Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Review Article
  • Published:

Solid tumours hijack the histone variant network

Abstract

Cancer is a complex disease characterized by loss of cellular homeostasis through genetic and epigenetic alterations. Emerging evidence highlights a role for histone variants and their dedicated chaperones in cancer initiation and progression. Histone variants are involved in processes as diverse as maintenance of genome integrity, nuclear architecture and cell identity. On a molecular level, histone variants add a layer of complexity to the dynamic regulation of transcription, DNA replication and repair, and mitotic chromosome segregation. Because these functions are critical to ensure normal proliferation and maintenance of cellular fate, cancer cells are defined by their capacity to subvert them. Hijacking histone variants and their chaperones is emerging as a common means to disrupt homeostasis across a wide range of cancers, particularly solid tumours. Here we discuss histone variants and histone chaperones as tumour-promoting or tumour-suppressive players in the pathogenesis of cancer.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Fig. 1: Localization and deposition pathways of histone variants across the genome.
Fig. 2: Histone variant mutational spectrum and altered expression across solid tumours.
Fig. 3: Histone variant hallmarks in tumorigenesis.

Similar content being viewed by others

References

  1. Mendiratta, S., Gatto, A. & Almouzni, G. Histone supply: multitiered regulation ensures chromatin dynamics throughout the cell cycle. J. Cell Biol. 218, 39–54 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Buschbeck, M. & Hake, S. B. Variants of core histones and their roles in cell fate decisions, development and cancer. Nat. Rev. Mol. Cell Bio 18, 299–314 (2017).

    Article  CAS  Google Scholar 

  3. Talbert, P. B. & Henikoff, S. Histone variants on the move: substrates for chromatin dynamics. Nat. Rev. Mol. Cell Bio 18, 115–126 (2017).

    Article  CAS  Google Scholar 

  4. Mattiroli, F., D’Arcy, S. & Luger, K. The right place at the right time: chaperoning core histone variants. EMBO Rep. 16, 1454–1466 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Filipescu, D., Müller, S. & Almouzni, G. Histone H3 variants and their chaperones during development and disease: contributing to epigenetic control. Annu. Rev. Cell Dev. Biol. 30, 615–646 (2014).

    Article  CAS  PubMed  Google Scholar 

  6. Corujo, D. & Buschbeck, M. Post-translational modifications of H2A histone variants and their role in cancer. Cancers 10, 59 (2018).

    Article  PubMed Central  CAS  Google Scholar 

  7. Zhao, Z. & Shilatifard, A. Epigenetic modifications of histones in cancer. Genome Biol. 20, 245 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  8. Wang, T. et al. Histone variants: critical determinants in tumour heterogeneity. Front. Med. 13, 289–297 (2019).

    Article  PubMed  Google Scholar 

  9. Faast, R. et al. Histone variant H2A.Z is required for early mammalian development. Curr. Biol. 11, 1183–1187 (2001).

    Article  CAS  PubMed  Google Scholar 

  10. Eirín-López, J. M., González-Romero, R., Dryhurst, D., Ishibashi, T. & Ausió, J. The evolutionary differentiation of two histone H2A.Z variants in chordates (H2A.Z-1 and H2A.Z-2) is mediated by a stepwise mutation process that affects three amino acid residues. BMC Evol. Biol. 9, 31 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  11. Dryhurst, D. et al. Characterization of the histone H2A.Z-1 and H2A.Z-2 isoforms in vertebrates. BMC Biol. 7, 86 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  12. Horikoshi, N. et al. Structural polymorphism in the L1 loop regions of human H2A.Z.1 and H2A.Z.2. Acta Crystallogr. D Biol. Crystallogr. 69, 2431–2439 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Vardabasso, C. et al. Histone variant H2A.Z.2 mediates proliferation and drug sensitivity of malignant melanoma. Mol. Cell 59, 75–88 (2015). This article demonstrates that H2A.Z.2 together with BRD2 promotes expression of E2F target genes and proposes BETi as a therapeutic strategy for melanoma.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Dunn, C. J. et al. Histone hypervariants H2A.Z.1 and H2A.Z.2 play independent and context-specific roles in neuronal activity-induced transcription of Arc/Arg3.1 and other immediate early genes. eNeuro https://doi.org/10.1523/ENEURO.0040-17.2017 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  15. Greenberg, R. S., Long, H. K., Swigut, T. & Wysocka, J. Single amino acid change underlies distinct roles of H2A.Z subtypes in human syndrome. Cell 178, 1421–1436.e24 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Lamaa, A. et al. Integrated analysis of H2A.Z isoforms function reveals a complex interplay in gene regulation. eELife 9, e53375 (2020).

    Article  Google Scholar 

  17. Fan, J. Y., Gordon, F., Luger, K., Hansen, J. C. & Tremethick, D. J. The essential histone variant H2A.Z regulates the equilibrium between different chromatin conformational states. Nat. Struct. Biol. 9, 172–176 (2002).

    Article  CAS  PubMed  Google Scholar 

  18. Jin, C. & Felsenfeld, G. Nucleosome stability mediated by histone variants H3.3 and H2A.Z. Gene Dev. 21, 1519–1529 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Jin, C. et al. H3.3/H2A.Z double variant–containing nucleosomes mark “nucleosome-free regions” of active promoters and other regulatory regions. Nat. Genet. 41, 941–945 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Ku, M. et al. H2A.Z landscapes and dual modifications in pluripotent and multipotent stem cells underlie complex genome regulatory functions. Genome Biol. 13, R85 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Perell, G. T. et al. Specific acetylation patterns of H2A.Z form transient interactions with the BPTF bromodomain. Biochemistry 56, 4607–4615 (2017).

    Article  CAS  PubMed  Google Scholar 

  22. Giaimo, B. D., Ferrante, F., Herchenröther, A., Hake, S. B. & Borggrefe, T. The histone variant H2A.Z in gene regulation. Epigenet. Chromatin 12, 37 (2019).

    Article  Google Scholar 

  23. Dalvai, M. et al. H2A.Z-dependent crosstalk between enhancer and promoter regulates Cyclin D1 expression. Oncogene 32, 4243–4251 (2013).

    Article  CAS  PubMed  Google Scholar 

  24. Giaimo, B. D. et al. Histone variant H2A.Z deposition and acetylation directs the canonical Notch signaling response. Nucleic Acids Res. 46, gky551 (2018).

    Article  CAS  Google Scholar 

  25. Choi, J., Heo, K. & An, W. Cooperative action of TIP48 and TIP49 in H2A.Z exchange catalyzed by acetylation of nucleosomal H2A. Nucleic Acids Res. 37, 5993–6007 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Hsu, C.-C. et al. Gas41 links histone acetylation to H2A.Z deposition and maintenance of embryonic stem cell identity. Cell Discov. 4, 28 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  27. Cho, H. J. et al. GAS41 recognizes diacetylated histone H3 through a bivalent binding mode. ACS Chem. Biol. 13, 2739–2746 (2018).

    Article  CAS  PubMed  Google Scholar 

  28. Obri, A. et al. ANP32E is a histone chaperone that removes H2A.Z from chromatin. Nature 505, 648–653 (2014).

    Article  CAS  PubMed  Google Scholar 

  29. Alatwi, H. E. & Downs, J. A. Removal of H2A.Z by INO80 promotes homologous recombination. EMBO Rep. 16, 986–994 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Gursoy-Yuzugullu, O., Ayrapetov, M. K. & Price, B. D. Histone chaperone Anp32e removes H2A.Z from DNA double-strand breaks and promotes nucleosome reorganization and DNA repair. Proc. Natl Acad. Sci. USA 112, 7507–7512 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Greaves, I. K., Rangasamy, D., Ridgway, P. & Tremethick, D. J. H2A.Z contributes to the unique 3D structure of the centromere. Proc. Natl Acad. Sci. USA 104, 525–530 (2007).

    Article  CAS  PubMed  Google Scholar 

  32. Xu, Y. et al. Histone H2A.Z controls a critical chromatin remodeling step required for DNA double-strand break repair. Mol. Cell 48, 723–733 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Boyarchuk, E., Filipescu, D., Vassias, I., Cantaloube, S. & Almouzni, G. The histone variant composition of centromeres is controlled by the pericentric heterochromatin state during the cell cycle. J. Cell Sci. 127, 3347–3359 (2014).

    CAS  PubMed  Google Scholar 

  34. Creyghton, M. P. et al. H2AZ is enriched at polycomb complex target genes in ES cells and is necessary for lineage commitment. Cell 135, 649–661 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Subramanian, V. et al. H2A.Z acidic patch couples chromatin dynamics to regulation of gene expression programs during ESC differentiation. PLoS Genet. 9, e1003725 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Yang, H. D. et al. Oncogenic potential of histone-variant H2A.Z.1 and its regulatory role in cell cycle and epithelial-mesenchymal transition in liver cancer. Oncotarget 7, 11412–11423 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  37. Rispal, J. et al. The H2A.Z histone variant integrates Wnt signaling in intestinal epithelial homeostasis. Nat. Commun. 10, 1827 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  38. Frank, S. R. et al. MYC recruits the TIP60 histone acetyltransferase complex to chromatin. EMBO Rep. 4, 575–580 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Jeong, K. W. et al. Recognition of enhancer element–specific histone methylation by TIP60 in transcriptional activation. Nat. Struct. Mol. Biol. 18, 1358–1365 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Slupianek, A., Yerrum, S., Safadi, F. F. & Monroy, M. A. The chromatin remodeling factor SRCAP modulates expression of prostate specific antigen and cellular proliferation in prostate cancer cells. J. Cell Physiol. 224, 369–375 (2010).

    Article  CAS  PubMed  Google Scholar 

  41. Hua, S. et al. Genomic analysis of estrogen cascade reveals histone variant H2A.Z associated with breast cancer progression. Mol. Syst. Biol. 4, 188 (2008). This article identifies E-boxes at H2A.Z promoters and shows that oestrogen synergizes with MYC to induce H2A.Z expression.

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  42. Svotelis, A., Gévry, N., Grondin, G. & Gaudreau, L. H2A.Z overexpression promotes cellular proliferation of breast cancer cells. Cell Cycle 9, 364–370 (2014).

    Article  Google Scholar 

  43. Gévry, N. et al. Histone H2A.Z is essential for estrogen receptor signaling. Gene Dev. 23, 1522–1533 (2009).

    Article  PubMed  CAS  PubMed Central  Google Scholar 

  44. Tsai, C.-H. et al. SMYD3-mediated H2A.Z.1 methylation promotes cell cycle and cancer proliferation. Cancer Res. 76, 6043–6053 (2016).

    Article  CAS  PubMed  Google Scholar 

  45. He, H. H. et al. Nucleosome dynamics define transcriptional enhancers. Nat. Genet. 42, 343–347 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Valdés-Mora, F. et al. Acetylated histone variant H2A.Z is involved in the activation of neo-enhancers in prostate cancer. Nat. Commun. 8, 1346 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  47. Dryhurst, D., McMullen, B., Fazli, L., Rennie, P. S. & Ausió, J. Histone H2A.Z prepares the prostate specific antigen (PSA) gene for androgen receptor-mediated transcription and is upregulated in a model of prostate cancer progression. Cancer Lett. 315, 38–47 (2012).

    Article  CAS  PubMed  Google Scholar 

  48. Ito, S. et al. MRGBP promotes AR-mediated transactivation of KLK3 and TMPRSS2 via acetylation of histone H2A.Z in prostate cancer cells. Biochim. Biophys. Acta Gene Regul. Mech. 1861, 794–802 (2018).

    Article  CAS  Google Scholar 

  49. Valdés-Mora, F. et al. Acetylation of H2A.Z is a key epigenetic modification associated with gene deregulation and epigenetic remodeling in cancer. Genome Res. 22, 307–321 (2012).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  50. Sarcinella, E., Zuzarte, P. C., Lau, P. N. I., Draker, R. & Cheung, P. Monoubiquitylation of H2A.Z distinguishes its association with euchromatin or facultative heterochromatin. Mol. Cell Biol. 27, 6457–6468 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Draker, R., Sarcinella, E. & Cheung, P. USP10 deubiquitylates the histone variant H2A.Z and both are required for androgen receptor-mediated gene activation. Nucleic Acids Res. 39, 3529–3542 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Cerami, E. et al. The cBio Cancer Genomics Portal: an open platform for exploring multidimensional cancer genomics data. Cancer Discov. 2, 401–404 (2012).

    Article  PubMed  Google Scholar 

  53. Hsu, C.-C. et al. Recognition of histone acetylation by the GAS41 YEATS domain promotes H2A.Z deposition in non-small cell lung cancer. Gene Dev. 32, 58–69 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  54. Mattera, L. et al. The p400/Tip60 ratio is critical for colorectal cancer cell proliferation through DNA damage response pathways. Oncogene 28, 1506–1517 (2009).

    Article  CAS  PubMed  Google Scholar 

  55. Chevillard-Briet, M. et al. Interplay between chromatin-modifying enzymes controls colon cancer progression through Wnt signaling. Hum. Mol. Genet. 23, 2120–2131 (2014).

    Article  CAS  PubMed  Google Scholar 

  56. Draker, R. et al. A combination of H2A.Z and H4 acetylation eecruits Brd2 to chromatin during transcriptional activation. PLoS Genet. 8, e1003047 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. Filippakopoulos, P. et al. Selective inhibition of BET bromodomains. Nature 468, 1067–1073 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Yang, B. et al. H2A.Z regulates tumorigenesis, metastasis and sensitivity to cisplatin in intrahepatic cholangiocarcinoma. Int. J. Oncol. 52, 1235–1245 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Long, H. et al. H2A.Z facilitates licensing and activation of early replication origins. Nature 577, 1–18 (2020).

    Article  CAS  Google Scholar 

  60. Latrick, C. M. et al. Molecular basis and specificity of H2A.Z–H2B recognition and deposition by the histone chaperone YL1. Nat. Struct. Mol. Biol. 23, 309–316 (2016).

    Article  CAS  PubMed  Google Scholar 

  61. Liang, X. et al. Structural basis of H2A.Z recognition by SRCAP chromatin-remodeling subunit YL1. Nat. Struct. Mol. Biol. 23, 317–323 (2016).

    Article  CAS  PubMed  Google Scholar 

  62. Pehrson & Fried, V. MacroH2A, a core histone containing a large nonhistone region. Science 257, 1398–1400 (1992).

    Article  CAS  PubMed  Google Scholar 

  63. Chakravarthy, S. et al. Structural characterization of the histone variant macroH2A. Mol. Cell Biol. 25, 7616–7624 (2005).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Pehrson, J. R., Costanzi, C. & Dharia, C. Developmental and tissue expression patterns of histone macroH2A1 subtypes. J. Cell. Biochem. 65, 107–113 (1997).

    Article  CAS  PubMed  Google Scholar 

  65. Costanzi, C. & Pehrson, J. R. Histone macroH2A1 is concentrated in the inactive X chromosome of female mammals. Nature 393, 599–601 (1998).

    Article  CAS  PubMed  Google Scholar 

  66. Costanzi, C. & Pehrson, J. R. MACROH2A2, a new member of the MACROH2A core histone family. J. Biol. Chem. 276, 21776–21784 (2001).

    Article  CAS  PubMed  Google Scholar 

  67. Zhang, R. et al. Formation of macroH2A-containing senescence-associated heterochromatin foci and senescence driven by ASF1a and HIRA. Dev. Cell 8, 19–30 (2005).

    Article  CAS  PubMed  Google Scholar 

  68. Buschbeck, M. et al. The histone variant macroH2A is an epigenetic regulator of key developmental genes. Nat. Struct. Mol. Biol. 16, 1074–1079 (2009).

    Article  CAS  PubMed  Google Scholar 

  69. Changolkar, L. N. et al. Genome-wide distribution of macroH2A1 histone variants in mouse liver chromatin. Mol. Cell Biol. 30, 5473–5483 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Gamble, M. J., Frizzell, K. M., Yang, C., Krishnakumar, R. & Kraus, W. L. The histone variant macroH2A1 marks repressed autosomal chromatin, but protects a subset of its target genes from silencing. Gene Dev. 24, 21–32 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  71. Ratnakumar, K. et al. ATRX-mediated chromatin association of histone variant macroH2A1 regulates α-globin expression. Gene Dev. 26, 433–438 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Gaspar-Maia, A. et al. MacroH2A histone variants act as a barrier upon reprogramming towards pluripotency. Nat. Commun. 4, 1565 (2013).

    Article  PubMed  CAS  Google Scholar 

  73. Douet, J. et al. MacroH2A histone variants maintain nuclear organization and heterochromatin architecture. J. Cell Sci. 130, jcs.199216 (2017).

    Article  CAS  Google Scholar 

  74. Sun, Z. et al. Transcription-associated histone pruning demarcates macroH2A chromatin domains. Nat. Struct. Mol. Biol. 25, 958–970 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Kustatscher, G., Hothorn, M., Pugieux, C., Scheffzek, K. & Ladurner, A. G. Splicing regulates NAD metabolite binding to histone macroH2A. Nat. Struct. Mol. Biol. 12, 624–625 (2005).

    Article  CAS  PubMed  Google Scholar 

  76. Kozlowski, M. et al. MacroH2A histone variants limit chromatin plasticity through two distinct mechanisms. EMBO Rep. 19, e44445 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  77. Simonet, N. G. et al. SirT7 auto-ADP-ribosylation regulates glucose starvation response through mH2A1. Sci. Adv. 6, eaaz2590 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Ouararhni, K. et al. The histone variant mH2A1.1 interferes with transcription by down-regulating PARP-1 enzymatic activity. Gene Dev. 20, 3324–3336 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Timinszky, G. et al. A macrodomain-containing histone rearranges chromatin upon sensing PARP1 activation. Nat. Struct. Mol. Biol. 16, 923–929 (2009).

    Article  CAS  PubMed  Google Scholar 

  80. Xu, C., Xu, Y., Gursoy-Yuzugullu, O. & Price, B. D. The histone variant macroH2A1.1 is recruited to DSBs through a mechanism involving PARP1. Febs Lett. 586, 3920–3925 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Marjanović, M. P. et al. MacroH2A1.1 regulates mitochondrial respiration by limiting nuclear NAD+ consumption. Nat. Struct. Mol. Biol. 24, 902–910 (2017).

    Article  CAS  Google Scholar 

  82. Tong, L. & Denu, J. M. Function and metabolism of sirtuin metabolite O-acetyl-ADP-ribose. Biochim. Biophys. Acta Proteins Proteom. 1804, 1617–1625 (2010).

    Article  CAS  Google Scholar 

  83. Nusinow, D. A. et al. Poly(ADP-ribose) polymerase 1 is inhibited by a histone H2A variant, macroH2A, and contributes to silencing of the inactive X chromosome. J. Biol. Chem. 282, 12851–12859 (2007).

    Article  CAS  PubMed  Google Scholar 

  84. Pasque, V. et al. Histone variant macroH2A marks embryonic differentiation in vivo and acts as an epigenetic barrier to induced pluripotency. J. Cell Sci. 125, 6094–6104 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Pasque, V., Gillich, A., Garrett, N. & Gurdon, J. B. Histone variant macroH2A confers resistance to nuclear reprogramming. EMBO J. 30, 2373–2387 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  86. Barrero, M. J. et al. Macrohistone variants preserve cell identity by preventing the gain of H3K4me2 during reprogramming to pluripotency. Cell Rep. 3, 1005–1011 (2013).

    Article  CAS  PubMed  Google Scholar 

  87. Sporn, J. C. et al. Histone macroH2A isoforms predict the risk of lung cancer recurrence. Oncogene 28, 3423–3428 (2009).

    Article  CAS  PubMed  Google Scholar 

  88. Novikov, L. et al. QKI-mediated alternative splicing of the histone variant macroH2A1 regulates cancer cell proliferation. Mol. Cell Biol. 31, 4244–4255 (2011). This is the first article to identify a regulator of macroH2A1 alternative splicing, and a differential impact of splice variants on cell proliferation.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  89. Li, F. et al. QKI5-mediated alternative splicing of the histone variant macroH2A1 regulates gastric carcinogenesis. Oncotarget 7, 32821–32834 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  90. Vieira-Silva, T. S. et al. Histone variant macroH2A1 is downregulated in prostate cancer and influences malignant cell phenotype. Cancer Cell Int. 19, 112 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  91. Dardenne, E. et al. Splicing switch of an epigenetic regulator by RNA helicases promotes tumor-cell invasiveness. Nat. Struct. Mol. Biol. 19, 1139–1146 (2012). This study shows that splicing factors regulate macroH2A1 splice variant switching, contributing to breast cancer invasiveness.

    Article  CAS  PubMed  Google Scholar 

  92. Yip, B. H. et al. The U2AF1S34F mutation induces lineage-specific splicing alterations in myelodysplastic syndromes. J. Clin. Invest. 127, 2206–2221 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  93. Xu, D. et al. Skp2–MacroH2A1–CDK8 axis orchestrates G2/M transition and tumorigenesis. Nat. Commun. 6, 6641 (2015).

    Article  CAS  PubMed  Google Scholar 

  94. Hodge, D. Q., Cui, J., Gamble, M. J. & Guo, W. Histone variant macroH2A1 plays an isoform-specific role in suppressing epithelial-mesenchymal transition. Sci. Rep. 8, 841 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  95. Kim, J.-M. et al. MacroH2A1.2 inhibits prostate cancer-induced osteoclastogenesis through cooperation with HP1α and H1.2. Oncogene 37, 5749–5765 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Kim, J. et al. Regulation of breast cancer-induced osteoclastogenesis by macroH2A1.2 involving EZH2-mediated H3K27me3. Cell Rep. 24, 224–237 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  97. Kapoor, A. et al. The histone variant macroH2A suppresses melanoma progression through regulation of CDK8. Nature 468, 1105–1109 (2010). This is the first report of a functional role for macroH2A variants in cancer.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  98. Boulard, M. et al. Histone variant macroH2A1 deletion in mice causes female-specific steatosis. Epigenet. Chromatin 3, 8 (2010).

    Article  CAS  Google Scholar 

  99. Changolkar, L. N. et al. Developmental changes in histone macroH2A1-mediated gene regulation. Mol. Cell Biol. 27, 2758–2764 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Sheedfar, F. et al. Genetic ablation of macrohistone H2A1 leads to increased leanness, glucose tolerance and energy expenditure in mice fed a high-fat diet. Int. J. Obes. 39, 331–338 (2015).

    Article  CAS  Google Scholar 

  101. Pehrson, J. R., Changolkar, L. N., Costanzi, C. & Leu, N. A. Mice without macroH2A histone variants. Mol. Cell Biol. 34, 4523–4533 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  102. Rappa, F. et al. Immunopositivity for histone macroH2A1 isoforms marks steatosis-associated hepatocellular carcinoma. PLoS ONE 8, e54458 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Borghesan, M. et al. DNA hypomethylation and histone variant macroH2A1 synergistically attenuate chemotherapy-induced senescence to promote hepatocellular carcinoma progression. Cancer Res. 76, 594–606 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Re, O. L. et al. Histone variant macroH2A1 rewires carbohydrate and lipid metabolism of hepatocellular carcinoma cells towards cancer stem cells. Epigenetics 13, 829–845 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  105. Hu, W. H. et al. Loss of histone variant macroH2A2 expression associates with progression of anal neoplasm. J. Clin. Pathol. 69, 627 (2016).

    Article  CAS  PubMed  Google Scholar 

  106. Garcia, H. et al. Facilitates chromatin transcription complex is an “accelerator” of tumor transformation and potential marker and target of aggressive cancers. Cell Rep. 4, 159–173 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  107. Cedeno, R. J. et al. The histone variant macroH2A confers functional robustness to the intestinal stem cell compartment. PLoS ONE 12, e0185196 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  108. Curtin, N. J. & Szabo, C. Poly(ADP-ribose) polymerase inhibition: past, present and future. Nat. Rev. Drug Discov. 19, 711–736 (2020).

    Article  CAS  PubMed  Google Scholar 

  109. Wu, R. S., Tsai, S. & Bonner, W. M. Patterns of histone variant synthesis can distinguish go from G1 cells. Cell 31, 367–374 (1982).

    Article  CAS  PubMed  Google Scholar 

  110. Piña, B. & Suau, P. Changes in histones H2A and H3 variant composition in differentiating and mature rat brain cortical neurons. Dev. Biol. 123, 51–58 (1987).

    Article  PubMed  Google Scholar 

  111. Hake, S. B. et al. Expression patterns and post-translational modifications associated with mammalian histone H3 variants. J. Biol. Chem. 281, 559–568 (2006).

    Article  CAS  PubMed  Google Scholar 

  112. Goldberg, A. D. et al. Distinct factors control histone variant H3.3 localization at specific genomic regions. Cell 140, 678–691 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Ray-Gallet, D. et al. Dynamics of histone H3 deposition in vivo reveal a nucleosome gap-filling mechanism for H3.3 to maintain chromatin integrity. Mol. Cell 44, 928–941 (2011).

    Article  CAS  PubMed  Google Scholar 

  114. Zhang, H. et al. RPA interacts with HIRA and regulates H3.3 deposition at gene regulatory elements in mammalian cells. Mol. Cell 65, 272–284 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Adam, S., Polo, S. E. & Almouzni, G. Transcription recovery after DNA damage requires chromatin priming by the H3.3 histone chaperone HIRA. Cell 155, 94–106 (2013).

    Article  CAS  PubMed  Google Scholar 

  116. Banaszynski, L. A. et al. Hira-dependent histone H3.3 deposition facilitates PRC2 recruitment at developmental loci in ES cells. Cell 155, 107–120 (2013).

    Article  CAS  PubMed  Google Scholar 

  117. Filipescu, D., Szenker, E. & Almouzni, G. Developmental roles of histone H3 variants and their chaperones. Trends Genet. 29, 630–640 (2013).

    Article  CAS  PubMed  Google Scholar 

  118. Chen, P. et al. H3.3 actively marks enhancers and primes gene transcription via opening higher-ordered chromatin. Gene Dev. 27, 2109–2124 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  119. Armache, A. et al. Histone H3.3 phosphorylation amplifies stimulation-induced transcription. Nature 583, 852–857 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Lewis, P. W., Elsaesser, S. J., Noh, K.-M., Stadler, S. C. & Allis, C. D. Daxx is an H3.3-specific histone chaperone and cooperates with ATRX in replication-independent chromatin assembly at telomeres. Proc. Natl Acad. Sci. USA 107, 14075–14080 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Elsässer, S. J., Noh, K.-M., Diaz, N., Allis, C. D. & Banaszynski, L. A. Histone H3.3 is required for endogenous retroviral element silencing in embryonic stem cells. Nature 522, 240–244 (2015).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  122. Voon, H. P. J. et al. ATRX plays a key role in maintaining silencing at interstitial heterochromatic loci and imprinted genes. Cell Rep. 11, 405–418 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. He, Q. et al. The Daxx/Atrx complex protects tandem repetitive elements during DNA hypomethylation by promoting H3K9 trimethylation. Cell Stem Cell 17, 273–286 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. Udugama, M. et al. Histone variant H3.3 provides the heterochromatic H3 lysine 9 tri-methylation mark at telomeres. Nucleic Acids Res. 43, 10227–10237 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  125. Nikbakht, H. et al. Spatial and temporal homogeneity of driver mutations in diffuse intrinsic pontine glioma. Nat. Commun. 7, 11185 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Gomes, A. P. et al. Dynamic incorporation of histone H3 variants into chromatin is essential for acquisition of aggressive traits and metastatic colonization. Cancer Cell 36, 402–417.e13 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Schwartzentruber, J. et al. Driver mutations in histone H3.3 and chromatin remodelling genes in paediatric glioblastoma. Nature 482, 226–231 (2012). This study is among the first, along with Wu et al. (2012), to identify ‘oncohistone’ H3.3, ATRX and DAXX mutations in paediatric midline gliomas.

    Article  CAS  PubMed  Google Scholar 

  128. Khuong-Quang, D.-A. et al. K27M mutation in histone H3.3 defines clinically and biologically distinct subgroups of pediatric diffuse intrinsic pontine gliomas. Acta Neuropathol. 124, 439–447 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  129. Wu, G. et al. Somatic histone H3 alterations in pediatric diffuse intrinsic pontine gliomas and non-brainstem glioblastomas. Nat. Genet. 44, 251–253 (2012). This study is among the first, along with Schwartzentruber et al. (2012), to identify ‘oncohistone’ H3.3 and H3 mutations in paediatric midline gliomas.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  130. Sturm, D. et al. Hotspot mutations in H3F3A and IDH1 define distinct epigenetic and biological subgroups of glioblastoma. Cancer Cell 22, 425–437 (2012).

    Article  CAS  PubMed  Google Scholar 

  131. Fontebasso, A. M. et al. Recurrent somatic mutations in ACVR1 in pediatric midline high-grade astrocytoma. Nat. Genet. 46, 462–466 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  132. Aihara, K. et al. H3F3A K27M mutations in thalamic gliomas from young adult patients. Neuro Oncol. 16, 140–146 (2014).

    Article  CAS  PubMed  Google Scholar 

  133. Buczkowicz, P. et al. Genomic analysis of diffuse intrinsic pontine gliomas identifies three molecular subgroups and recurrent activating ACVR1 mutations. Nat. Genet. 46, 451–456 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  134. Bechet, D. et al. Specific detection of methionine 27 mutation in histone 3 variants (H3K27M) in fixed tissue from high-grade astrocytomas. Acta Neuropathol. 128, 733–741 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Wu, G. et al. The genomic landscape of diffuse intrinsic pontine glioma and pediatric non-brainstem high-grade glioma. Nat. Genet. 46, 444–450 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Taylor, K. R. et al. Recurrent activating ACVR1 mutations in diffuse intrinsic pontine glioma. Nat. Genet. 46, 457–461 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  137. Louis, D. N. et al. The 2016 World Health Organization classification of tumors of the central nervous system: a summary. Acta Neuropathol. 131, 803–820 (2016).

    Article  PubMed  Google Scholar 

  138. Parsons, D. W. et al. An integrated genomic analysis of human glioblastoma multiforme. Science 321, 1807–1812 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  139. Castel, D. et al. Histone H3F3A and HIST1H3B K27M mutations define two subgroups of diffuse intrinsic pontine gliomas with different prognosis and phenotypes. Acta Neuropathol. 130, 815–827 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Nacev, B. A. et al. The expanding landscape of ‘oncohistone’ mutations in human cancers. Nature 567, 473–478 (2019). This study analyses an extensive dataset across human malignancies and reveals that histone mutations are widespread.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Holoch, D. & Margueron, R. Mechanisms regulating PRC2 recruitment and enzymatic activity. Trends Biochem. Sci. 42, 531–542 (2017).

    Article  CAS  PubMed  Google Scholar 

  142. Bender, S. et al. Reduced H3K27me3 and DNA hypomethylation are major drivers of gene expression in K27M mutant pediatric high-grade gliomas. Cancer Cell 24, 660–672 (2013). This study identifies a dominant-negative effect of K27M on H3K27 trimethylation via PRC2 inhibition.

    Article  CAS  PubMed  Google Scholar 

  143. Lewis, P. W. et al. Inhibition of PRC2 activity by a gain-of-function H3 mutation found in pediatric glioblastoma. Science 340, 857–861 (2013). This article demonstrates a common mechanism by which H3 methionine for lysine substitutions inhibit SET domain-containing lysine methyltransferases.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  144. Chan, K.-M. et al. The histone H3.3K27M mutation in pediatric glioma reprograms H3K27 methylation and gene expression. Gene Dev. 27, 985–990 (2013). This article describes a global decrease and a local redistribution of H3K27me3 as a consequence of K27M with consequent gene expression changes.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  145. Venneti, S. et al. Evaluation of histone 3 lysine 27 trimethylation (H3K27me3) and enhancer of zest 2 (EZH2) in pediatric glial and glioneuronal tumors shows decreased H3K27me3 in H3F3A K27M mutant glioblastomas. Brain Pathol. 23, 558–564 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Justin, N. et al. Structural basis of oncogenic histone H3K27M inhibition of human polycomb repressive complex 2. Nat. Commun. 7, 11316 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Funato, K., Major, T., Lewis, P. W., Allis, C. D. & Tabar, V. Use of human embryonic stem cells to model pediatric gliomas with H3.3K27M histone mutation. Science 346, 1529–1533 (2014). This study models differentiation using the H3.3K27M mutation and proposes an epigenetic therapy for H3-mutant gliomas.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  148. Harutyunyan, A. S. et al. H3K27M induces defective chromatin spread of PRC2-mediated repressive H3K27me2/me3 and is essential for glioma tumorigenesis. Nat. Commun. 10, 1262 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  149. Fang, D. et al. H3.3K27M mutant proteins reprogram epigenome by sequestering the PRC2 complex to poised enhancers. eLife 7, e36696 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  150. Stafford, J. M. et al. Multiple modes of PRC2 inhibition elicit global chromatin alterations in H3K27M pediatric glioma. Sci. Adv. 4, eaau5935 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  151. Pathania, M. et al. H3.3K27M cooperates with Trp53 loss and PDGFRA gain in mouse embryonic neural progenitor cells to induce invasive high-grade gliomas. Cancer Cell 32, 684–700.e9 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  152. Cordero, F. J. et al. Histone H3.3K27M represses p16 to accelerate gliomagenesis in a murine model of DIPG. Mol. Cancer Res. 15, 1243–1254 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  153. Larson, J. D. et al. Histone H3.3 K27M accelerates spontaneous brainstem glioma and drives restricted changes in bivalent gene expression. Cancer Cell 35, 140–155.e7 (2019).

    Article  CAS  PubMed  Google Scholar 

  154. Silveira, A. B. et al. H3.3 K27M depletion increases differentiation and extends latency of diffuse intrinsic pontine glioma growth in vivo. Acta Neuropathol. 137, 637–655 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. Behjati, S. et al. Distinct H3F3A and H3F3B driver mutations define chondroblastoma and giant cell tumor of bone. Nat. Genet. 45, 1479–1482 (2013). This article identifies highly prevalent H3 mutations in rare cancer types affecting cartilage and growth plates of bones in children and young adults.

    Article  CAS  PubMed  Google Scholar 

  156. Schaefer, I. et al. Immunohistochemistry for histone H3G34W and H3K36M is highly specific for giant cell tumor of bone and chondroblastoma, respectively, in FNA and core needle biopsy. Cancer Cytopathol. 126, 552–566 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  157. Lu, C. et al. Histone H3K36 mutations promote sarcomagenesis through altered histone methylation landscape. Science 352, 844–849 (2016). This article characterizes the mechanism of action of K36M mutations, which affect H3K27me3 distribution and inhibit mesenchymal differentiation.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  158. Papillon-Cavanagh, S. et al. Impaired H3K36 methylation defines a subset of head and neck squamous cell carcinomas. Nat. Genet. 49, 180–185 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  159. Fang, D. et al. The histone H3.3K36M mutation reprograms the epigenome of chondroblastomas. Science 352, 1344–1348 (2016). This article characterizes the mechanism of action of K36M mutations by inhibiting H3K36 lysine methyltransferases.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  160. Schmitges, F. W. et al. Histone methylation by PRC2 is inhibited by active chromatin marks. Mol. Cell 42, 330–341 (2011).

    Article  CAS  PubMed  Google Scholar 

  161. Bjerke, L. et al. Histone H3.3 mutations drive pediatric glioblastoma through upregulation of MYCN. Cancer Discov. 3, 512–519 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  162. Yang, S. et al. Molecular basis for oncohistone H3 recognition by SETD2 methyltransferase. Gene Dev. 30, 1611–1616 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  163. Fang, J. et al. Cancer-driving H3G34V/R/D mutations block H3K36 methylation and H3K36me3–MutSα interaction. Proc. Natl Acad. Sci. USA 115, 201806355 (2018).

    Article  Google Scholar 

  164. Voon, H. P. J. et al. Inhibition of a K9/K36 demethylase by an H3.3 point mutation found in paediatric glioblastoma. Nat. Commun. 9, 3142 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  165. Fontebasso, A. M. et al. Mutations in SETD2 and genes affecting histone H3K36 methylation target hemispheric high-grade gliomas. Acta Neuropathol. 125, 659–669 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  166. Williams, M. J., Singleton, W. G. B., Lowis, S. P., Malik, K. & Kurian, K. M. Therapeutic targeting of histone modifications in adult and pediatric high-grade glioma. Front. Oncol. 7, 45 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  167. Xu, W. et al. Oncometabolite 2-hydroxyglutarate is a competitive inhibitor of α-ketoglutarate-dependent dioxygenases. Cancer Cell 19, 17–30 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  168. Chowdhury, R. et al. The oncometabolite 2-hydroxyglutarate inhibits histone lysine demethylases. EMBO Rep. 12, 463–469 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  169. Shi, L., Shi, J., Shi, X., Li, W. & Wen, H. Histone H3.3 G34 mutations alter histone H3K36 and H3K27 methylation in cis. J. Mol. Biol. 430, 1562–1565 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  170. Lim, J. et al. The histone variant H3.3 G34W substitution in giant cell tumor of the bone link chromatin and RNA processing. Sci. Rep. 7, 13459 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  171. Fellenberg, J. et al. Knock-down of oncohistone H3F3A-G34W counteracts the neoplastic phenotype of giant cell tumor of bone derived stromal cells. Cancer Lett. 448, 61–69 (2019).

    Article  CAS  PubMed  Google Scholar 

  172. Karremann, M. et al. Diffuse high-grade gliomas with H3 K27M mutations carry a dismal prognosis independent of tumor location. Neuro Oncol. 20, 123–131 (2018).

    Article  CAS  PubMed  Google Scholar 

  173. Huang, T. et al. Detection of histone H3 K27M mutation and post-translational modifications in pediatric diffuse midline glioma via tissue immunohistochemistry informs diagnosis and clinical outcomes. Oncotarget 9, 37112–37124 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  174. Hashizume, R. et al. Pharmacologic inhibition of histone demethylation as a therapy for pediatric brainstem glioma. Nat. Med. 20, 1394–1396 (2014). This is one of the first studies to propose an epigenetic therapy for H3-mutant gliomas.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  175. Yokoyama, A. et al. The menin tumor suppressor protein is an essential oncogenic cofactor for MLL-associated leukemogenesis. Cell 123, 207–218 (2005).

    Article  CAS  PubMed  Google Scholar 

  176. Mohammad, F. et al. EZH2 is a potential therapeutic target for H3K27M-mutant pediatric gliomas. Nat. Med. 23, 483–492 (2017).

    Article  CAS  PubMed  Google Scholar 

  177. Piunti, A. et al. Therapeutic targeting of polycomb and BET bromodomain proteins in diffuse intrinsic pontine gliomas. Nat. Med. 23, 493–500 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  178. Krug, B. et al. Pervasive H3K27 acetylation leads to ERV expression and a therapeutic vulnerability in H3K27M gliomas. Cancer Cell 35, 782–797.e8 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  179. Chheda, Z. S. et al. Novel and shared neoantigen derived from histone 3 variant H3.3K27M mutation for glioma T cell therapy. J. Exp. Med. 215, 141–157 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  180. Heaphy, C. M. et al. Altered Telomeres in Tumors with ATRX and DAXX Mutations. Science 333, 425–425 (2011). This article identifies a connection between ATRX/DAXX inactivation and acquisition of the ALT phenotype.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  181. Jiao, Y. et al. DAXX/ATRX, MEN1, and mTOR pathway genes are frequently altered in pancreatic neuroendocrine tumors. Science 331, 1199–1203 (2011). This study identifies ATRX/DAXX mutations in a poorly understood pancreatic cancer subtype.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  182. Cheung, N.-K. V. et al. Association of age at diagnosis and genetic mutations in patients with neuroblastoma. Jama 307, 1062–1071 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  183. Mahmud, I. & Liao, D. DAXX in cancer: phenomena, processes, mechanisms and regulation. Nucleic Acids Res. 47, 7734–7752 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  184. Kwan, P. S. et al. Daxx regulates mitotic progression and prostate cancer predisposition. Carcinogenesis 34, 750–759 (2013).

    Article  CAS  PubMed  Google Scholar 

  185. Tsourlakis, M. C. et al. Overexpression of the chromatin remodeler death-domain–associated protein in prostate cancer is an independent predictor of early prostate-specific antigen recurrence. Hum. Pathol. 44, 1789–1796 (2013).

    Article  CAS  PubMed  Google Scholar 

  186. Pan, W. W. et al. Death domain-associated protein DAXX promotes ovarian cancer development and chemoresistance. J. Biol. Chem. 288, 13620–13630 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  187. Dyer, M. A., Qadeer, Z. A., Valle-Garcia, D. & Bernstein, E. ATRX and DAXX: mechanisms and mutations. Cold Spring Harb. Perspect. Med. 7, a026567 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  188. Sarma, K. et al. ATRX directs binding of PRC2 to Xist RNA and polycomb targets. Cell 159, 869–883 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  189. Yost, K. E. et al. Rapid and reversible suppression of ALT by DAXX in osteosarcoma cells. Sci. Rep. 9, 4544 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  190. Killela, P. J. et al. TERT promoter mutations occur frequently in gliomas and a subset of tumors derived from cells with low rates of self-renewal. Proc. Natl Acad. Sci. USA 110, 6021–6026 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  191. Singhi, A. D. et al. Alternative lengthening of telomeres and loss of DAXX/ATRX expression predicts metastatic disease and poor survival in patients with pancreatic neuroendocrine tumors. Clin. Cancer Res. 23, 600–609 (2017).

    Article  CAS  PubMed  Google Scholar 

  192. Clynes, D. et al. Suppression of the alternative lengthening of telomere pathway by the chromatin remodelling factor ATRX. Nat. Commun. 6, 7538 (2015).

    Article  PubMed  Google Scholar 

  193. Law, M. J. et al. ATR-X syndrome protein targets tandem repeats and influences allele-specific expression in a size-dependent manner. Cell 143, 367–378 (2010).

    Article  CAS  PubMed  Google Scholar 

  194. Levy, M. A., Kernohan, K. D., Jiang, Y. & Bérubé, N. G. ATRX promotes gene expression by facilitating transcriptional elongation through guanine-rich coding regions. Hum. Mol. Genet. 24, 1824–1835 (2015).

    Article  CAS  PubMed  Google Scholar 

  195. O’Sullivan, R. J. et al. Rapid induction of alternative lengthening of telomeres by depletion of the histone chaperone ASF1. Nat. Struct. Mol. Biol. 21, 167–174 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  196. Marinoni, I. et al. Loss of DAXX and ATRX are associated with chromosome instability and reduced survival of patients with pancreatic neuroendocrine tumors. Gastroenterology 146, 453–460.e5 (2014).

    Article  CAS  PubMed  Google Scholar 

  197. Park, J. K. et al. DAXX/ATRX and MEN1 genes are strong prognostic markers in pancreatic neuroendocrine tumor. Oncotarget 5, 49796–49806 (2017).

    Article  Google Scholar 

  198. Chou, A. et al. ATRX loss is an independent predictor of poor survival in pancreatic neuroendocrine tumors. Hum. Pathol. 82, 249–257 (2018).

    Article  CAS  PubMed  Google Scholar 

  199. Chan, C. S. et al. ATRX, DAXX or MEN1 mutant pancreatic neuroendocrine tumors are a distinct alpha-cell signature subgroup. Nat. Commun. 9, 4158 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  200. Molenaar, J. J. et al. Sequencing of neuroblastoma identifies chromothripsis and defects in neuritogenesis genes. Nature 483, 589–593 (2012).

    Article  CAS  PubMed  Google Scholar 

  201. Qadeer, Z. A. et al. ATRX in-frame fusion neuroblastoma is sensitive to EZH2 inhibition via modulation of neuronal gene signatures. Cancer Cell 36, 512–527.e9 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  202. Pugh, T. J. et al. The genetic landscape of high-risk neuroblastoma. Nat. Genet. 45, 279–284 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  203. Kurihara, S., Hiyama, E., Onitake, Y., Yamaoka, E. & Hiyama, K. Clinical features of ATRX or DAXX mutated neuroblastoma. J. Pediatr. Surg. 49, 1835–1838 (2014).

    Article  PubMed  Google Scholar 

  204. Flynn, R. L. et al. Alternative lengthening of telomeres renders cancer cells hypersensitive to ATR inhibitors. Science 347, 273–277 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  205. Huh, M. S. et al. Stalled replication forks within heterochromatin require ATRX for protection. Cell Death Dis. 7, e2220 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  206. George, S. L. et al. Therapeutic vulnerabilities in the DNA damage response for the treatment of ATRX mutant neuroblastoma. Ebiomedicine 59, 102971 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  207. Sullivan, K. F., Hechenberger, M. & Masri, K. Human CENP-A contains a histone H3 related histone fold domain that is required for targeting to the centromere. J. Cell Biol. 127, 581–592 (1994).

    Article  CAS  PubMed  Google Scholar 

  208. Musacchio, A. & Desai, A. A molecular view of kinetochore assembly and function. Biology 6, 5 (2017).

    Article  PubMed Central  CAS  Google Scholar 

  209. Black, B. E. & Cleveland, D. W. Epigenetic centromere propagation and the nature of CENP-A nucleosomes. Cell 144, 471–479 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  210. Shelby, R. D., Vafa, O. & Sullivan, K. F. Assembly of CENP-A into centromeric chromatin requires a cooperative array of nucleosomal DNA contact sites. J. Cell Biol. 136, 501–513 (1997).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  211. Shelby, R. D., Monier, K. & Sullivan, K. F. Chromatin assembly at kinetochores is uncoupled from DNA replication. J. Cell Biol. 151, 1113–1118 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  212. Jansen, L. E. T., Black, B. E., Foltz, D. R. & Cleveland, D. W. Propagation of centromeric chromatin requires exit from mitosis. J. Cell Biol. 176, 795–805 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  213. Silva, M. C. C. et al. Cdk activity couples epigenetic centromere inheritance to cell cycle progression. Dev. Cell 22, 52–63 (2012).

    Article  CAS  PubMed  Google Scholar 

  214. Dunleavy, E. M. et al. HJURP is a cell-cycle-dependent maintenance and deposition factor of CENP-A at centromeres. Cell 137, 485–497 (2009).

    Article  CAS  PubMed  Google Scholar 

  215. Foltz, D. R. et al. Centromere-specific assembly of CENP-a nucleosomes is mediated by HJURP. Cell 137, 472–484 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  216. Hu, H. et al. Structure of a CENP-A–histone H4 heterodimer in complex with chaperone HJURP. Gene Dev. 25, 901–906 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  217. Stankovic, A. et al. A dual inhibitory mechanism sufficient to maintain cell-cycle-restricted CENP-A assembly. Mol. Cell 65, 231–246 (2017).

    Article  CAS  PubMed  Google Scholar 

  218. Howman, E. V. et al. Early disruption of centromeric chromatin organization in centromere protein A (Cenpa) null mice. Proc. Natl Acad. Sci. USA 97, 1148–1153 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  219. Régnier, V. et al. CENP-A is required for accurate chromosome segregation and sustained kinetochore association of BubR1. Mol. Cell Biol. 25, 3967–3981 (2005).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  220. Fachinetti, D. et al. A two-step mechanism for epigenetic specification of centromere identity and function. Nat. Cell Biol. 15, 1056–1066 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  221. Zhang, W. et al. Centromere and kinetochore gene misexpression predicts cancer patient survival and response to radiotherapy and chemotherapy. Nat. Commun. 7, 12619 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  222. Hu, Z. et al. The expression level of HJURP has an independent prognostic impact and predicts the sensitivity to radiotherapy in breast cancer. Breast Cancer Res. 12, R18 (2010).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  223. Li, Y. et al. ShRNA-targeted centromere protein a inhibits hepatocellular carcinoma growth. PLoS ONE 6, e17794 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  224. McGovern, S. L., Qi, Y., Pusztai, L., Symmans, W. F. & Buchholz, T. A. Centromere protein-A, an essential centromere protein, is a prognostic marker for relapse in estrogen receptor-positive breast cancer. Breast Cancer Res. 14, R72 (2012).

    Article  PubMed  PubMed Central  Google Scholar 

  225. Gu, X.-M. et al. Expression and prognostic relevance of centromere protein A in primary osteosarcoma. Pathol. Res. Pract. 210, 228–233 (2014).

    Article  CAS  PubMed  Google Scholar 

  226. Luo, J., Solimini, N. L. & Elledge, S. J. Principles of cancer therapy: oncogene and non-oncogene addiction. Cell 136, 823–837 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  227. Filipescu, D. et al. Essential role for centromeric factors following p53 loss and oncogenic transformation. Gene Dev. 31, 463–480 (2017). This work identifies a dependency on increased levels of CENP-A and HJURP for aberrant proliferation of p53-deficient cells.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  228. Müller, G. A. et al. The CHR promoter element controls cell cycle-dependent gene transcription and binds the DREAM and MMB complexes. Nucleic Acids Res. 40, 1561–1578 (2012).

    Article  PubMed  CAS  Google Scholar 

  229. Müller, G. A. et al. The CHR site: definition and genome-wide identification of a cell cycle transcriptional element. Nucleic Acids Res. 42, 10331–10350 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  230. Hooser, A. A. V. et al. Specification of kinetochore-forming chromatin by the histone H3 variant CENP-A. J. Cell Sci. 114, 3529–3542 (2001).

    Article  PubMed  Google Scholar 

  231. Lacoste, N. et al. Mislocalization of the centromeric histone variant CenH3/CENP-A in human cells depends on the chaperone DAXX. Mol. Cell 53, 631–644 (2014).

    Article  CAS  PubMed  Google Scholar 

  232. Athwal, R. K. et al. CENP-A nucleosomes localize to transcription factor hotspots and subtelomeric sites in human cancer cells. Epigenet Chromatin 8, 2 (2015).

    Article  CAS  Google Scholar 

  233. Nye, J., Sturgill, D., Athwal, R. & Dalal, Y. HJURP antagonizes CENP-A mislocalization driven by the H3.3 chaperones HIRA and DAXX. PLoS ONE 13, e0205948 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  234. Barra, V. & Fachinetti, D. The dark side of centromeres: types, causes and consequences of structural abnormalities implicating centromeric DNA. Nat. Commun. 9, 4340 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  235. Santaguida, S. & Amon, A. Short- and long-term effects of chromosome mis-segregation and aneuploidy. Nat. Rev. Mol. Cell Biol. 16, 473–485 (2015).

    Article  CAS  PubMed  Google Scholar 

  236. Zasadzińska, E. et al. Inheritance of CENP-A nucleosomes during DNA replication requires HJURP. Dev. Cell 47, 348–362.e7 (2018).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  237. Thompson, S. L. & Compton, D. A. Proliferation of aneuploid human cells is limited by a p53-dependent mechanismp53 suppresses aneuploidy. J. Cell Biol. 188, 369–381 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  238. Ohashi, A. et al. Aneuploidy generates proteotoxic stress and DNA damage concurrently with p53-mediated post-mitotic apoptosis in SAC-impaired cells. Nat. Commun. 6, 7668 (2015).

    Article  PubMed  Google Scholar 

  239. Santaguida, S. et al. Chromosome mis-segregation generates cell-cycle-arrested cells with complex karyotypes that are eliminated by the immune system. Dev. Cell 41, 638–651.e5 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  240. Galluzzi, L., Buqué, A., Kepp, O., Zitvogel, L. & Kroemer, G. Immunogenic cell death in cancer and infectious disease. Nat. Rev. Immunol. 17, 97–111 (2017).

    Article  CAS  PubMed  Google Scholar 

  241. Srivastava, S., Zasadzińska, E. & Foltz, D. R. Posttranslational mechanisms controlling centromere function and assembly. Curr. Opin. Cell Biol. 52, 126–135 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  242. Müller, S. et al. Phosphorylation and DNA binding of HJURP determine its centromeric recruitment and function in CenH3CENP-A loading. Cell Rep. 8, 190–203 (2014).

    Article  PubMed  CAS  Google Scholar 

  243. Yu, Z. et al. Dynamic phosphorylation of CENP-A at Ser68 orchestrates its cell-cycle-dependent deposition at centromeres. Dev. Cell 32, 68–81 (2015).

    Article  PubMed  CAS  Google Scholar 

  244. Pan, D. et al. CDK-regulated dimerization of M18BP1 on a Mis18 hexamer is necessary for CENP-A loading. eLife 6, e23352 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  245. Spiller, F. et al. Molecular basis for Cdk1-regulated timing of Mis18 complex assembly and CENP-A deposition. EMBO Rep. 18, 894–905 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  246. Takada, M. et al. FBW7 loss promotes chromosomal instability and tumorigenesis via cyclin E1/CDK2–mediated phosphorylation of CENP-A. Cancer Res. 77, 4881–4893 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  247. McKinley, K. L. & Cheeseman, I. M. Polo-like kinase 1 licenses CENP-A deposition at centromeres. Cell 158, 397–411 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  248. Niikura, Y. et al. CENP-A K124 ubiquitylation is required for CENP-A deposition at the centromere. Dev. Cell 32, 589–603 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  249. Smith, L. et al. The responses of cancer cells to PLK1 inhibitors reveal a novel protective role for p53 in maintaining centrosome separation. Sci. Rep. 7, 16115 (2017).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  250. Hanahan, D. & Weinberg, R. A. Hallmarks of cancer: the next generation. Cell 144, 646–674 (2011).

    Article  CAS  PubMed  Google Scholar 

  251. Atak, Z. K. et al. Comprehensive analysis of transcriptome variation uncovers known and novel driver events in T-cell acute lymphoblastic leukemia. PLoS Genet. 9, e1003997 (2013).

    Article  PubMed  CAS  Google Scholar 

  252. Huether, R. et al. The landscape of somatic mutations in epigenetic regulators across 1,000 paediatric cancer genomes. Nat. Commun. 5, 3630 (2014).

    Article  PubMed  CAS  Google Scholar 

  253. Lehnertz, B. et al. H3 K27M/I mutations promote context-dependent transformation in acute myeloid leukemia with RUNX1 alterations. Blood 130, 2204–2214 (2017).

    Article  CAS  PubMed  Google Scholar 

  254. Collord, G. et al. Recurrent histone mutations in T-cell acute lymphoblastic leukaemia. Brit J. Haematol. 184, 676–679 (2019).

    Article  Google Scholar 

  255. Boileau, M. et al. Mutant H3 histones drive human pre-leukemic hematopoietic stem cell expansion and promote leukemic aggressiveness. Nat. Commun. 10, 2891 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  256. Hussey, K. M. et al. The histone variant macroH2A1 regulates target gene expression in part by recruiting the transcriptional coregulator PELP1. Mol. Cell Biol. 34, 2437–2449 (2014).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  257. Chen, H. et al. MacroH2A1 and ATM play opposing roles in paracrine senescence and the senescence-associated secretory phenotype. Mol. Cell 59, 719–731 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  258. Rhodes, D. R. et al. Large-scale meta-analysis of cancer microarray data identifies common transcriptional profiles of neoplastic transformation and progression. Proc. Natl Acad. Sci. USA 101, 9309–9314 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  259. Hardy, S. et al. The euchromatic and heterochromatic landscapes are shaped by antagonizing effects of transcription on H2A.Z deposition. PLoS Genet. 5, e1000687 (2009).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  260. Segala, G., Bennesch, M. A., Pandey, D. P., Hulo, N. & Picard, D. Monoubiquitination of histone H2B blocks eviction of histone variant H2A.Z from inducible enhancers. Mol. Cell 64, 334–346 (2016).

    Article  CAS  PubMed  Google Scholar 

  261. Ma, X.-J. et al. Gene expression profiles of human breast cancer progression. Proc. Natl Acad. Sci. 100, 5974–5979 (2003).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  262. Oca, R. M. de et al. The histone chaperone HJURP is a new independent prognostic marker for luminal A breast carcinoma. Mol. Oncol. 9, 657–674 (2015).

    Article  CAS  Google Scholar 

  263. Lei, S., Long, J. & Li, J. MacroH2A suppresses the proliferation of the B16 melanoma cell line. Mol. Med. Rep. 10, 1845–1850 (2014).

    Article  CAS  PubMed  Google Scholar 

  264. Hu, B. et al. Holliday junction–recognizing protein promotes cell proliferation and correlates with unfavorable clinical outcome of hepatocellular carcinoma. Onco. Targets Ther. 10, 2601–2607 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  265. Wu, Q. et al. Expression and prognostic significance of centromere protein A in human lung adenocarcinoma. Lung Cancer 77, 407–414 (2012).

    Article  PubMed  Google Scholar 

  266. Kato, T. et al. Activation of Holliday junction–recognizing protein involved in the chromosomal stability and immortality of cancer cells. Cancer Res. 67, 8544–8553 (2007).

    Article  CAS  PubMed  Google Scholar 

  267. Barrios, O. de et al. ZEB1-induced tumourigenesis requires senescence inhibition via activation of DKK1/mutant p53/Mdm2/CtBP and repression of macroH2A1. Gut 66, 666 (2017).

    Article  PubMed  CAS  Google Scholar 

  268. Tomonaga, T. et al. Overexpression and mistargeting of centromere protein-A in human primary colorectal cancer. Cancer Res. 63, 3511–3516 (2003).

    CAS  PubMed  Google Scholar 

  269. Sang, Y. et al. CDK5-dependent phosphorylation and nuclear translocation of TRIM59 promotes macroH2A1 ubiquitination and tumorigenicity. Nat. Commun. 10, 4013 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  270. Stangeland, B. et al. Combined expressional analysis, bioinformatics and targeted proteomics identify new potential therapeutic targets in glioblastoma stem cells. Oncotarget 6, 26192–26215 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  271. Tayrac, M. de et al. Prognostic significance of EDN/RB, HJURP, p60/CAF-1 and PDLI4, four new markers in high-grade gliomas. PLoS ONE 8, e73332 (2013).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  272. Valente, V. et al. Modulation of HJURP (Holliday junction-recognizing protein) levels is correlated with glioblastoma cells survival. PLoS ONE 8, e62200 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  273. Qiu, J.-J. et al. Prognostic value of centromere protein-A expression in patients with epithelial ovarian cancer. Tumor Biol. 34, 2971–2975 (2013).

    Article  CAS  Google Scholar 

  274. Li, L., Li, X., Meng, Q., Khan, A. Q. & Chen, X. Increased expression of Holliday junction-recognizing protein (HJURP) as an independent prognostic biomarker in advanced-stage serous ovarian carcinoma. Med. Sci. Monit. 24, 3050–3055 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  275. Liu, Y., Hua, T., Chi, S. & Wang, H. Identification of key pathways and genes in endometrial cancer using bioinformatics analyses. Oncol. Lett. 17, 897–906 (2018).

    PubMed  PubMed Central  Google Scholar 

  276. Chen, X. et al. Recurrent somatic structural variations contribute to tumorigenesis in pediatric osteosarcoma. Cell Rep. 7, 104–112 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  277. Mason-Osann, E. et al. Identification of a novel gene fusion in ALT positive osteosarcoma. Oncotarget 9, 32868–32880 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  278. Lee, P. J. et al. Spectrum of mutations in leiomyosarcomas identified by clinical targeted next-generation sequencing. Exp. Mol. Pathol. 102, 156–161 (2017).

    Article  CAS  PubMed  Google Scholar 

  279. Mäkinen, N. et al. Exome sequencing of uterine leiomyosarcomas identifies frequent mutations in TP53, ATRX, and MED12. PLoS Genet. 12, e1005850 (2016).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  280. Liau, J.-Y. et al. Leiomyosarcoma with alternative lengthening of telomeres is associated with aggressive histologic features, loss of ATRX expression, and poor clinical outcome. Am. J. Surg. Pathol. 39, 236–244 (2015).

    Article  PubMed  Google Scholar 

  281. Liau, J.-Y. et al. Comprehensive screening of alternative lengthening of telomeres phenotype and loss of ATRX expression in sarcomas. Mod. Pathol. 28, 1545–1554 (2015).

    Article  CAS  PubMed  Google Scholar 

  282. Yang, C.-Y. et al. Targeted next-generation sequencing of cancer genes identified frequent TP53 and ATRX mutations in leiomyosarcoma. Am. J. Transl. Res. 7, 2072–2081 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  283. Tachiwana, H. et al. Structures of human nucleosomes containing major histone H3 variants. Acta Crystallogr. D Biol. Crystallogr. 67, 578–583 (2011).

    Article  CAS  PubMed  Google Scholar 

  284. Tachiwana, H. et al. Crystal structure of the human centromeric nucleosome containing CENP-A. Nature 476, 232–235 (2011).

    Article  CAS  PubMed  Google Scholar 

  285. Chakravarthy, S., Luger, K. PDB Entry - 2F8N. Worldwide Protein Data Bank. https://www.wwpdb.org/pdb?id=pdb_00002f8n (2011).

  286. Reynolds, C. R., Islam, S. A. & Sternberg, M. J. E. EzMol: a web server wizard for the rapid visualisation and image production of protein and nucleic acid structures. J. Mol. Biol. https://doi.org/10.1016/j.jmb.2018.01.013 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  287. Luger, K., Mäder, A. W., Richmond, R. K., Sargent, D. F. & Richmond, T. J. Crystal structure of the nucleosome core particle at 2.8 Å resolution. Nature 389, 251–260 (1997).

    Article  CAS  PubMed  Google Scholar 

  288. Tagami, H., Ray-Gallet, D., Almouzni, G. & Nakatani, Y. Histone H3.1 and H3.3 complexes mediate nucleosome assembly pathways dependent or independent of DNA synthesis. Cell 116, 51–61 (2004).

    Article  CAS  PubMed  Google Scholar 

  289. Latreille, D., Bluy, L., Benkirane, M. & Kiernan, R. E. Identification of histone 3 variant 2 interacting factors. Nucleic Acids Res. 42, 3542–3550 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  290. Shibahara, K. & Stillman, B. Replication-dependent marking of DNA by PCNA facilitates CAF-1-coupled inheritance of chromatin. Cell 96, 575–585 (1999).

    Article  CAS  PubMed  Google Scholar 

  291. Moggs, J. G. et al. A CAF-1–PCNA-mediated chromatin assembly pathway triggered by sensing DNA damage. Mol. Cell Biol. 20, 1206–1218 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  292. Mello, J. A. et al. Human Asf1 and CAF-1 interact and synergize in a repair-coupled nucleosome assembly pathway. EMBO Rep. 3, 329–334 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  293. Tang, Y. et al. Structure of a human ASF1a–HIRA complex and insights into specificity of histone chaperone complex assembly. Nat. Struct. Mol. Biol. 13, 921–929 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  294. Huang, H. et al. A unique binding mode enables MCM2 to chaperone histones H3–H4 at replication forks. Nat. Struct. Mol. Biol. 22, 618–626 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  295. Groth, A. et al. Regulation of replication fork progression through histone supply and demand. Science 318, 1928–1931 (2007).

    Article  CAS  PubMed  Google Scholar 

  296. Herranz, N. & Gil, J. Mechanisms and functions of cellular senescence. J. Clin. Invest. 128, 1238–1246 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  297. Chandra, T. et al. Independence of repressive histone marks and chromatin compaction during senescent heterochromatic layer formation. Mol. Cell 47, 203–214 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  298. Lee, K., Lau, Z. Z., Meredith, C. & Park, J. H. Decrease of p400 ATPase complex and loss of H2A.Z within the p21 promoter occur in senescent IMR-90 human fibroblasts. Mech. Ageing Dev. 133, 686–694 (2012).

    Article  CAS  PubMed  Google Scholar 

  299. Duarte, L. F. et al. Histone H3.3 and its proteolytically processed form drive a cellular senescence programme. Nat. Commun. 5, 5210 (2014). This study demonstrates a novel tumour-suppressive role for H3.3 proteolytic processing.

    Article  CAS  PubMed  Google Scholar 

  300. Chicas, A. et al. H3K4 demethylation by Jarid1a and Jarid1b contributes to retinoblastoma-mediated gene silencing during cellular senescence. Proc. Natl Acad. Sci.USA 109, 8971–8976 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  301. Corpet, A., Olbrich, T., Gwerder, M., Fink, D. & Stucki, M. Dynamics of histone H3.3 deposition in proliferating and senescent cells reveals a DAXX-dependent targeting to PML-NBs important for pericentromeric heterochromatin organization. Cell Cycle 13, 249–267 (2014).

    Article  CAS  PubMed  Google Scholar 

  302. Rai, T. S. et al. HIRA orchestrates a dynamic chromatin landscape in senescence and is required for suppression of neoplasia. Gene Dev. 28, 2712–2725 (2014). This article identifies a novel tumour-suppressive role for the histone chaperone HIRA.

    Article  PubMed  CAS  PubMed Central  Google Scholar 

Download references

Acknowledgements

The authors thank members of the Bernstein laboratory for discussions relevant to this Review. This work was supported by NIH/NCI grant R01CA154683, NIH/NCI grant R01CA218024, NIH/NINDS grant R01NS110837 (E.B.) and the American Skin Association Ping Y. Tai Foundation Research Grant in Skin Cancer/Melanoma (D.F.). The authors apologize to those whose work was not cited due to space constraints.

Author information

Authors and Affiliations

Authors

Contributions

D.F. and F.G.G. researched data for the article: All authors substantially contributed to discussion of the content, and wrote, reviewed and edited the manuscript before submission.

Corresponding author

Correspondence to Emily Bernstein.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information

Nature Reviews Cancer thanks M. Buschbeck and Z. Zhang for their contribution to the peer review of this work.

Publisher’s note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Related links

cBioPortal for Cancer Genomics: http://cbioportal.org

Supplementary information

Glossary

Oncohistones

Missense mutations in a histone variant, most commonly H3.3, often altering a key post-translational modification site that leads to a tumorigenic gene expression programme driving specific malignancies in children or young adults.

YEATS domain

Histone post-translational modification reader module found in transcriptional regulator proteins that selectively binds to acetylated and crotonylated lysine residues.

Pericentric heterochromatin

Constitutively condensed chromatin localized adjacent to the centromere, formed on repetitive DNA satellite sequences repressed by DNA methylation and histone H3 lysine 9 trimethylation (H3K9me3).

Bivalent chromatin domains

Genomic regions that harbour both the repressive mark histone H3 lysine 27 trimethylation (H3K27me3) and the activating mark histone H3 lysine 4 trimethylation (H3K4me3) and poise a subset of genes silenced in embryonic stem cells for activation during subsequent cell fate decisions to drive normal development.

PRC2

Polycomb repressive complex 2 (PRC2) catalyses trimethylation of histone H3 lysine 27 to silence developmental genes during differentiation, and contributes to facultative heterochromatin formation at the inactive X chromosome and imprinted genes.

PRC1

Polycomb repressive complex 1 (PRC1) monoubiquitylates histone H2A lysine 119 to induce gene repression, in coordination with histone H3 lysine 27 trimethylation (H3K27me3) deposited by PRC2.

Poly(ADP-ribose) polymerase

An enzyme that attaches chains of ADP-ribose from an NAD+ donor molecule to acceptor proteins (including poly(ADP-ribose) polymerases themselves) to regulate DNA repair and carbohydrate and lipid metabolism.

Triple-negative breast cancer

Breast cancer subtype that does not express oestrogen and progesterone receptors or HER2, and therefore is not suited to hormone therapy or HER2 inhibition.

Euchromatin

The loosely compacted, transcriptionally active and gene-enriched fraction of chromatin.

Alternative lengthening of telomeres

(ALT). Telomerase-independent mechanism that exploits homologous recombination of repetitive telomere sequences to promote telomere length maintenance in certain immortalized cell lines and tumours.

Promyelocytic leukaemia bodies

Membraneless nuclear compartments formed on a scaffold of promyelocytic leukaemia protein involved in telomere lengthening and DNA damage response.

ADD domain

Protein domain consisting of an amino-terminal GATA-like zinc-finger, a plant homeodomain finger and a long carboxy-terminal α-helix which binds histone H3 unmethylated at lysine 4 and trimethylated at lysine 9.

SNF2 helicase domain

Protein domain that uses the energy of ATP hydrolysis to apply torsional strain to DNA in order to remodel nucleosomes and other DNA–protein complexes.

G-quadruplex structures

Stable secondary DNA structures present in G-rich DNA held together by G-G base pairs formed within the same DNA strand.

ATM

Ataxia telangiectasia mutated (ATM) is a DNA damage kinase that phosphorylates itself and downstream effectors in response to DNA double-strand breaks to coordinate the cellular DNA damage response.

Kinetochore

Multicomplex protein structure localized at the centromere during mitosis to attach, orient and move sister chromatids along the mitotic spindle, ensuring accurate chromosome segregation.

Neocentromeres

Abnormal centromeres formed de novo at a location distinct from the primary constriction of a chromosome.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Ghiraldini, F.G., Filipescu, D. & Bernstein, E. Solid tumours hijack the histone variant network. Nat Rev Cancer 21, 257–275 (2021). https://doi.org/10.1038/s41568-020-00330-0

Download citation

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41568-020-00330-0

This article is cited by

Search

Quick links

Nature Briefing: Cancer

Sign up for the Nature Briefing: Cancer newsletter — what matters in cancer research, free to your inbox weekly.

Get what matters in cancer research, free to your inbox weekly. Sign up for Nature Briefing: Cancer