying put." Molecular cell 49(5): 808-824.
  100. Margueron, R., G. Li, et al. (2008). "Ezh1 and Ezh2 maintain repressive chromatin through different mechanisms." Molecular cell 32(4): 503-518.
  101. Margueron, R. and D. Reinberg (2011). "The Polycomb complex PRC2 and its mark in life." Nature 469(7330): 343-349.
  102. Margueron, R., N. Justin, et al. (2009). "Role of the polycomb protein EED in the propagation of repressive histone marks." Nature 461(7265): 762-767.
  103. Towbin, H., T. Staehelin, et al. (1979). "Electrophoretic transfer of proteins from polyacrylamide gels to nitrocellulose sheets: procedure and some applications." Proceedings of the National Academy of Sciences 76(9): 4350-4354.
  104. Marushige, K. (1976). "Activation of chromatin by acetylation of histone side chains." Proceedings of the National Academy of Sciences 73(11): 3937-3941.
  105. Martianov, I., S. Brancorsini, et al. (2005). "Polar nuclear localization of H1T2, a histone H1 variant, required for spermatid elongation and DNA condensation during spermiogenesis." Proceedings of the National Academy of Sciences of the United States of America 102(8): 2808-2813.
  106. McEntyre, J. and D. Lipman (2001). "PubMed: bridging the information gap." Canadian Medical Association Journal 164(9): 1317-1319.
  107. Dey, A., J. Ellenberg, et al. (2000). "A bromodomain protein, MCAP, associates with mitotic chromosomes and affects G2-to-M transition." Molecular and cellular biology 20(17): 6537-6549.
  108. Pasini, D., A. P. Bracken, et al. (2004). "Suz12 is essential for mouse development and for EZH2 histone methyltransferase activity." EMBO J 23(20): 4061-4071.
  109. Zhang, S., D. Li, et al. (2008). "Expression localization of Bmi1 in mice testis." Mol Cell Endocrinol 287(1-2): 47-56.
  110. Niederreither, K., E. Remboutsika, et al. (1999). "Expression of the transcriptional intermediary factor TIF1alpha during mouse development and in the reproductive organs." Mech Dev 88(1): 111-117.
  111. Morinière, J., S. Rousseaux, et al. (2009). "Cooperative binding of two acetylation marks on a histone tail by a single bromodomain." Nature 461(7264): 664-668.


* Das Dokument ist im Internet frei zugänglich - Hinweise zu den Nutzungsrechten