1932

Abstract

Many transcription factors (TFs) function as tumor suppressor genes with heterozygous phenotypes, yet haploinsufficiency generally has an underappreciated role in neoplasia. This is no less true in myeloid cells, which are normally regulated by a delicately balanced and interconnected transcriptional network. Detailed understanding of TF dose in this circuitry sheds light on the leukemic transcriptome. In this review, we discuss the emerging features of haploinsufficient transcription factors (HITFs). We posit that: () monoallelic and biallelic losses can have distinct cellular outcomes; () the activity of a TF exists in a greater range than the traditional Mendelian genetic doses; and () how a TF is deleted or mutated impacts the cellular phenotype. The net effect of a HITF is a myeloid differentiation block and increased intercellular heterogeneity in the course of myeloid neoplasia.

Loading

Article metrics loading...

/content/journals/10.1146/annurev-pathmechdis-051222-013421
2024-01-24
2024-04-29
Loading full text...

Full text loading...

/deliver/fulltext/pathol/19/1/annurev-pathmechdis-051222-013421.html?itemId=/content/journals/10.1146/annurev-pathmechdis-051222-013421&mimeType=html&fmt=ahah

Literature Cited

  1. 1.
    Barski G, Cornefert F. 1962. Characteristics of “hybrid”-type clonal cell lines obtained from mixed cultures in vitro. J. Natl. Cancer Inst. 28:4801–21
    [Google Scholar]
  2. 2.
    Harris H, Miller OJ, Klein G, Worst P, Tachibana T. 1969. Suppression of malignancy by cell fusion. Nature 223:5204363–68
    [Google Scholar]
  3. 3.
    Knudson AG. 1971. Mutation and cancer: statistical study of retinoblastoma. PNAS 68:4820–23
    [Google Scholar]
  4. 4.
    Fero ML, Randel E, Gurley KE, Roberts JM, Kemp CJ. 1998. The murine gene p27Kip1 is haplo-insufficient for tumour suppression. Nature 396:6707177–80
    [Google Scholar]
  5. 5.
    Bartha I, di Iulio J, Venter JC, Telenti A. 2017. Human gene essentiality. Nat. Rev. Genet. 19:151–62
    [Google Scholar]
  6. 6.
    Delneri D, Hoyle DC, Gkargkas K, Cross EJM, Rash B et al. 2007. Identification and characterization of high-flux-control genes of yeast through competition analyses in continuous cultures. Nat. Genet. 40:1113–17
    [Google Scholar]
  7. 7.
    Ohnuki S, Ohya Y. 2018. High-dimensional single-cell phenotyping reveals extensive haploinsufficiency. PLOS Biol. 16:5e2005130
    [Google Scholar]
  8. 8.
    White JK, Gerdin AK, Karp NA, Ryder E, Buljan M et al. 2013. Genome-wide generation and systematic phenotyping of knockout mice reveals new roles for many genes. Cell 154:2452–64
    [Google Scholar]
  9. 9.
    Karczewski KJ, Francioli LC, Tiao G, Cummings BB, Alföldi J et al. 2020. The mutational constraint spectrum quantified from variation in 141,456 humans. Nature 581:7809434–43This paper highlights the extent of genetic constraint in the human genome due to haploinsufficiency.
    [Google Scholar]
  10. 10.
    Ni Z, Zhou XY, Aslam S, Niu DK. 2019. Characterization of human dosage-sensitive transcription factor genes. Front. Genet. 10:1208
    [Google Scholar]
  11. 11.
    Collins RL, Glessner JT, Porcu E, Lepamets M, Brandon R et al. 2022. A cross-disorder dosage sensitivity map of the human genome. Cell 185:163041–55.e25
    [Google Scholar]
  12. 12.
    Davoli T, Xu AW, Mengwasser KE, Sack LM, Yoon JC et al. 2013. Cumulative haploinsufficiency and triplosensitivity drive aneuploidy patterns and shape the cancer genome. Cell 155:4948–62This paper demonstrates the great extent of haploinsufficiency in a pan-cancer analysis.
    [Google Scholar]
  13. 13.
    Chen Z, Trotman LC, Shaffer D, Lin HK, Dotan ZA et al. 2005. Crucial role of p53-dependent cellular senescence in suppression of Pten-deficient tumorigenesis. Nature 436:7051725–30
    [Google Scholar]
  14. 14.
    Berger AH, Knudson AG, Pandolfi PP. 2011. A continuum model for tumour suppression. Nature 476:7359163–69
    [Google Scholar]
  15. 15.
    Lee JY, Nakada D, Yilmaz OH, Tothova Z, Joseph NM et al. 2010. mTOR activation induces tumor suppressors that inhibit leukemogenesis and deplete hematopoietic stem cells after Pten deletion. Cell Stem Cell 7:5593–605
    [Google Scholar]
  16. 16.
    Alimonti A, Carracedo A, Clohessy JG, Trotman LC, Nardella C et al. 2010. Subtle variations in Pten dose determine cancer susceptibility. Nat. Genet. 42:5454–58
    [Google Scholar]
  17. 17.
    Will B, Vogler TO, Narayanagari S, Bartholdy B, Todorova TI et al. 2015. Minimal PU.1 reduction induces a preleukemic state and promotes development of acute myeloid leukemia. Nat. Med. 21:101172–81
    [Google Scholar]
  18. 18.
    Poluri RTK, Audet-Walsh É. 2018. Genomic deletion at 10q23 in prostate cancer: more than PTEN loss?. Front. Oncol. 8:246
    [Google Scholar]
  19. 19.
    Olson OC, Kang YA, Passegué E. 2020. Normal hematopoiesis is a balancing act of self-renewal and regeneration. Cold Spring Harb. Perspect. Med. 10:12a035519
    [Google Scholar]
  20. 20.
    Sanjuan-Pla A, Macaulay IC, Jensen CT, Woll PS, Luis TC et al. 2013. Platelet-biased stem cells reside at the apex of the haematopoietic stem-cell hierarchy. Nature 502:7470232–36
    [Google Scholar]
  21. 21.
    Bowling S, Sritharan D, Osorio FG, Nguyen M, Cheung P et al. 2020. An engineered CRISPR-Cas9 mouse line for simultaneous readout of lineage histories and gene expression profiles in single cells. Cell 181:61410–22.e27
    [Google Scholar]
  22. 22.
    Pei W, Shang F, Wang X, Fanti AK, Greco A et al. 2020. Resolving fates and single-cell transcriptomes of hematopoietic stem cell clones by PolyloxExpress barcoding. Cell Stem Cell 27:3383–95.e8
    [Google Scholar]
  23. 23.
    Rodriguez-Fraticelli AE, Wolock SL, Weinreb CS, Panero R, Patel SH et al. 2018. Clonal analysis of lineage fate in native haematopoiesis. Nature 553:7687212–16
    [Google Scholar]
  24. 24.
    Velten L, Haas SF, Raffel S, Blaszkiewicz S, Islam S et al. 2017. Human haematopoietic stem cell lineage commitment is a continuous process. Nat. Cell Biol. 19:4271–81
    [Google Scholar]
  25. 25.
    Lu F, Lionnet T. 2021. Transcription factor dynamics. Cold Spring Harb. Perspect. Biol. 13:11a040949
    [Google Scholar]
  26. 26.
    Aly M, Ramdzan ZM, Nagata Y, Balasubramanian SK, Hosono N et al. 2019. Distinct clinical and biological implications of CUX1 in myeloid neoplasms. Blood Adv. 3:142164–78
    [Google Scholar]
  27. 27.
    McNerney ME, Brown CD, Wang X, Bartom ET, Karmakar S et al. 2013. CUX1 is a haploinsufficient tumor suppressor gene on chromosome 7 frequently inactivated in acute myeloid leukemia. Blood 121:6975–83
    [Google Scholar]
  28. 28.
    Supper E, Rudat S, Iyer V, Droop A, Wong K et al. 2021. Cut-like homeobox1 (CUX1) tumor suppressor gene haploinsufficiency induces apoptosis evasion to sustain myeloid leukaemia. Nat. Commun. 12:2482
    [Google Scholar]
  29. 29.
    An N, Khan S, Imgruet MK, Gurbuxani SK, Konecki SN et al. 2018. Gene dosage effect of CUX1 in a murine model disrupts HSC homeostasis and controls the severity and mortality of MDS. Blood 131:242682–97
    [Google Scholar]
  30. 30.
    Larsson AJM, Johnsson P, Hagemann-Jensen M, Hartmanis L, Faridani OR et al. 2019. Genomic encoding of transcriptional burst kinetics. Nature 565:7738251–54
    [Google Scholar]
  31. 31.
    Wheat JC, Sella Y, Willcockson M, Skoultchi AI, Bergman A et al. 2020. Single-molecule imaging of transcription dynamics in somatic stem cells. Nature 583:7816431–36
    [Google Scholar]
  32. 32.
    Uhlén M, Fagerberg L, Hallström BM, Lindskog C, Oksvold P et al. 2015. Tissue-based map of the human proteome. Science 347:62201260419
    [Google Scholar]
  33. 33.
    Cook DL, Gerber AN, Tapscott SJ. 1998. Modeling stochastic gene expression: implications for haploinsufficiency. PNAS 95:2615641–46
    [Google Scholar]
  34. 34.
    Kemkemer R, Schrank S, Vogel W, Gruler H, Kaufmann D. 2002. Increased noise as an effect of haploinsufficiency of the tumor-suppressor gene neurofibromatosis type 1 in vitro. PNAS 99:2113783–88
    [Google Scholar]
  35. 35.
    Shaffer SM, Emert BL, Reyes Hueros RA, Cote C, Harmange G et al. 2020. Memory sequencing reveals heritable single-cell gene expression programs associated with distinct cellular behaviors. Cell 182:4947–59.e17
    [Google Scholar]
  36. 36.
    Ryl T, Kuchen EE, Bell E, Shao C, Flórez AF et al. 2017. Cell-cycle position of single MYC-driven cancer cells dictates their susceptibility to a chemotherapeutic drug. Cell Syst. 5:3237–50.e8
    [Google Scholar]
  37. 37.
    Waddington CH. 1957. The Strategy of the Genes: A Discussion of Some Aspects of Theoretical Biology London: Routledge, 1st ed.
  38. 38.
    Hansen AS, Amitai A, Cattoglio C, Tjian R, Darzacq X. 2019. Guided nuclear exploration increases CTCF target search efficiency. Nat. Chem. Biol. 16:3257–66
    [Google Scholar]
  39. 39.
    Izeddin I, Récamier V, Bosanac L, Cissé II, Boudarene L et al. 2014. Single-molecule tracking in live cells reveals distinct target-search strategies of transcription factors in the nucleus. eLife 3:e02230
    [Google Scholar]
  40. 40.
    Marklund E, van Oosten B, Mao G, Amselem E, Kipper K et al. 2020. DNA surface exploration and operator bypassing during target search. Nature 583:7818858–61
    [Google Scholar]
  41. 41.
    Chen J, Zhang Z, Li L, Chen BC, Revyakin A et al. 2014. Single-molecule dynamics of enhanceosome assembly in embryonic stem cells. Cell 156:61274–85
    [Google Scholar]
  42. 42.
    Stavreva DA, Garcia DA, Fettweis G, Gudla PR, Zaki GF et al. 2019. Transcriptional bursting and co-bursting regulation by steroid hormone release pattern and transcription factor mobility. Mol. Cell. 75:61161–77.e11
    [Google Scholar]
  43. 43.
    Purvis JE, Karhohs KW, Mock C, Batchelor E, Loewer A, Lahav G. 2012. p53 dynamics control cell fate. Science 336:60871440–44
    [Google Scholar]
  44. 44.
    Imayoshi I, Isomura A, Harima Y, Kawaguchi K, Kori H et al. 2013. Oscillatory control of factors determining multipotency and fate in mouse neural progenitors. Science 342:61631203–8
    [Google Scholar]
  45. 45.
    Kull T, Wehling A, Etzrodt M, Auler M, Dettinger P et al. 2022. NfκB signaling dynamics and their target genes differ between mouse blood cell types and induce distinct cell behavior. Blood 140:299–111
    [Google Scholar]
  46. 46.
    Natarajan A, Yardimci GG, Sheffield NC, Crawford GE, Ohler U. 2012. Predicting cell-type-specific gene expression from regions of open chromatin. Genome Res. 22:91711–22
    [Google Scholar]
  47. 47.
    Benabdallah NS, Williamson I, Illingworth RS, Kane L, Boyle S et al. 2019. Decreased enhancer-promoter proximity accompanying enhancer activation. Mol. Cell 76:3473–84.e7
    [Google Scholar]
  48. 48.
    Allahyar A, Vermeulen C, Bouwman BAM, Krijger PHL, Verstegen MJAM et al. 2018. Enhancer hubs and loop collisions identified from single-allele topologies. Nat. Genet. 50:81151–60
    [Google Scholar]
  49. 49.
    Schoenfelder S, Sexton T, Chakalova L, Cope NF, Horton A et al. 2009. Preferential associations between co-regulated genes reveal a transcriptional interactome in erythroid cells. Nat. Genet. 42:153–61
    [Google Scholar]
  50. 50.
    Li J, Dong A, Saydaminova K, Chang H, Wang G et al. 2019. Single-molecule nanoscopy elucidates RNA polymerase II transcription at single genes in live cells. Cell 178:2491–506.e28
    [Google Scholar]
  51. 51.
    Mir M, Stadler MR, Ortiz SA, Hannon CE, Harrison MM et al. 2018. Dynamic multifactor hubs interact transiently with sites of active transcription in drosophila embryos. eLife 7:e40497
    [Google Scholar]
  52. 52.
    Gu B, Swigut T, Spencley A, Bauer MR, Chung M et al. 2018. Transcription-coupled changes in nuclear mobility of mammalian cis-regulatory elements. Science 359:63791050–55
    [Google Scholar]
  53. 53.
    Petrovic J, Zhou Y, Fasolino M, Goldman N, Schwartz GW et al. 2019. Oncogenic Notch promotes long-range regulatory interactions within hyperconnected 3D cliques. Mol. Cell 73:61174–90.e12
    [Google Scholar]
  54. 54.
    Therizols P, Illingworth RS, Courilleau C, Boyle S, Wood AJ, Bickmore WA. 2014. Chromatin decondensation is sufficient to alter nuclear organization in embryonic stem cells. Science 346:62141238–42
    [Google Scholar]
  55. 55.
    Lundgren M, Chow CM, Sabbattini P, Georgiou A, Minaee S, Dillon N. 2000. Transcription factor dosage affects changes in higher order chromatin structure associated with activation of a heterochromatic gene. Cell 103:5733–43This paper demonstrates how monoallelic loss of a HITF can influence the nuclear position of target genes.
    [Google Scholar]
  56. 56.
    Liu J, Perumal NB, Oldfield CJ, Su EW, Uversky VN, Dunker AK. 2006. Intrinsic disorder in transcription factors. Biochemistry 45:226873–88
    [Google Scholar]
  57. 57.
    Henley MJ, Linhares BM, Morgan BS, Cierpicki T, Fierke CA, Mapp AK. 2020. Unexpected specificity within dynamic transcriptional protein-protein complexes. PNAS 117:4427346–53
    [Google Scholar]
  58. 58.
    Zou H, Pan T, Gao Y, Chen R, Li S et al. 2022. Pan-cancer assessment of mutational landscape in intrinsically disordered hotspots reveals potential driver genes. Nucleic Acids Res. 50:9e499
    [Google Scholar]
  59. 59.
    Brodsky S, Jana T, Mittelman K, Chapal M, Krishna Kumar D et al. 2020. Intrinsically disordered regions direct transcription factor in vivo binding specificity. Mol. Cell 79:459–71.e4
    [Google Scholar]
  60. 60.
    Ravarani CN, Erkina TY, de Baets G, Dudman DC, Erkine AM, Babu MM. 2018. High-throughput discovery of functional disordered regions: investigation of transactivation domains. Mol. Syst. Biol. 14:5e8190
    [Google Scholar]
  61. 61.
    Boija A, Klein IA, Sabari BR, Dall'Agnese A, Coffey EL et al. 2018. Transcription factors activate genes through the phase-separation capacity of their activation domains. Cell 175:71842–55.e16
    [Google Scholar]
  62. 62.
    Chong S, Dugast-Darzacq C, Liu Z, Dong P, Dailey GM et al. 2018. Imaging dynamic and selective low-complexity domain interactions that control gene transcription. Science 361:6400eaar2555
    [Google Scholar]
  63. 63.
    Sabari BR, Dall'Agnese A, Boija A, Klein IA, Coffey EL et al. 2018. Coactivator condensation at super-enhancers links phase separation and gene control. Science 361:6400eaar3958
    [Google Scholar]
  64. 64.
    Furlong EEM, Levine M. 2018. Developmental enhancers and chromosome topology. Science 361:64091341–45
    [Google Scholar]
  65. 65.
    Cho WK, Spille JH, Hecht M, Lee C, Li C et al. 2018. Mediator and RNA polymerase II clusters associate in transcription-dependent condensates. Science 361:6400412–15
    [Google Scholar]
  66. 66.
    Deviri D, Safran SA. 2021. Physical theory of biological noise buffering by multicomponent phase separation. PNAS 118:25e2100099118
    [Google Scholar]
  67. 67.
    Ng KP, Hu Z, Ebrahem Q, Negrotto S, Lausen J, Saunthararajah Y. 2013. Runx1 deficiency permits granulocyte lineage commitment but impairs subsequent maturation. Oncogenesis 2:11e78
    [Google Scholar]
  68. 68.
    Sinha T, Lammerts van Bueren K, Dickel DE, Zlatanova I, Thomas R et al. 2022. Differential Etv2 threshold requirement for endothelial and erythropoietic development. Cell Rep. 39:9110881
    [Google Scholar]
  69. 69.
    Giorgetti L, Siggers T, Tiana G, Caprara G, Notarbartolo S et al. 2010. Noncooperative interactions between transcription factors and clustered DNA binding sites enable graded transcriptional responses to environmental inputs. Mol. Cell 37:3418–28
    [Google Scholar]
  70. 70.
    Staber PB, Zhang P, Ye M, Welner RS, Nombela-Arrieta C et al. 2013. Sustained PU.1 levels balance cell-cycle regulators to prevent exhaustion of adult hematopoietic stem cells. Mol. Cell 49:5934–46
    [Google Scholar]
  71. 71.
    Naqvi S, Kim S, Hoskens H, Matthews HS, Spritz RA et al. 2023. Precise modulation of transcription factor levels identifies features underlying dosage sensitivity. Nat. Genet. 55:841–51This paper represents the most advanced analysis of the mechanism of a HITF using a novel degron system.
    [Google Scholar]
  72. 72.
    Staber PB, Zhang P, Ye M, Welner RS, Levantini E et al. 2014. The Runx-PU.1 pathway preserves normal and AML/ETO9a leukemic stem cells. Blood 124:152391–99
    [Google Scholar]
  73. 73.
    Ye M, Zhang H, Yang H, Koche R, Staber PB et al. 2015. Hematopoietic differentiation is required for initiation of acute myeloid leukemia. Cell Stem Cell 17:5611–23This paper demonstrates the requirement of myeloid cell identity for the development of myeloid neoplasm.
    [Google Scholar]
  74. 74.
    Huang N, Lee I, Marcotte EM, Hurles ME. 2010. Characterising and predicting haploinsufficiency in the human genome. PLOS Genet. 6:10e1001154
    [Google Scholar]
  75. 75.
    Assi SA, Imperato MR, Coleman DJL, Pickin A, Potluri S et al. 2018. Subtype-specific regulatory network rewiring in acute myeloid leukemia. Nat. Genet. 51:1151–62
    [Google Scholar]
  76. 76.
    Jotte MRM, McNerney ME. 2022. The significance of CUX1 and chromosome 7 in myeloid malignancies. Curr. Opin. Hematol. 29:292–102
    [Google Scholar]
  77. 77.
    Ben-David U, Amon A. 2019. Context is everything: aneuploidy in cancer. Nat. Rev. Genet. 21:144–62
    [Google Scholar]
  78. 78.
    Gao T, Ptashkin R, Bolton KL, Sirenko M, Fong C et al. 2021. Interplay between chromosomal alterations and gene mutations shapes the evolutionary trajectory of clonal hematopoiesis. Nat. Commun. 12:338
    [Google Scholar]
  79. 79.
    Chunduri NK, Menges P, Zhang X, Wieland A, Gotsmann VL et al. 2021. Systems approaches identify the consequences of monosomy in somatic human cells. Nat. Commun. 12:5576
    [Google Scholar]
  80. 80.
    Rücker FG, Schlenk RF, Bullinger L, Kayser S, Teleanu V et al. 2012. TP53 alterations in acute myeloid leukemia with complex karyotype correlate with specific copy number alterations, monosomal karyotype, and dismal outcome. Blood 119:92114–21
    [Google Scholar]
  81. 81.
    Beach RR, Ricci-Tam C, Brennan CM, Silberman RE, Springer M, Amon A. 2017. Aneuploidy causes non-genetic individuality. Cell 169:229–42
    [Google Scholar]
  82. 82.
    Farquhar KS, Charlebois DA, Szenk M, Cohen J, Nevozhay D, Balázsi G. 2019. Role of network-mediated stochasticity in mammalian drug resistance. Nat. Commun. 10:2766
    [Google Scholar]
  83. 83.
    Schukken KM, Sheltzer JM. 2022. Extensive protein dosage compensation in aneuploid human cancers. Genome Res. 32:71254–70
    [Google Scholar]
  84. 84.
    Liu Y, Chen C, Xu Z, Scuoppo C, Rillahan CD et al. 2016. Deletions linked to TP53 loss drive cancer through p53-independent mechanisms. Nature 531:7595471–75
    [Google Scholar]
  85. 85.
    Stoddart A, Nakitandwe J, Chen SC, Downing JR, Le Beau MM 2015. Haploinsufficient loss of multiple 5q genes may fine-tune Wnt signaling in del(5q) therapy-related myeloid neoplasms. Blood 126:262899–901
    [Google Scholar]
  86. 86.
    Joslin JM, Fernald AA, Tennant TR, Davis EM, Kogan SC et al. 2007. Haploinsufficiency of EGR1, a candidate gene in the del(5q), leads to the development of myeloid disorders. Blood 110:2719–26
    [Google Scholar]
  87. 87.
    Stoddart A, Fernald AA, Davis EM, McNerney ME, le Beau MM. 2022. EGR1 haploinsufficiency confers a fitness advantage to hematopoietic stem cells following chemotherapy. Exp Hematol. 115:54–67
    [Google Scholar]
  88. 88.
    Stoddart A, Fernald AA, Wang J, Davis EM, Karrison T et al. 2014. Haploinsufficiency of del(5q) genes, Egr1 and Apc, cooperate with Tp53 loss to induce acute myeloid leukemia in mice. Blood 123:71069–78
    [Google Scholar]
  89. 89.
    Pitsouli C, Perrimon N. 2013. The homeobox transcription factor cut coordinates patterning and growth during Drosophila airway remodeling. Sci. Signal. 6:263ra12
    [Google Scholar]
  90. 90.
    Grueber WB, Jan LY, Jan YN. 2003. Different levels of the homeodomain protein cut regulate distinct dendrite branching patterns of Drosophila multidendritic neurons. Cell 112:6805–18
    [Google Scholar]
  91. 91.
    Arthur RK, An N, Khan S, McNerney ME. 2017. The haploinsufficient tumor suppressor, CUX1, acts as an analog transcriptional regulator that controls target genes through distal enhancers that loop to target promoters. Nucleic Acids Res. 45:116350–61
    [Google Scholar]
  92. 92.
    Wong CC, Martincorena I, Rust AG, Rashid M, Alifrangis C et al. 2014. Inactivating CUX1 mutations promote tumorigenesis. Nat. Genet. 46:133–38
    [Google Scholar]
  93. 93.
    Schwartz JR, Ma J, Lamprecht T, Walsh M, Wang S et al. 2017. The genomic landscape of pediatric myelodysplastic syndromes. Nat. Commun. 8:1557
    [Google Scholar]
  94. 94.
    Smith SM, Le Beau MM, Huo D, Karrison T, Sobecks RM et al. 2003. Clinical-cytogenetic associations in 306 patients with therapy-related myelodysplasia and myeloid leukemia: the University of Chicago series. Blood 102:143–52
    [Google Scholar]
  95. 95.
    Imgruet MK, Lutze J, An NN, Hu B, Khan S et al. 2021. Loss of a 7q gene, CUX1, disrupts epigenetic-driven DNA repair and drives therapy-related myeloid neoplasms. Blood 138:790–805
    [Google Scholar]
  96. 96.
    Baker SJ, Ma'ayan A, Lieu YK, John P, Reddy MVR et al. 2014. B-myb is an essential regulator of hematopoietic stem cell and myeloid progenitor cell development. PNAS 111:83122–27
    [Google Scholar]
  97. 97.
    Clarke M, Dumon S, Ward C, Jäger R, Freeman S et al. 2012. MYBL2 haploinsufficiency increases susceptibility to age-related haematopoietic neoplasia. Leukemia 27:3661–70
    [Google Scholar]
  98. 98.
    Heinrichs S, Conover LF, Bueso-Ramos CE, Kilpivaara O, Stevenson K et al. 2013. MYBL2 is a sub-haploinsufficient tumor suppressor gene in myeloid malignancy. eLife 2:e00825
    [Google Scholar]
  99. 99.
    Braulke F, Müller-Thomas C, Götze K, Platzbecker U, Germing U et al. 2015. Frequency of del(12p) is commonly underestimated in myelodysplastic syndromes: results from a German diagnostic study in comparison with an international control group. Genes Chromosomes Cancer 54:12809–17
    [Google Scholar]
  100. 100.
    Feurstein S, Rücker FG, Bullinger L, Hofmann W, Manukjan G et al. 2014. Haploinsufficiency of ETV6 and CDKN1B in patients with acute myeloid leukemia and complex karyotype. BMC Genom. 151:784
    [Google Scholar]
  101. 101.
    Green SM, Coyne HJ, McIntosh LP, Graves BJ. 2010. DNA binding by the ETS protein TEL (ETV6) is regulated by autoinhibition and self-association. J. Biol. Chem. 285:2418496–504
    [Google Scholar]
  102. 102.
    Churpek JE, Bresnick EH. 2019. Transcription factor mutations as a cause of familial myeloid neoplasms. J. Clin. Investig. 129:2476–88
    [Google Scholar]
  103. 103.
    Fisher MH, Kirkpatrick GD, Stevens B, Jones C, Callaghan M et al. 2020. ETV6 germline mutations cause HDAC3/NCOR2 mislocalization and upregulation of interferon response genes. JCI Insight 5:18e140332
    [Google Scholar]
  104. 104.
    Bernard E, Nannya Y, Hasserjian RP, Devlin SM, Tuechler H et al. 2020. Implications of TP53 allelic state for genome stability, clinical presentation and outcomes in myelodysplastic syndromes. Nat. Med. 26:101549–56
    [Google Scholar]
  105. 105.
    Boettcher S, Miller PG, Sharma R, McConkey M, Leventhal M et al. 2019. A dominant-negative effect drives selection of TP53 missense mutations in myeloid malignancies. Science 365:6453599–604
    [Google Scholar]
  106. 106.
    Venkatachalam S, Shi YP, Jones SN, Vogel H, Bradley A et al. 1998. Retention of wild-type p53 in tumors from p53 heterozygous mice: Reduction of p53 dosage can promote cancer formation. EMBO J. 17:164657–67
    [Google Scholar]
  107. 107.
    Solimini NL, Xu Q, Mermel CH, Liang AC, Schlabach MR et al. 2012. Recurrent hemizygous deletions in cancers may optimize proliferative potential. Science 336:6090104–9
    [Google Scholar]
  108. 108.
    Davoli T, Uno H, Wooten EC, Elledge SJ. 2017. Tumor aneuploidy correlates with markers of immune evasion and with reduced response to immunotherapy. Science 355:6322eaaf8399
    [Google Scholar]
  109. 109.
    Fischer K, Fröhling S, Scherer SW, Brown JMA, Scholl C et al. 1997. Molecular cytogenetic delineation of deletions and translocations involving chromosome band 7q22 in myeloid leukemias. Blood 89:62036–41
    [Google Scholar]
  110. 110.
    Wong JC, Weinfurtner KM, del pilar Alzamora M, Kogan SC, Burgess MR et al. 2015. Functional evidence implicating chromosome 7q22 haploinsufficiency in myelodysplastic syndrome pathogenesis. eLife 4:e07839
    [Google Scholar]
  111. 111.
    Ouillette P, Collins R, Shakhan S, Li J, Li C et al. 2011. The prognostic significance of various 13q14 deletions in chronic lymphocytic leukemia. Clin. Cancer Res. 17:216778–90
    [Google Scholar]
  112. 112.
    Rego EM, Wang ZG, Peruzzi D, He LZ, Cordon-Cardo C, Pandolfi PP. 2001. Role of promyelocytic leukemia (Pml) protein in tumor suppression. J. Exp. Med. 193:4521–30
    [Google Scholar]
  113. 113.
    Welch JS, Klco JM, Varghese N, Nagarajan R, Ley TJ. 2011. Rara haploinsufficiency modestly influences the phenotype of acute promyelocytic leukemia in mice. Blood 117:82460–68
    [Google Scholar]
  114. 114.
    Gröschel S, Sanders MA, Hoogenboezem R, de Wit E, Bouwman BAM et al. 2014. A single oncogenic enhancer rearrangement causes concomitant EVI1 and GATA2 deregulation in leukemia. Cell 157:2369–81
    [Google Scholar]
  115. 115.
    Katayama S, Suzuki M, Yamaoka A, Keleku-Lukwete N, Katsuoka F et al. 2017. GATA2 haploinsufficiency accelerates EVI1-driven leukemogenesis. Blood 130:7908–19
    [Google Scholar]
  116. 116.
    Gröschel S, Sanders MA, Hoogenboezem R, Zeilemaker A, Havermans M et al. 2015. Mutational spectrum of myeloid malignancies with inv(3)/t(3;3) reveals a predominant involvement of RAS/RTK signaling pathways. Blood 125:1133–39
    [Google Scholar]
  117. 117.
    de Bruijn M, Dzierzak E. 2017. Runx transcription factors in the development and function of the definitive hematopoietic system. Blood 129:152061–69
    [Google Scholar]
  118. 118.
    Yan M, Burel SA, Peterson LF, Kanbe E, Iwasaki H et al. 2004. Deletion of an AML1-ETO C-terminal NcoR/SMRT-interacting region strongly induces leukemia development. PNAS 101:4917186–91
    [Google Scholar]
  119. 119.
    Pabst T, Mueller BU, Harakawa N, Schoch C, Haferlach T et al. 2001. AML1-ETO downregulates the granulocytic differentiation factor C/EBPα in t(8;21) myeloid leukemia. Nat. Med. 7:4444–51
    [Google Scholar]
  120. 120.
    Choi Y, Elagib KE, Delehanty LL, Goldfarb AN. 2006. Erythroid inhibition by the leukemic fusion AML1-ETO is associated with impaired acetylation of the major erythroid transcription factor GATA-1. Cancer Res. 66:62990–96
    [Google Scholar]
  121. 121.
    Vangala RK, Heiss-Neumann MS, Rangatia JS, Singh SM, Schoch C et al. 2003. The myeloid master regulator transcription factor PU.1 is inactivated by AML1-ETO in t(8;21) myeloid leukemia. Blood 101:1270–77
    [Google Scholar]
  122. 122.
    Rubin C, Larson R, Anastasi J, Winter J, Thangavelu M et al. 1990. t(3;21)(q26;q22): a recurring chromosomal abnormality in therapy-related myelodysplastic syndrome and acute myeloid leukemia. Blood 76:122594–98
    [Google Scholar]
  123. 123.
    Loke J, Assi SA, Imperato MR, Ptasinska A, Cauchy P et al. 2017. RUNX1-ETO and RUNX1-EVI1 differentially reprogram the chromatin landscape in t(8;21) and t(3;21) AML. Cell Rep. 19:81654–68
    [Google Scholar]
  124. 124.
    Ptasinska A, Assi SA, Martinez-Soria N, Imperato MR, Piper J et al. 2014. Identification of a dynamic core transcriptional network in t(8;21) AML that regulates differentiation block and self-renewal. Cell Rep. 8:61974–88
    [Google Scholar]
  125. 125.
    Ben-Ami O, Friedman D, Leshkowitz D, Goldenberg D, Orlovsky K et al. 2013. Addiction of t(8;21) and inv(16) acute myeloid leukemia to native RUNX1. Cell Rep. 4:61131–43
    [Google Scholar]
  126. 126.
    Lukasik SM, Zhang L, Corpora T, Tomanicek S, Li Y et al. 2002. Altered affinity of CBFβ-SMMHC for Runx1 explains its role in leukemogenesis. Nat. Struct. Biol. 9:9674–79
    [Google Scholar]
  127. 127.
    Zhen T, Cao Y, Ren G, Zhao L, Hyde RK et al. 2020. RUNX1 and CBFβ-SMMHC transactivate target genes together in abnormal myeloid progenitors for leukemia development. Blood 136:212373–85
    [Google Scholar]
  128. 128.
    Hyde RK, Zhao L, Alemu L, Liu PP. 2015. Runx1 is required for hematopoietic defects and leukemogenesis in Cbfb-MYH11 knock-in mice. Leukemia 29:81771–78
    [Google Scholar]
  129. 129.
    Qi J, Singh S, Hua WK, Cai Q, Chao SW et al. 2015. HDAC8 inhibition specifically targets inv(16) acute myeloid leukemic stem cells by restoring p53 acetylation. Cell Stem Cell 17:5597–610
    [Google Scholar]
  130. 130.
    Papaemmanuil E, Gerstung M, Bullinger L, Gaidzik VI, Paschka P et al. 2016. Genomic classification and prognosis in acute myeloid leukemia. N. Engl. J. Med. 374:232209–21
    [Google Scholar]
  131. 131.
    Sakurai M, Kunimoto H, Watanabe N, Fukuchi Y, Yuasa S et al. 2014. Impaired hematopoietic differentiation of RUNX1-mutated induced pluripotent stem cells derived from FPD/AML patients. Leukemia 28:122344–54
    [Google Scholar]
  132. 132.
    Estevez B, Borst S, Jarocha D, Sudunagunta V, Gonzalez M et al. 2021. RUNX-1 haploinsufficiency causes a marked deficiency of megakaryocyte-biased hematopoietic progenitor cells. Blood 137:192662–75
    [Google Scholar]
  133. 133.
    Bellissimo DC, Speck NA. 2017. RUNX1 mutations in inherited and sporadic leukemia. Front. Cell Dev. Biol. 5:111
    [Google Scholar]
  134. 134.
    Antony-Debré I, Manchev VT, Balayn N, Bluteau D, Tomowiak C et al. 2015. Level of RUNX1 activity is critical for leukemic predisposition but not for thrombocytopenia. Blood 125:930–40
    [Google Scholar]
  135. 135.
    Tsai SC, Shih LY, Liang ST, Huang YJ, Kuo MC et al. 2015. Biological activities of RUNX1 mutants predict secondary acute leukemia transformation from chronic myelomonocytic leukemia and myelodysplastic syndromes. Clin. Cancer Res. 21:153541–51
    [Google Scholar]
  136. 136.
    McDevitt MA, Shivdasani RA, Fujiwara Y, Yang H, Orkin SH. 1997. A “knockdown” mutation created by cis-element gene targeting reveals the dependence of erythroid cell maturation on the level of transcription factor GATA-1. PNAS 94:136781–85
    [Google Scholar]
  137. 137.
    Gregory T, Yu C, Ma A, Orkin SH, Blobel GA, Weiss MJ. 1999. GATA-1 and erythropoietin cooperate to promote erythroid cell survival by regulating bcl-xL expression. Blood 94:187–96
    [Google Scholar]
  138. 138.
    Yoshida K, Toki T, Okuno Y, Kanezaki R, Shiraishi Y et al. 2013. The landscape of somatic mutations in Down syndrome-related myeloid disorders. Nat. Genet. 45:111293–99
    [Google Scholar]
  139. 139.
    Wechsler J, Greene M, McDevitt MA, Anastasi J, Karp JE et al. 2002. Acquired mutations in GATA1 in the megakaryoblastic leukemia of Down syndrome. Nat. Genet. 32:1148–52
    [Google Scholar]
  140. 140.
    Rodrigues NP, Janzen V, Forkert R, Dombkowski DM, Boyd AS et al. 2005. Haploinsufficiency of GATA-2 perturbs adult hematopoietic stem-cell homeostasis. Blood 106:2477–84
    [Google Scholar]
  141. 141.
    Persons DA, Allay JA, Allay ER, Ashmun RA, Orlic D et al. 1999. Enforced expression of the GATA-2 transcription factor blocks normal hematopoiesis. Blood 93:2488–99
    [Google Scholar]
  142. 142.
    Collin M, Dickinson R, Bigley V. 2015. Haematopoietic and immune defects associated with GATA2 mutation. Br. J. Haematol. 169:2173–87
    [Google Scholar]
  143. 143.
    Ping N, Sun A, Song Y, Wang Q, Yin J et al. 2016. Exome sequencing identifies highly recurrent somatic GATA2 and CEBPA mutations in acute erythroid leukemia. Leukemia 31:1195–202
    [Google Scholar]
  144. 144.
    di Genua C, Valletta S, Buono M, Stoilova B, Sweeney C et al. 2020. C/EBPα and GATA-2 mutations induce bilineage acute erythroid leukemia through transformation of a neomorphic neutrophil-erythroid progenitor. Cancer Cell 37:5690–704
    [Google Scholar]
  145. 145.
    Sperling AS, Guerra VA, Kennedy JA, Yan Y, Hsu JI et al. 2022. Lenalidomide promotes the development of TP53-mutated therapy-related myeloid neoplasms. Blood 140:161753–63
    [Google Scholar]
  146. 146.
    Chen S, Wang Q, Yu H, Capitano ML, Vemula S et al. 2019. Mutant p53 drives clonal hematopoiesis through modulating epigenetic pathway. Nat. Commun. 101:5649
    [Google Scholar]
  147. 147.
    Loizou E, Banito A, Livshits G, Ho YJ, Koche RP et al. 2019. A gain-of-function p53-mutant oncogene promotes cell fate plasticity and myeloid leukemia through the pluripotency factor FOXH1. Cancer Discov. 9:7962–79
    [Google Scholar]
  148. 148.
    Iwasaki H, Somoza C, Shigematsu H, Duprez EA, Iwasaki-Arai J et al. 2005. Distinctive and indispensable roles of PU.1 in maintenance of hematopoietic stem cells and their differentiation. Blood 106:51590–600
    [Google Scholar]
  149. 149.
    Scott EW, Simon MC, Anastasi J, Singh H. 1994. Requirement of transcription factor PU.1 in the development of multiple hematopoietic lineages. Science 265:51781573–77
    [Google Scholar]
  150. 150.
    Rosenbauer F, Wagner K, Kutok JL, Iwasaki H, le Beau MM et al. 2004. Acute myeloid leukemia induced by graded reduction of a lineage-specific transcription factor, PU.1. Nat. Genet. 36:6624–30
    [Google Scholar]
  151. 151.
    Huang G, Zhang P, Hirai H, Elf S, Yan X et al. 2007. PU.1 is a major downstream target of AML1 (RUNX1) in adult mouse hematopoiesis. Nat. Genet. 40:151–60
    [Google Scholar]
  152. 152.
    Trinh BQ, Ummarino S, Zhang Y, Ebralidze AK, Bassal MA et al. 2021. Myeloid lncRNA LOUP mediates opposing regulatory effects of RUNX1 and RUNX1-ETO in t(8;21) AML. Blood 138:151331–44
    [Google Scholar]
  153. 153.
    van der Kouwe E, Heller G, Czibere A, Pulikkan JA, Agreiter C et al. 2021. Core-binding factor leukemia hijacks the T-cell-prone PU.1 antisense promoter. Blood 138:151345–58
    [Google Scholar]
  154. 154.
    Huang G, Zhao X, Wang L, Elf S, Xu H et al. 2011. The ability of MLL to bind RUNX1 and methylate H3K4 at PU.1 regulatory regions is impaired by MDS/AML-associated RUNX1/AML1 mutations. Blood 118:256544–52
    [Google Scholar]
  155. 155.
    Mueller BU, Pabst T, Fos J, Petkovic V, Fey MF et al. 2006. ATRA resolves the differentiation block in t(15;17) acute myeloid leukemia by restoring PU.1 expression. Blood 107:83330–38
    [Google Scholar]
  156. 156.
    Mizuki M, Schwäble J, Steur C, Choudhary C, Agrawal S et al. 2003. Suppression of myeloid transcription factors and induction of STAT response genes by AML-specific Flt3 mutations. Blood 101:83164–73
    [Google Scholar]
  157. 157.
    Yun H, Narayan N, Vohra S, Giotopoulos G, Mupo A et al. 2021. Mutational synergy during leukemia induction remodels chromatin accessibility, histone modifications and three-dimensional DNA topology to alter gene expression. Nat. Genet. 53:101443–55
    [Google Scholar]
  158. 158.
    Yeamans C, Wang D, Paz-Priel I, Torbett BE, Tenen DG, Friedman AD. 2007. C/EBPα binds and activates the PU.1 distal enhancer to induce monocyte lineage commitment. Blood 110:93136–42
    [Google Scholar]
  159. 159.
    Gu X, Ebrahem Q, Mahfouz RZ, Hasipek M, Enane F et al. 2018. Leukemogenic nucleophosmin mutation disrupts the transcription factor hub that regulates granulomonocytic fates. J. Clin. Investig. 128:104260–79
    [Google Scholar]
  160. 160.
    Pianigiani G, Betti C, Bigerna B, Rossi R, Brunetti L. 2020. PU.1 subcellular localization in acute myeloid leukaemia with mutated NPM1. Br. J. Haematol. 188:1184–87
    [Google Scholar]
  161. 161.
    Aivalioti MM, Bartholdy BA, Pradhan K, Bhagat TD, Zintiridou A et al. 2022. PU.1-dependent enhancer inhibition separates Tet2-deficient hematopoiesis from malignant transformation. Blood Cancer Discov. 3:5444–67This paper demonstrates the importance of minor changes in TF dose for the progression of leukemia.
    [Google Scholar]
  162. 162.
    Kaasinen E, Kuismin O, Rajamäki K, Ristolainen H, Aavikko M et al. 2019. Impact of constitutional TET2 haploinsufficiency on molecular and clinical phenotype in humans. Nat. Commun. 101:1252
    [Google Scholar]
  163. 163.
    Tulstrup M, Soerensen M, Hansen JW, Gillberg L, Needhamsen M et al. 2021. TET2 mutations are associated with hypermethylation at key regulatory enhancers in normal and malignant hematopoiesis. Nat. Commun. 121:6061
    [Google Scholar]
  164. 164.
    Ohlsson E, Schuster MB, Hasemann M, Porse BT. 2015. The multifaceted functions of C/EBPα in normal and malignant haematopoiesis. Leukemia 30:4767–75
    [Google Scholar]
  165. 165.
    Mannelli F, Ponziani V, Bencini S, Bonetti MI, Benelli M et al. 2017. CEBPA-double-mutated acute myeloid leukemia displays a unique phenotypic profile: a reliable screening method and insight into biological features. Haematologica 102:3529–40
    [Google Scholar]
  166. 166.
    Kirstetter P, Schuster MB, Bereshchenko O, Moore S, Dvinge H et al. 2008. Modeling of C/EBPα mutant acute myeloid leukemia reveals a common expression signature of committed myeloid leukemia-initiating cells. Cancer Cell 13:4299–310
    [Google Scholar]
  167. 167.
    Bereshchenko O, Mancini E, Moore S, Bilbao D, Månsson R et al. 2009. Hematopoietic stem cell expansion precedes the generation of committed myeloid leukemia-initiating cells in C/EBPα mutant AML. Cancer Cell 16:5390–400
    [Google Scholar]
  168. 168.
    Khanna-Gupta A, Abayasekara N, Levine M, Sun H, Virgilio M et al. 2012. Up-regulation of translation eukaryotic initiation factor 4E in nucleophosmin 1 haploinsufficient cells results in changes in CCAAT enhancer-binding protein α activity: implications in myelodysplastic syndrome and acute myeloid leukemia. J. Biol. Chem. 287:3932728–37
    [Google Scholar]
  169. 169.
    Tanaka T, Tanaka K, Ogawa S, Kurokawa M, Mitani K et al. 1995. An acute myeloid leukemia gene, AML1, regulates hemopoietic myeloid cell differentiation and transcriptional activation antagonistically by two alternative spliced forms. EMBO J. 14:2341–50
    [Google Scholar]
  170. 170.
    Gialesaki S, Bräuer-Hartmann D, Bhayadia R, Alejo-Valle O, Issa H et al. 2022. RUNX1 isoform disequilibrium in the development of trisomy 21 associated myeloid leukemia. bioRxiv 2022.03.07.483334. https://doi.org/10.1101/2022.03.07.483334 This paper demonstrates how gain of a chromosome can negatively affect the dose of a TF.
  171. 171.
    Sakurai H, Harada Y, Ogata Y, Kagiyama Y, Shingai N et al. 2017. Overexpression of RUNX1 short isoform has an important role in the development of myelodysplastic/myeloproliferative neoplasms. Blood Adv. 1:181382–86
    [Google Scholar]
  172. 172.
    Figueroa ME, Lugthart S, Li Y, Erpelinck-Verschueren C, Deng X et al. 2010. DNA methylation signatures identify biologically distinct subtypes in acute myeloid leukemia. Cancer Cell 17:113–27
    [Google Scholar]
  173. 173.
    Mulet-Lazaro R, van Herk S, Erpelinck C, Bindels E, Sanders MA et al. 2021. Allele-specific expression of GATA2 due to epigenetic dysregulation in CEBPA double-mutant AML. Blood 138:2160–77
    [Google Scholar]
  174. 174.
    Lin TC, Hou HA, Chou WC, Ou DL, Yu SL et al. 2010. CEBPA methylation as a prognostic biomarker in patients with de novo acute myeloid leukemia. Leukemia 25:132–40
    [Google Scholar]
  175. 175.
    Mill CP, Fiskus W, DiNardo CD, Qian Y, Raina K et al. 2019. RUNX1-targeted therapy for AML expressing somatic or germline mutation in RUNX1. Blood 134:159–73This paper highlights one possible treatment option for neoplasms dependent on a HITF.
    [Google Scholar]
  176. 176.
    Krönke J, Fink EC, Hollenbach PW, MacBeth KJ, Hurst SN et al. 2015. Lenalidomide induces ubiquitination and degradation of CK1α in del(5q) MDS. Nature 523:7559183–88
    [Google Scholar]
  177. 177.
    Pich O, Reyes-Salazar I, Gonzalez-Perez A, Lopez-Bigas N. 2022. Discovering the drivers of clonal hematopoiesis. Nat. Commun. 131:4267
    [Google Scholar]
  178. 178.
    Jaiswal S, Fontanillas P, Flannick J, Manning A, Grauman PV et al. 2014. Age-related clonal hematopoiesis associated with adverse outcomes. N. Engl. J. Med. 371:262488–98
    [Google Scholar]
  179. 179.
    Bolton KL, Koh Y, Foote MB, Im H, Jee J et al. 2021. Clonal hematopoiesis is associated with risk of severe Covid-19. Nat. Commun. 121:5975
    [Google Scholar]
  180. 180.
    Gallì A, Todisco G, Catamo E, Sala C, Elena C et al. 2021. Relationship between clone metrics and clinical outcome in clonal cytopenia. Blood 138:11965–76
    [Google Scholar]
  181. 181.
    Onida F, Kantarjian HM, Smith TL, Ball G, Keating MJ et al. 2002. Prognostic factors and scoring systems in chronic myelomonocytic leukemia: a retrospective analysis of 213 patients. Blood 99:3840–49
    [Google Scholar]
  182. 182.
    Such E, Cervera J, Costa D, Solé F, Vallespí T et al. 2011. Cytogenetic risk stratification in chronic myelomonocytic leukemia. Haematologica 96:3375–83
    [Google Scholar]
  183. 183.
    Papaemmanuil E, Gerstung M, Malcovati L, Tauro S, Gundem G et al. 2013. Clinical and biological implications of driver mutations in myelodysplastic syndromes. Blood 122:223616–27
    [Google Scholar]
  184. 184.
    Grinfeld J, Nangalia J, Baxter EJ, Wedge DC, Angelopoulos N et al. 2018. Classification and personalized prognosis in myeloproliferative neoplasms. N. Engl. J. Med. 379:151416–30
    [Google Scholar]
  185. 185.
    Marcellino BK, Hoffman R, Tripodi J, Lu M, Kosiorek H et al. 2018. Advanced forms of MPNs are accompanied by chromosomal abnormalities that lead to dysregulation of TP53. Blood Adv. 2:243581–89
    [Google Scholar]
  186. 186.
    Yoshizato T, Nannya Y, Atsuta Y, Shiozawa Y, Iijima-Yamashita Y et al. 2017. Genetic abnormalities in myelodysplasia and secondary acute myeloid leukemia: impact on outcome of stem cell transplantation. Blood 129:172347–58
    [Google Scholar]
  187. 187.
    Bernard E, Tuechler H, Greenberg PL, Hasserjian RP, Ossa JEA et al. 2022. Molecular International Prognostic Scoring System for myelodysplastic syndromes. NEJM Evid. 1:7 https://doi.org/10.1056/EVIDoa2200008
    [Google Scholar]
  188. 188.
    Bejar R, Stevenson K, Abdel-Wahab O, Galili N, Nilsson B et al. 2011. Clinical effect of point mutations in myelodysplastic syndromes. N. Engl. J. Med. 364:262496–506
    [Google Scholar]
  189. 189.
    Grimwade D, Walker H, Harrison G, Oliver F, Chatters S et al. 2001. The predictive value of hierarchical cytogenetic classification in older adults with acute myeloid leukemia (AML): analysis of 1065 patients entered into the United Kingdom Medical Research Council AML11 trial. Blood 98:51312–20
    [Google Scholar]
  190. 190.
    Tyner JW, Tognon CE, Bottomly D, Wilmot B, Kurtz SE et al. 2018. Functional genomic landscape of acute myeloid leukaemia. Nature 562:7728526–31
    [Google Scholar]
  191. 191.
    Fröhling S, Schlenk RF, Kayser S, Morhardt M, Benner A et al. 2006. Cytogenetics and age are major determinants of outcome in intensively treated acute myeloid leukemia patients older than 60 years: results from AMLSG trial AML HD98-B. Blood 108:103280–88
    [Google Scholar]
  192. 192.
    Ley TJ, Miller C, Ding L, Raphael BJ, Mungall AJ et al. 2013. Genomic and epigenomic landscapes of adult de novo acute myeloid leukemia. N. Engl. J. Med. 368:222059–74
    [Google Scholar]
  193. 193.
    Lindsley RC, Mar BG, Mazzola E, Grauman PV, Shareef S et al. 2015. Acute myeloid leukemia ontogeny is defined by distinct somatic mutations. Blood 125:91367–76
    [Google Scholar]
/content/journals/10.1146/annurev-pathmechdis-051222-013421
Loading
/content/journals/10.1146/annurev-pathmechdis-051222-013421
Loading

Data & Media loading...

  • Article Type: Review Article
This is a required field
Please enter a valid email address
Approval was a Success
Invalid data
An Error Occurred
Approval was partially successful, following selected items could not be processed due to error