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.

  • Article
  • Published:

Germline NPM1 mutations lead to altered rRNA 2′-O-methylation and cause dyskeratosis congenita

Abstract

RNA modifications are emerging as key determinants of gene expression. However, compelling genetic demonstrations of their relevance to human disease are lacking. Here, we link ribosomal RNA 2′-O-methylation (2′-O-Me) to the etiology of dyskeratosis congenita. We identify nucleophosmin (NPM1) as an essential regulator of 2′-O-Me on rRNA by directly binding C/D box small nucleolar RNAs, thereby modulating translation. We demonstrate the importance of 2′-O-Me-regulated translation for cellular growth, differentiation and hematopoietic stem cell maintenance, and show that Npm1 inactivation in adult hematopoietic stem cells results in bone marrow failure. We identify NPM1 germline mutations in patients with dyskeratosis congenita presenting with bone marrow failure and demonstrate that they are deficient in small nucleolar RNA binding. Mice harboring a dyskeratosis congenita germline Npm1 mutation recapitulate both hematological and nonhematological features of dyskeratosis congenita. Thus, our findings indicate that impaired 2′-O-Me can be etiological to human disease.

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

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

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

Fig. 1: NPM1 regulates 2′-O-Me through snoRNA binding.
Fig. 2: NPM1 regulates cellular growth and differentiation through 2′-O-Me.
Fig. 3: Acute deletion of Npm1 in adult mouse HSCs leads to BMF.
Fig. 4: NPM1 germline mutations identified in patients with dyskeratosis congenita.
Fig. 5: NPM1D180del mice show multi-organ features of dyskeratosis congenita.

Similar content being viewed by others

Data availability

Raw and preprocessed sequencing and microarray data may be accessed from the Gene Expression Omnibus with accession number GSE135726. Since the informed consent obtained from dyskeratosis congenita patients does not allow for public deposition of the data, the WES data from patient CM108 (NPMD178H) and healthy controls can be communicated upon reasonable request to J.S. and the WES data from the patient harboring the NPM1D180del mutation can be communicated upon reasonable request to I.D. and T.J.V.

Code availability

Code for microarray analysis is available in Supplementary Information 1 file. The code for the WES analysis is available at https://github.com/UCLGeneticsInstitute/DNASeq_pipeline.

References

  1. Davalos, V., Blanco, S. & Esteller, M. SnapShot: messenger RNA modifications. Cell 174, 498–498.e1 (2018).

    Article  CAS  PubMed  Google Scholar 

  2. Esteller, M. & Pandolfi, P. P. The epitranscriptome of noncoding RNAs in cancer. Cancer Discov. 7, 359–368 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Frye, M., Harada, B. T., Behm, M. & He, C. RNA modifications modulate gene expression during development. Science 361, 1346–1349 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Safra, M. et al. The m1A landscape on cytosolic and mitochondrial mRNA at single-base resolution. Nature 551, 251–255 (2017).

    Article  CAS  PubMed  Google Scholar 

  5. Chawla, M., Oliva, R., Bujnicki, J. M. & Cavallo, L. An atlas of RNA base pairs involving modified nucleobases with optimal geometries and accurate energies. Nucleic Acids Res. 43, 9573 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Polikanov, Y. S., Melnikov, S. V., Söll, D. & Steitz, T. A. Structural insights into the role of rRNA modifications in protein synthesis and ribosome assembly. Nat. Struct. Mol. Biol. 22, 342–344 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Sharma, S. & Lafontaine, D. L. J. ‘View from a bridge’: a new perspective on eukaryotic rRNA base modification. Trends Biochem. Sci. 40, 560–575 (2015).

    Article  CAS  PubMed  Google Scholar 

  8. Kiss, T. Small nucleolar RNAs: an abundant group of noncoding RNAs with diverse cellular functions. Cell 109, 145–148 (2002).

    Article  CAS  PubMed  Google Scholar 

  9. Watkins, N. J. & Bohnsack, M. T. The box C/D and H/ACA snoRNPs: key players in the modification, processing and the dynamic folding of ribosomal RNA. Wiley Interdiscip. Rev. RNA 3, 397–414 (2012).

    Article  CAS  PubMed  Google Scholar 

  10. Borer, R. A., Lehner, C. F., Eppenberger, H. M. & Nigg, E. A. Major nucleolar proteins shuttle between nucleus and cytoplasm. Cell 56, 379–390 (1989).

    Article  CAS  PubMed  Google Scholar 

  11. Grisendi, S., Mecucci, C., Falini, B. & Pandolfi, P. P. Nucleophosmin and cancer. Nat. Rev. Cancer 6, 493–505 (2006).

    Article  CAS  PubMed  Google Scholar 

  12. Falini, B. et al. Cytoplasmic nucleophosmin in acute myelogenous leukemia with a normal karyotype. N. Engl. J. Med. 352, 254–266 (2005).

    Article  CAS  PubMed  Google Scholar 

  13. Box, J. K. et al. Nucleophosmin: from structure and function to disease development. BMC Mol. Biol. 17, 19 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  14. Grisendi, S. et al. Role of nucleophosmin in embryonic development and tumorigenesis. Nature 437, 147–153 (2005).

    Article  CAS  PubMed  Google Scholar 

  15. Schwartz, S. et al. Transcriptome-wide mapping reveals widespread dynamic-regulated pseudouridylation of ncRNA and mRNA. Cell 159, 148–162 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Dong, Z. W. et al. RTL-P: a sensitive approach for detecting sites of 2′-O-methylation in RNA molecules. Nucleic Acids Res. 40, e157 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Krogh, N. et al. Profiling of 2′-O-Me in human rRNA reveals a subset of fractionally modified positions and provides evidence for ribosome heterogeneity. Nucleic Acids Res. 44, 7884–7895 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Natchiar, S. K., Myasnikov, A. G., Hazemann, I. & Klaholz, B. P. Visualizing the role of 2′-OH rRNA methylations in the human ribosome structure. Biomolecules 8, E125 (2018).

    Article  PubMed  Google Scholar 

  19. Natchiar, S. K., Myasnikov, A. G., Kratzat, H., Hazemann, I. & Klaholz, B. P. Visualization of chemical modifications in the human 80S ribosome structure. Nature 551, 472–477 (2017).

    Article  CAS  PubMed  Google Scholar 

  20. Sharma, S., Marchand, V., Motorin, Y. & Lafontaine, D. L. J. Identification of sites of 2′-O-methylation vulnerability in human ribosomal RNAs by systematic mapping. Sci. Rep. 7, 11490 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  21. Steitz, T. A. A structural understanding of the dynamic ribosome machine. Nat. Rev. Mol. Cell Biol. 9, 242–253 (2008).

    Article  CAS  PubMed  Google Scholar 

  22. Liang, X. H., Liu, Q. & Fournier, M. J. Loss of rRNA modifications in the decoding center of the ribosome impairs translation and strongly delays pre-rRNA processing. RNA 15, 1716–1728 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Proud, C. G. mTORC1 signalling and mRNA translation. Biochem. Soc. Trans. 37, 227–231 (2009).

    Article  CAS  PubMed  Google Scholar 

  24. Wullschleger, S., Loewith, R. & Hall, M. N. TOR signaling in growth and metabolism. Cell 124, 471–484 (2006).

    Article  CAS  PubMed  Google Scholar 

  25. Reschke, M. et al. Characterization and analysis of the composition and dynamics of the mammalian riboproteome. Cell Rep. 4, 1276–1287 (2013).

    Article  CAS  PubMed  Google Scholar 

  26. Ludwig, L. S. et al. Altered translation of GATA1 in Diamond–Blackfan anemia. Nat. Med. 20, 748–753 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Mills, E. W. & Green, R. Ribosomopathies: there’s strength in numbers.Science 358, eaan2755 (2017).

    Article  PubMed  Google Scholar 

  28. Bugaud, O. et al. Kinetics of CrPV and HCV IRES-mediated eukaryotic translation using single-molecule fluorescence microscopy. RNA 23, 1626–1635 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Sonenberg, N. & Hinnebusch, A. G. Regulation of translation initiation in eukaryotes: mechanisms and biological targets. Cell 136, 731–745 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Weingarten-Gabbay, S. et al. Systematic discovery of cap-independent translation sequences in human and viral genomes. Science 351, aad4939 (2016).

    Article  PubMed  Google Scholar 

  31. Holcik, M. & Korneluk, R. G. Functional characterization of the X-linked inhibitor of apoptosis (XIAP) internal ribosome entry site element: role of La autoantigen in XIAP translation. Mol. Cell. Biol. 20, 4648–4657 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Holcik, M., Lefebvre, C., Yeh, C., Chow, T. & Korneluk, R. G. A new internal-ribosome-entry-site motif potentiates XIAP-mediated cytoprotection. Nat. Cell Biol. 1, 190–192 (1999).

    Article  CAS  PubMed  Google Scholar 

  33. Jiang, H., Coleman, J., Miskimins, R., Srinivasan, R. & Miskimins, W. K. Cap-independent translation through the p27 5′-UTR. Nucleic Acids Res. 35, 4767–4778 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Kullmann, M., Göpfert, U., Siewe, B. & Hengst, L. ELAV/Hu proteins inhibit p27 translation via an IRES element in the p27 5′UTR. Genes Dev. 16, 3087–3099 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Miskimins, W. K., Wang, G., Hawkinson, M. & Miskimins, R. Control of cyclin-dependent kinase inhibitor p27 expression by cap-independent translation. Mol. Cell. Biol. 21, 4960–4967 (2001).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Knight, Z. A. et al. Molecular profiling of activated neurons by phosphorylated ribosome capture. Cell 151, 1126–1137 (2012).

    Article  CAS  PubMed  Google Scholar 

  37. O, E. et al. Integrin αvβ3 is not significantly implicated in the anti-migratory effect of anti-angiogenic urokinase kringle domain. Oncol. Rep. 20, 631–636 (2008).

    PubMed  Google Scholar 

  38. Zhou, F. et al. AML1-ETO requires enhanced C/D box snoRNA/RNP formation to induce self-renewal and leukaemia. Nat. Cell Biol. 19, 844–855 (2017).

    Article  CAS  PubMed  Google Scholar 

  39. Baliga, B. S., Mankad, M., Shah, A. K. & Mankad, V. N. Mechanism of differentiation of human erythroleukaemic cell line K562 by hemin. Cell Prolif. 26, 519–529 (1993).

    Article  CAS  PubMed  Google Scholar 

  40. Ayadi, L., Galvanin, A., Pichot, F., Marchand, V. & Motorin, Y. RNA ribose methylation (2′-O-methylation): occurrence, biosynthesis and biological functions. Biochim. Biophys. Acta Gene Regul. Mech. 1862, 253–269 (2019).

    Article  CAS  PubMed  Google Scholar 

  41. Dai, Q. et al. Nm-seq maps 2′-O-methylation sites in human mRNA with base precision. Nat. Methods 14, 695–698 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Dauwerse, J. G. et al. Mutations in genes encoding subunits of RNA polymerases I and III cause Treacher Collins syndrome. Nat. Genet. 43, 20–22 (2011).

    Article  CAS  PubMed  Google Scholar 

  43. Draptchinskaia, N. et al. The gene encoding ribosomal protein S19 is mutated in Diamond–Blackfan anaemia. Nat. Genet. 21, 169–175 (1999).

    Article  CAS  PubMed  Google Scholar 

  44. Luzzatto, L. & Karadimitris, A. Dyskeratosis and ribosomal rebellion. Nat. Genet. 19, 6–7 (1998).

    Article  CAS  PubMed  Google Scholar 

  45. Narla, A. & Ebert, B. L. Ribosomopathies: human disorders of ribosome dysfunction. Blood 115, 3196–3205 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Ridanpää, M. et al. Mutations in the RNA component of RNase MRP cause a pleiotropic human disease, cartilage-hair hypoplasia. Cell 104, 195–203 (2001).

    Article  PubMed  Google Scholar 

  47. Ruggero, D. et al. Dyskeratosis congenita and cancer in mice deficient in ribosomal RNA modification. Science 299, 259–262 (2003).

    Article  CAS  PubMed  Google Scholar 

  48. Armistead, J. & Triggs-Raine, B. Diverse diseases from a ubiquitous process: the ribosomopathy paradox. FEBS Lett. 588, 1491–1500 (2014).

    Article  CAS  PubMed  Google Scholar 

  49. Dokal, I. Dyskeratosis congenita. Hematology Am. Soc. Hematol. Educ. Program 2011, 480–486 (2011).

    Article  PubMed  Google Scholar 

  50. Danilova, N. & Gazda, H. T. Ribosomopathies: how a common root can cause a tree of pathologies. Dis. Model Mech. 8, 1013–1026 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Fumagalli, S. & Thomas, G. The role of p53 in ribosomopathies. Semin. Hematol. 48, 97–105 (2011).

    Article  CAS  PubMed  Google Scholar 

  52. Yoon, A. et al. Impaired control of IRES-mediated translation in X-linked dyskeratosis congenita. Science 312, 902–906 (2006).

    Article  CAS  PubMed  Google Scholar 

  53. Hisaoka, M., Nagata, K. & Okuwaki, M. Intrinsically disordered regions of nucleophosmin/B23 regulate its RNA binding activity through their inter- and intra-molecular association. Nucleic Acids Res. 42, 1180–1195 (2014).

    Article  CAS  PubMed  Google Scholar 

  54. Dunkle, J. A. et al. Molecular recognition and modification of the 30S ribosome by the aminoglycoside-resistance methyltransferase NpmA. Proc. Natl Acad. Sci. USA 111, 6275–6280 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Gutierrez, B., Douthwaite, S. & Gonzalez-Zorn, B. Indigenous and acquired modifications in the aminoglycoside binding sites of Pseudomonas aeruginosa rRNAs. RNA Biol. 10, 1324–1332 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  56. Quadros, R. M. et al. Easi-CRISPR: a robust method for one-step generation of mice carrying conditional and insertion alleles using long ssDNA donors and CRISPR ribonucleoproteins. Genome Biol. 18, 92 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  57. De Vos, D., Bruggeman, F. J., Westerhoff, H. V. & Bakker, B. M. How molecular competition influences fluxes in gene expression networks. PLoS ONE 6, e28494 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Khajuria, R. K. et al. Ribosome levels selectively regulate translation and lineage commitment in human hematopoiesis. Cell 173, 90–103.e19 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Lodish, H. F. Model for the regulation of mRNA translation applied to haemoglobin synthesis. Nature 251, 385–388 (1974).

    Article  CAS  PubMed  Google Scholar 

  60. Schwanhäusser, B. et al. Global quantification of mammalian gene expression control. Nature 473, 337–342 (2011).

    Article  PubMed  Google Scholar 

  61. Buszczak, M., Signer, R. A. & Morrison, S. J. Cellular differences in protein synthesis regulate tissue homeostasis. Cell 159, 242–251 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Dinman, J. D. Pathways to specialized ribosomes: the Brussels lecture. J. Mol. Biol. 428, 2186–2194 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Shi, Z. & Barna, M. Translating the genome in time and space: specialized ribosomes, RNA regulons, and RNA-binding proteins. Annu. Rev. Cell Dev. Biol. 31, 31–54 (2015).

    Article  CAS  PubMed  Google Scholar 

  64. Simsek, D. et al. The mammalian ribo-interactome reveals ribosome functional diversity and heterogeneity. Cell 169, 1051–1065.e18 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Sloan, K. E. et al. Tuning the ribosome: the influence of rRNA modification on eukaryotic ribosome biogenesis and function.RNA Biol. 14, 1138–1152 (2017).

    Article  PubMed  Google Scholar 

  66. Vassiliou, G. S. et al. Mutant nucleophosmin and cooperating pathways drive leukemia initiation and progression in mice. Nat. Genet. 43, 470–475 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Uren, P. J. et al. Site identification in high-throughput RNA-protein interaction data. Bioinformatics 28, 3013–3020 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  68. Machnicka, M. A. et al. MODOMICS: a database of RNA modification pathways—2013 update. Nucleic Acids Res. 41, D262–D267 (2013).

    Article  CAS  PubMed  Google Scholar 

  69. Heinz, S. et al. Simple combinations of lineage-determining transcription factors prime cis-regulatory elements required for macrophage and B cell identities. Mol. Cell 38, 576–589 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Arai, F. et al. Tie2/angiopoietin-1 signaling regulates hematopoietic stem cell quiescence in the bone marrow niche. Cell 118, 149–161 (2004).

    Article  CAS  PubMed  Google Scholar 

  71. Kanehisa, M., Furumichi, M., Tanabe, M., Sato, Y. & Morishima, K. KEGG: new perspectives on genomes, pathways, diseases and drugs. Nucleic Acids Res. 45, D353–D361 (2017).

    Article  CAS  PubMed  Google Scholar 

  72. Bluteau, O. et al. A landscape of germ line mutations in a cohort of inherited bone marrow failure patients. Blood 131, 717–732 (2018).

    Article  CAS  PubMed  Google Scholar 

  73. Pontikos, N. et al. Phenopolis: an open platform for harmonization and analysis of genetic and phenotypic data. Bioinformatics 33, 2421–2423 (2017).

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

D.N. was supported by an EMBO long-term fellowship (no. EMBO-LTF498-2014). K.I. was supported by National Institutes of Health grants (no. R01DK98263, R01DK115577 and R01HL148852), and is a Scholar of The Leukemia and Lymphoma Society. A.H.B. was supported by the Damon Runyon Cancer Research Foundation (no. 2142-12). This work was supported in part by the European Research Council Consolidator (grant no. 311660) and Cancéropole Ile-de-France (no. 2011-1-LABEL-1-AXE2-UP7-3) to J.S; Medical Research Council (grant no. MR/PO18440/1) and Bloodwise (grant no. 14032) to I.D.; the Fondazione AIRC per la Ricerca sul Cancro IG 2016 (grant no. 18568) and the European Research Council Advanced Grant 2016 (no. 740230) to B.F.; and by an Outstanding Investigator Award R35 (grant no. CA197529) and the SHINE grant (no. 5R01DK115536) awarded by National Institutes of Health to P.P.P.

Author information

Authors and Affiliations

Authors

Contributions

D.N., J.G.C. and P.P.P. designed the experiments and discussed the data. D.N., J.G.C., L.L. and P.P.P. wrote the manuscript. D.N. designed and performed the biochemical and translation-related experiments and analyzed the data. D.N. generated the NPMD180del mouse model, and designed and executed the experiments. S.G. generated the Npm1 conditional knockout mouse model. K.I., S.G. and P.S. performed the conditional Npm1 knockout experiments. A.H.B., A.M. and R.B.D. designed and performed the HiTS-CLIP experiments. D.B. and E.M. performed the computational analysis of the HiTS-CLIP data. Y.Z. designed and performed the northern blot experiments. A.C.B. provided critical reagents. A.G. and C.A.M. provided animal technical assistance. M.P.M. and B.F. provided the human AML samples. K.C. and J.D.L. performed and analyzed the microarray experiments and data. L.L., L.M.M. and O.P. performed the pathology analyses of the hematopoietic features of the NPM1 mutants. T.J.V., I.D., O.B. and J.S. provided and analyzed the dyskeratosis congenita exome dataset. M.S. and S.S. performed and analyzed the PSI-seq experiments and data.

Corresponding author

Correspondence to Pier Paolo Pandolfi.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

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

Integrated supplementary information

Supplementary Fig. 1 NPM1 regulates 2’-O-Me through binding of C/D box snoRNAs.

a, Pearson’s Correlation of psi-ratio mean Npm1+/+ and mean Npm1−/− MEFs (n=3 independent experiments). P=0.9. b, NPM1-IP was performed with Npm1+/+ MEFs and snoRNA enrichment was calculated relative to total input. SnordCtrl is snoRD13 that was not identified as an NPM1-bound snoRNA. Data are presented as mean±SD, n=4 independent experiments. c, Western blot analysis of FBL expression levels in Npm1+/+ and Npm1−/− MEFs. HSP90 served as a loading control. Shown is a representative blot out of n=5 independent experiments. d, qPCR analysis of the relative abundance of the specific snoRNAs (x-axis) in Npm1+/+ and Npm1−/− MEFs. Data are presented as mean±SD, n=4 independent experiments.. e, Npm1−/− MEFs were stably transduced to overexpress NPM1 and assessed for specific 2′-O-Me residues (x-axis) levels, mediated by C/D box snoRNAs identified in NPM1 HITS-CLIP. Data are presented as mean±SD of fold change of 2′-O-Me in Npm1−/−+NPM1 MEFs relative to Npm1+/+, n=3 independent experiments. Ctrl residue is the U1804 2′-O-Me modification on the 18S rRNA that is mediated by snoRD20, which was not identified to interact with NPM1 by HITS-CLIP. f, NPM-IP was performed using nuclear extracts, then blotted for NOP58 and DKC1. Neither of the proteins were detected. n=3 independent experiments. g, NPM1-IP was performed using Npm1+/+ MEFs. IP pellets were treated either with RNaseA or MNase for 15min at 37c to digest RNA. FBL was not detected in the treated supernatant (SN) indicating it does not interact with NPM1 via RNA binding. n=3 independent experiments. h, As a positive control, pS6-IP was performed using Npm1+/+ MEFs. pS6-IP pellets were treated either with RNaseA or MNase for 15min at 37c to degrade RNA then blotted for pS6 (for IP validation) and for the RNA binding protein PABP which was released to the supernatant. SN – supernatant. n=3 independent experiments. For all relevant panels, and unless otherwise stated, statistical significance was determined by one-tailed student’s t-test. Uncropped blot images are presented in Supplementary Fig. 11.

Supplementary Fig. 2 NPM1-deletion affects specific modes of translation and does not affect global translation.

a, 35S-Methionine incorporation assay of Npm1+/+ and Npm1−/− MEFs showed no differences on global protein synthesis. Quantity of labeled proteins was analyzed by radiography (RG). n=3 independent experiments. b, HPG metabolic labeling of Npm1+/+ and Npm1−/− MEFs showed no difference in global protein synthesis. Detection of HPG labeling was by azide AlexaFlour-647 and flow cytometry. n=3 independent experiments. c, Flow cytometry analysis of forward scatter and side scatter of Npm1+/+ and Npm1−/− MEFs. n=3 independent experiments. d, Analysis of total RNA content of Npm1+/+ and Npm1−/− MEFs. n=3 independent experiments. e, Northen blot of rRNA processing was performed using Npm1+/+ and Npm1−/− cells. n=3 independent experiments. f, WB analysis of mTOR downstream effectors in Npm1+/+ and Npm1−/− MEFs. HSP90 served as a loading control. n=3 independent experiments. g, Top panel: gene set enrichment (GSEA) analysis using the full set of GFP-positive genes showed statistically significant negative enrichment of putative IRES genes, NES = -1.359, p-value = 1.167e-4. Bottom panel: GSEA analysis of the restricted list of 583 IRES genes identified in Weingarten-Gabbay et al. and found a negative trend of IRES genes associated with Npm1+/- polysome-depleted transcripts, NES = -1.049, p-value = 0.31. h, WB of NPM1 in Npm1+/+, Npm1+/- and Npm1−/− MEFs. Shown is a representative blot of n=3 independent experiments. i, SnoRNA abundance in Npm1+/+ and Npm1+/- MEFs. Data are presented as mean±SD, n=4 independent experiments. j, Levels of specific 2′-O-Me residues (x-axis) mediated by only C/D box snoRNAs identified to interact with NPM1 by HITS-CLIP in Npm1+/- MEFs. Data are presented as mean±SD (n=3 independent experiments) of fold change of 2′-O-Me in Npm1+/- MEFs relative to Npm1+/+. Ctrl residue is the U1804 2′-O-Me modification on the 18S rRNA that is mediated by snoRD20, which was not identified to interact with NPM1 by HITS-CLIP. k, qPCR analysis of Cdkn1b (p27 transcript) and Xiap levels in Npm1−/− MEFs relative to Npm1+/+. Data are presented as mean±SD, n=3 independent experiments. l-m, Ribosome fractionation was performed using Npm1+/+ and Npm1−/− MEFs. ActinB, Gapdh and β2-microglobulin served as cap-translated controls (l). Cdk1nb, Xiap, Vegf and Fgf2 are IRES-translated genes presented in m. mRNA enrichment in polysome fractions is presented as percent from total RNA±SD. n=3 independent experiments. For all relevant panels, and unless otherwise stated, statistical significance was determined by one-tailed student’s t-test. Uncropped blot images are presented in Supplementary Fig. 11.

Supplementary Fig. 3 NPM1-bound snoRNAs modulate translation.

a-d, NIH/3T3 cells were stably transfected with vectors of the PiggyBac transposon system (SBI) to express mouse Snord47 and Snord52, a, Relative snoRNA abundance. Data are presented as mean±SD, n=3 independent experiments. b, Fold increase in 2′-O-Me in snoRNAs overexpressing cells. Data are presented as mean±SD, n=3 independent experiments. c, WB of p27 and XIAP in snoRNA overexpressing cells. RPL22 served as a loading control. Blot is a representative of n=3 independent experiments. d, levels of Cdkn1b and Xiap in snoRNA overexpressing cells. Data are presented as mean±SD, of n=3 independent experiments. Actb served as a control. e-g, Npm1+/+ MEFs were transfected with specific anti-snoRNAs GapmeRs and anti-GFP as control (50nM, Exiqon). e, SnoRNA levels 48hrs after transfection. Data are presented as mean±SD of n=3 independent experiments. f, Levels of the IRES-translated genes XIAP and p27 were reduced due GapmeRs transfection. Blot is a representative of n=3 independent experiments. RPL22 served as a loading control. g, Cdkn1b and Xiap transcripts levels 48 hours after GapmeRs transfection. Data are presented as mean±SD of n=3 independent experiments. h, WB analysis of FLAG expression levels of NPM1, NPMc+ and NPM-RARα in dual luciferase reporter assays. Blot is a representative of n=2 independent experiments. i, Levels of specific 2′-O-Me residues in OCI-AML3 cell line compared to OCI-AML2. Data are presented as mean±SD of n=3 independent experiments. j, SnoRNA abundance in OCI-AML2 (set as 1) and OCI-AML3 cell lines. Data are presented as mean±SD of n=3 independent experiments. For all relevant panels, and unless otherwise stated, statistical significance was determined by one-tailed student’s t-test. Uncropped blot images are presented in Supplementary Fig. 11.

Supplementary Fig. 4 NPM1-depletion and snoRNA-inactivation in K562 cell line.

a, NPM1 KD efficiency in K562/shNPM1 cells compared to scramble transduced cells (Ctrl). αTUBULIN was used as a loading control. Blot is a representative of n=3 independent experiments. b, NPM1-depletion in K562 cells led to decreased 2′-O-Me of specific residues. Data are mean±SD of n=3 independent experiments. Ctrl 2′-O-Me is the 2′-O-Me at 1804 18S rRNA. c, SnoRNA abundance in K562/shNPM1 cells. Data are presented as mean±SD of n=4 independent experiments. d, Examples of indels in specific snoRNA genes generated by CRISPR/Cas9 genome editing in K562 cells. In blue is the specific snoRNA’s wild type sequence; in red are examples of mutations in the same snoRNA induced by CRISPR/Cas9 editing. e, Analysis of control 2′-O-Me 4506 in CRISPR/Cas9 snoRNA-inactivated K562 cells. Data are presented as mean±SD of n=3 independent experiments.. f, Levels of specific 2′-O-Me residues (x-axis) mediated by only C/D box snoRNAs identified to interact with NPM1 by HITS-CLIP. Data are presented as mean±SD, n=3 independent experiments, of fold change of 2′-O-Me in K562/shNPM1+FBL cells relative to K562 cells transduced with scramble vector. Ctrl residue is the U1804 2′-O-Me modification on the 18S rRNA that is mediated by snoRD20, which was not identified to interact with NPM1 by HITS-CLIP. For all relevant panels, and unless otherwise stated, statistical significance was determined by one-tailed student’s t-test.

Supplementary Fig. 5 Characterization of acute Npm1 knock-out in hematopoietic system.

a, Schematic representations of the conditional Npm1 knockout strategy. Map of Hypomorphic allele (top), Npm1 locus after flp-mediated excision of the Neo resistance cassette (2nd) and Npm1 locus after Cre-mediated excision of the floxed exon 1-6 by crossing with an Mx1Cre transgenic mouse (bottom) are shown. Npm1 conditional allele created by insertion of a Frt-flanked Neomycin cassette (NEO) and two loxP sites in appropriate position. The NEO cassette transcriptionally interferes with the Npm1 gene expression, thus giving origin to an intermediate hypomorphic allele. The genomic sequence is depicted as a black line, with black boxes representing exons 1-11. White box represents the neomycin resistance cassette (NEO) and triangles represent the loxP site, respectively. Primers to detect Hypomorphic allele, NEO-free allele and Excised allele were also shown. b, PCR analysis of genomic DNA from the Linneg cells of NpmF/FMx1Cre+ or NpmF/FMx1Cre- mice after pIpC injection with the primers Fw3 and Rev3. The positions of amplified fragment corresponding to excised allele are indicated. n=3 independent experiments. c-d, Reduction of Npm1 after pIpC injection is confirmed at mRNA level (c) and protein level (d). Residual Npm1 expression in MNCs and CD34negKSL cells (c) and levels of NPM1 in Linpos and Linneg cells (d) are analyzed 14 days after pIpC injection. c, Data are presented as mean±SD of n=6 independent experiments. d, n=3 independent experiments. For all relevant panels, and unless otherwise stated, statistical significance was determined by one-tailed student’s t-test.

Supplementary Fig. 6 Npm1-deletion leads to bone marrow failure in adult mice.

a, Peripheral blood counts of Npm1F/FMx1Cre- and Npm1F/FMx1Cre+ mice 4 weeks after pIpC injections, demonstrating low WBC and low platelets. Data are presented as mean±SD of n=4 biologically independent animals. *P=0.019, **P=0.003. b, Npm1-deletion in adult LSK cells leads to impairment of maintenance of quiescence. % of cells in G0 phase in Npm1-deleted LSK cells 4 days after pIpC injection. Data are presented as mean±SD of n=4 biologically independent animals. c, Purified LSK cells from Npm1F/F mice were transduced with MSCV Puro-IRES-GFP (Empty) or MSCV Puro-Cre-IRES-GFP (Cre). Single sorted GFP+ cells were cultured with SCF+TPO and cell division was monitored daily. First division day was determined in 100 cells per experiment. Data are presented as mean±SD of n=3 biologically independent animals. d, Sorted LSK cells from NpmF/FMx1Cre+ or Mx1Cre- LSK were transplanted with competitor BMMNCs. 8 weeks after BMT, pIpC was injected into the recipient mice, and %G0 phase in donor-derived (Ly45.2) LSK cells was determined by PyroninY staining. Data are presented as mean±SD of n=3 biologically independent animals. e, Colony forming capacity (left) and high proliferation potential (HPP) colony forming cells (right) were determined. Data are presented as mean±SD of n=4 biologically independent animals. f, Npm1-deletion leads to significant reduction of long-term culture-initiating cells (LTC-IC). LSK cells sorted from pIpC treated NpmF/FMx1Cre- or NpmF/FMx1Cre+ or Npm+/+Mx1Cre- or Npm+/+Mx1Cre+ mice were cultured with stromal cells for 6 weeks, and tested for colony formation. Frequencies of LTC-IC in limiting-dilution assay were determined using Poisson statistics. n=4 biologically independent animals. g, Survival of Npm1/Trp53 compound mutant (n=16 biologically independent animals per group). h, Npm1/Trp53 double-null displays myeloproliferation. Two weeks after pIpC injection, myeloid compartment in bone marrow was measured by CD11b/Gr-1 positivity. Data are presented as mean±SD of n=3 biologically independent animals per group. i, Two weeks after pIpC injection, B220, CD3 and CD11b and/or Gr-1 (Myeloid) positive cells in spleen were assessed. For all relevant panels, and unless otherwise stated, statistical significance was determined by one-tailed student’s t-test.

Supplementary Fig. 7 NPMD178H is deficient in snoRNA binding.

a, Npm1−/− MEFs were stably transduced either with NPM1D178H-FLAG, or NPM1D178H-FLAG or NPM1wt-FLAG (set as 1, not shown). FLAG-IP was performed and Snord enrichment was analyzed by qPCR. Fold enrichment was calculated relative to the enrichment in NPM1wt-FLAG IP. Data are presented as mean±SD of n=3 independent experiments, b, NPM1 mutants associate as wild type NPM with FBL. Npm1−/− MEFs were stably transduced either with NPM1D178H-FLAG, NPM1D180del-FLAG or NPM1wt-FLAG. FLAG-IP was performed and co-precipitation with FBL was assessed through WB. Representative blot of n=3 independent experiments. c, snoRNA levels in BV311 and CM108 cells. Data are presented as mean±SD of n=3 independent experiments. SNORD13 served as a control (SNORDCtrl). d, WB analysis of CM108 and BV311 cells. n=2 independent experiments e, qPCR analysis of CDKN1B and XIAP levels in CM108 patient cells compared to BV311 control cells. ACTB served as a non-IRES control. Data are presented as mean±SD of n=3 independent experiments. f, 35S-Methionine incorporation assay to test for global translation comparing BV311 and CM108 cells. Shown is a representative blot of n=3 independent experiments. g, WB analysis of mTOR downstream effectors in BV311 (control) and CM108 (DC patient) cells. GAPDH served as a loading control. Shown is a representative blot of n=3 independent experiments. h, Northen blot of rRNA processing was performed using BV311 and CM108 cells. Shown is a representative blot of n=2 independent experiments. i, Traces of polysome fractionation of BV311 and CM108 cells. The traces are shown offset from one another on the arbitrary y axis (absorbance at 254 nm) for ease of visualizing the data. Shown is a representative trace of n=3 independent experiments. j, NPM1 localizes to the nucleolus in CM108 patient samples. Immunofluorescence (IF) was performed with anti-NPM1 (red), anti-FBL (as a nucleolus marker, green) and DAPI (blue). Shown are representative pictures of n=3 independent experiments. Scale bar–10μm. k, Similar levels of p53 and p21 were observed in patient cells (CM108) and control (BV311). HSP90 served as loading control. Shown is a representative blot of n=2 independent experiments. l, IF of censtrosomes (γ-tubulin) showed similar numbers in patient (CM108) and control (BV311) cells. Data are presented as mean±SD of n=5 (BV311) and n=9 (CM108) independent experiments. For all relevant panels, and unless otherwise stated, statistical significance was determined by one-tailed student’s t-test. Uncropped blot images are presented in Supplementary Fig. 11.

Supplementary Fig. 8 NPM1D180del is deficient in snoRNA binding.

a, Traces of polysome fractionation of NPM1D180del CRISPR MEFs (orange) and Trp53−/− parental MEFs (black). Shown are representative traces of n=3 independent experiments. b, Firefly/Renilla activity ratio in NPM1D180del CRISPR MEFs was calculated relative to the ratio measured in Trp53−/− parental MEFs. Data are presented as mean±SD of n=5 independent experiments. c, Western blot analysis of the levels of NPM and of the IRES translated genes XIAP and p27 in NPM1D180del CRISPR MEFs. TUBULIN served as a loading control. Blot is a representative of n=2 independent experiments. d, FBL-IP was performed using NPM1D180del CRISPR MEFs and parental Trp53−/− MEFs as control. SnoRNA enrichment was calculated relative to total input and fold enrichment in control MEFs. SnordCtrl is Snord13. Data are presented as mean±SD of n=3 independent experiments. e, Analysis of specific 2′-O-Me (x-axis) in NPM1D180del CRISPR MEFs compared to control MEFs. Data are presented as mean±SD of n=3 independent experiments. Ctrl 2′-O-Me is the 2′-O-Me at 1804 18S rRNA. f, WB validation of FLAG expression in K562/shNPM1 cells. RPL22 served as a loading control. Shown is a representative blot of n=2 independent experiments. For all relevant panels, and unless otherwise stated, statistical significance was determined by one-tailed student’s t-test. Uncropped blot images are presented in Supplementary Fig. 11.

Supplementary Fig. 9 NPM1D180del mice display features of aberrant hematopoiesis.

a, HSC compartment analysis of 2 months old NPM1D180del mice (Heterozygous, n=4, Homozygous, n=3) and control littermates (n=6). Data are presented as mean±SD b, Progenitors analysis of 2 months old NPM1D180del mice (Heterozygous, n=3, Homozygous, n=2) and control littermates (n=4). Data are presented as mean±SD c, Analysis of erythroblast percentage in the bone marrow of 2 months old NPM1D180del mice (Heterozygous, n=4, homozygous, n=3) compared to littermates mice (n=7). Data are presented as mean±SD. d, Red blood counts (RBC) and platelet counts (PLT) of heterozygous (orange circles, n=5) and homozygous (red circles, n=5) NPM1D180del mice compared to control littermates (n=12). Data are presented as mean±SD. e, Representative peripheral blood smears (out of n=9 independent samples) of NPM1D180del mice at six months of age show dysplastic RBC cells (i-control, ii-pencil cells (green arrows); iii-target cells (black arrows)), and dysplastic neutrophils (iv-control, v-vi-dysplastic cells, red arrow heads). Scale bar −10µm. f-i, An NPM1D180del mouse developed myeloid proliferative disorder exhibited splenomegaly (f) and pail bones (g). h, FACS analysis of the BM of the NPM1D180del mouse showed very low frequency of erythrocytes (TER119+ cells) and significant high percentage of myeloid cells (Gr1+CD11b+ cells), compared to littermate control. i, FACS analysis of the spleen of the NPM1D180del mouse demonstrated high percentage of TER119+ cells and of myeloid cells (Gr1+CD11b+ cells), compared to littermate control. j, western blot analysis of the IRES translated genes p27 and XIAP, in total BM cells of NPM1D180del and control littermates, demonstrating reduced protein levels in NPM1D180del cells. n=3 independent experiments performed. k, snoRNAs are expressed in similar levels in NPM1D180del mice and control littermates. Data are presented as mean±SD of n=3 independent experiments. For all relevant panels, and unless otherwise stated, statistical significance was determined by one-tailed student’s t-test. Uncropped blot images are presented in Supplementary Fig. 11.

Supplementary Fig. 10 Model of NPM1 function as a 2′-O-Me modulator.

a, NPM1 proteins assemble in pentamers and bind snoRNAs as well as FBL to facilitate snoRNA loading, proper rRNA 2′-O-methylation and normal hematopoiesis. b, NPM mutants identified in DC patients affect the complex’s ability to bind snoRNAs and hence snoRNA loading into snoRNPs is compromised. This leads to low 2′-O-Me, translational deficiencies in HSCs and to aberrant hematopoiesis. c, TCGA database reveals additional mutations in the specific NPM1 acidic repeat identified in various solid tumors.

Supplementary Fig. 11

Uncropped scans of western blot data.

Supplementary information

Supplementary Information

Supplementary Figs. 1–11

Reporting Summary

Supplementary Tables

Supplementary Tables 1 and 2

Supplementary Software

Microarray analysis code

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Nachmani, D., Bothmer, A.H., Grisendi, S. et al. Germline NPM1 mutations lead to altered rRNA 2′-O-methylation and cause dyskeratosis congenita. Nat Genet 51, 1518–1529 (2019). https://doi.org/10.1038/s41588-019-0502-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41588-019-0502-z

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing