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:

Mutations in the histone methyltransferase gene KMT2B cause complex early-onset dystonia

A Corrigendum to this article was published on 26 May 2017

This article has been updated

Abstract

Histone lysine methylation, mediated by mixed-lineage leukemia (MLL) proteins, is now known to be critical in the regulation of gene expression, genomic stability, cell cycle and nuclear architecture. Despite MLL proteins being postulated as essential for normal development, little is known about the specific functions of the different MLL lysine methyltransferases. Here we report heterozygous variants in the gene KMT2B (also known as MLL4) in 27 unrelated individuals with a complex progressive childhood-onset dystonia, often associated with a typical facial appearance and characteristic brain magnetic resonance imaging findings. Over time, the majority of affected individuals developed prominent cervical, cranial and laryngeal dystonia. Marked clinical benefit, including the restoration of independent ambulation in some cases, was observed following deep brain stimulation (DBS). These findings highlight a clinically recognizable and potentially treatable form of genetic dystonia, demonstrating the crucial role of KMT2B in the physiological control of voluntary movement.

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

Figure 1: Molecular genetic findings in patients with KMT2B variants.
Figure 2: Clinical features of patients with KMT2B variants.
Figure 3: Radiological features of patients with KMT2B variants.
Figure 4: Comparative modeling of KMT2B protein structure.
Figure 5: KMT2B expression and effects on histone H3K4 methylation.

Similar content being viewed by others

Accession codes

Accessions

NCBI Reference Sequence

Change history

  • 20 April 2017

    Following publication of this article, the authors were asked to remove a clinical image and some video footage of one of the affected individuals. Although consent was obtained, in keeping with their ethical consent framework, the authors allow for withdrawal of consent and are carrying out the wishes of the research subjects under their consent process. This amendment has been made in the HTML and PDF versions of the article.

References

  1. Charlesworth, G., Bhatia, K.P. & Wood, N.W. The genetics of dystonia: new twists in an old tale. Brain 136, 2017–2037 (2013).

    PubMed  PubMed Central  Google Scholar 

  2. Albanese, A. et al. Phenomenology and classification of dystonia: a consensus update. Mov. Disord. 28, 863–873 (2013).

    PubMed  PubMed Central  Google Scholar 

  3. Ng, J., Papandreou, A., Heales, S.J. & Kurian, M.A. Monoamine neurotransmitter disorders—clinical advances and future perspectives. Nat. Rev. Neurol. 11, 567–584 (2015).

    CAS  PubMed  Google Scholar 

  4. Shanker, V. & Bressman, S.B. Diagnosis and management of dystonia. Continuum (Minneap. Minn.) 22 4 Movement Disorders 1227–1245 (2016).

    Google Scholar 

  5. Balint, B. & Bhatia, K.P. Isolated and combined dystonia syndromes—an update on new genes and their phenotypes. Eur. J. Neurol. 22, 610–617 (2015).

    CAS  PubMed  Google Scholar 

  6. Lin, J.P., Lumsden, D.E., Gimeno, H. & Kaminska, M. The impact and prognosis for dystonia in childhood including dystonic cerebral palsy: a clinical and demographic tertiary cohort study. J. Neurol. Neurosurg. Psychiatry 85, 1239–1244 (2014).

    PubMed  Google Scholar 

  7. Dale, R.C., Grattan-Smith, P., Nicholson, M. & Peters, G.B. Microdeletions detected using chromosome microarray in children with suspected genetic movement disorders: a single-centre study. Dev. Med. Child Neurol. 54, 618–623 (2012).

    PubMed  Google Scholar 

  8. Lek, M. et al. Analysis of protein-coding genetic variation in 60,706 humans. Nature 536, 285–291 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Kircher, M. et al. A general framework for estimating the relative pathogenicity of human genetic variants. Nat. Genet. 46, 310–315 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Laimer, J., Hofer, H., Fritz, M., Wegenkittl, S. & Lackner, P. MAESTRO—multi agent stability prediction upon point mutations. BMC Bioinformatics 16, 116 (2015).

    PubMed  PubMed Central  Google Scholar 

  11. Pires, D.E., Ascher, D.B. & Blundell, T.L. DUET: a server for predicting effects of mutations on protein stability using an integrated computational approach. Nucleic Acids Res. 42, W314–W319 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Liu, Z. et al. Structural and functional insights into the human Börjeson–Forssman–Lehmann syndrome–associated protein PHF6. J. Biol. Chem. 289, 10069–10083 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Musselman, C.A. & Kutateladze, T.G. Handpicking epigenetic marks with PHD fingers. Nucleic Acids Res. 39, 9061–9071 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Sanchez, R. & Zhou, M.M. The PHD finger: a versatile epigenome reader. Trends Biochem. Sci. 36, 364–372 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Hsieh, J.J., Ernst, P., Erdjument-Bromage, H., Tempst, P. & Korsmeyer, S.J. Proteolytic cleavage of MLL generates a complex of N- and C-terminal fragments that confers protein stability and subnuclear localization. Mol. Cell. Biol. 23, 186–194 (2003).

    CAS  PubMed  PubMed Central  Google Scholar 

  16. Pless, B. et al. The heterodimerization domains of MLL-FYRN and FYRC- are potential target structures in t(4;11) leukemia. Leukemia 25, 663–670 (2011).

    CAS  PubMed  Google Scholar 

  17. Wysocka, J. et al. WDR5 associates with histone H3 methylated at K4 and is essential for H3 K4 methylation and vertebrate development. Cell 121, 859–872 (2005).

    CAS  PubMed  Google Scholar 

  18. Song, J.J. & Kingston, R.E. WDR5 interacts with mixed lineage leukemia (MLL) protein via the histone H3–binding pocket. J. Biol. Chem. 283, 35258–35264 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Chubb, J.R. et al. Developmental timing in Dictyostelium is regulated by the Set1 histone methyltransferase. Dev. Biol. 292, 519–532 (2006).

    CAS  PubMed  Google Scholar 

  20. Malan, V. et al. 19q13.11 deletion syndrome: a novel clinically recognisable genetic condition identified by array comparative genomic hybridisation. J. Med. Genet. 46, 635–640 (2009).

    CAS  PubMed  Google Scholar 

  21. Kruer, M.C. et al. Neuroimaging features of neurodegeneration with brain iron accumulation. AJNR Am. J. Neuroradiol. 33, 407–414 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Meyer, E., Kurian, M.A. & Hayflick, S.J. Neurodegeneration with brain iron accumulation: genetic diversity and pathophysiological Mechanisms. Annu. Rev. Genomics Hum. Genet. 16, 257–279 (2015).

    CAS  PubMed  Google Scholar 

  23. Ozelius, L. & Lubarr, N. DYT1 early-onset Isolated dystonia. GeneReviews https://www.ncbi.nlm.nih.gov/books/NBK1492/ (1999).

  24. Marras, K.C. & Münchau, A. Dystonia overview. GeneReviews https://www.ncbi.nlm.nih.gov/books/NBK1155/ (2003).

  25. Kouzarides, T. Chromatin modifications and their function. Cell 128, 693–705 (2007).

    CAS  PubMed  Google Scholar 

  26. Black, J.C., Van Rechem, C. & Whetstine, J.R. Histone lysine methylation dynamics: establishment, regulation, and biological impact. Mol. Cell 48, 491–507 (2012).

    CAS  PubMed  Google Scholar 

  27. Creyghton, M.P. et al. Histone H3K27ac separates active from poised enhancers and predicts developmental state. Proc. Natl. Acad. Sci. USA 107, 21931–21936 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Wang, Y., Li, X. & Hu, H. H3K4me2 reliably defines transcription factor binding regions in different cells. Genomics 103, 222–228 (2014).

    CAS  PubMed  Google Scholar 

  29. Shao, G.B. et al. Dynamic patterns of histone H3 lysine 4 methyltransferases and demethylases during mouse preimplantation development. In Vitro Cell. Dev. Biol. Anim. 50, 603–613 (2014).

    CAS  PubMed  Google Scholar 

  30. Shen, E., Shulha, H., Weng, Z. & Akbarian, S. Regulation of histone H3K4 methylation in brain development and disease. Phil. Trans. R. Soc. Lond. B 369, 20130514 (2014).

    Google Scholar 

  31. Jones, W.D. et al. De novo mutations in MLL cause Wiedemann–Steiner syndrome. Am. J. Hum. Genet. 91, 358–364 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Kleefstra, T. et al. Disruption of an EHMT1-associated chromatin-modification module causes intellectual disability. Am. J. Hum. Genet. 91, 73–82 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Ng, S.B. et al. Exome sequencing identifies MLL2 mutations as a cause of Kabuki syndrome. Nat. Genet. 42, 790–793 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Singh, T. et al. Rare loss-of-function variants in SETD1A are associated with schizophrenia and developmental disorders. Nat. Neurosci. 19, 571–577 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Micale, L. et al. Molecular analysis, pathogenic mechanisms, and readthrough therapy on a large cohort of Kabuki syndrome patients. Hum. Mutat. 35, 841–850 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Ang, S.Y. et al. KMT2D regulates specific programs in heart development via histone H3 lysine 4 di-methylation. Development 143, 810–821 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Jakovcevski, M. et al. Neuronal Kmt2a/Mll1 histone methyltransferase is essential for prefrontal synaptic plasticity and working memory. J. Neurosci. 35, 5097–5108 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Benayoun, B.A. et al. H3K4me3 breadth is linked to cell identity and transcriptional consistency. Cell 158, 673–688 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Muramoto, T., Müller, I., Thomas, G., Melvin, A. & Chubb, J.R. Methylation of H3K4 is required for inheritance of active transcriptional states. Curr. Biol. 20, 397–406 (2010).

    CAS  PubMed  Google Scholar 

  40. Trabzuni, D. et al. Quality control parameters on a large dataset of regionally dissected human control brains for whole genome expression studies. J. Neurochem. 119, 275–282 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. 1000 Genomes Project Consortium. An integrated map of genetic variation from 1,092 human genomes. Nature 491, 56–65 (2012).

  42. UK10K Consortium. The UK10K project identifies rare variants in health and disease. Nature 526, 82–90 (2015).

  43. Raczy, C. et al. Isaac: ultra-fast whole-genome secondary analysis on Illumina sequencing platforms. Bioinformatics 29, 2041–2043 (2013).

    CAS  PubMed  Google Scholar 

  44. Li, H. & Durbin, R. Fast and accurate short read alignment with Burrows–Wheeler transform. Bioinformatics 25, 1754–1760 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Li, H. et al. The Sequence Alignment/Map format and SAMtools. Bioinformatics 25, 2078–2079 (2009).

    PubMed  PubMed Central  Google Scholar 

  46. McKenna, A. et al. The Genome Analysis Toolkit: a MapReduce framework for analyzing next-generation DNA sequencing data. Genome Res. 20, 1297–1303 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Danecek, P. et al. The variant call format and VCFtools. Bioinformatics 27, 2156–2158 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  48. McLaren, W. et al. Deriving the consequences of genomic variants with the Ensembl API and SNP Effect Predictor. Bioinformatics 26, 2069–2070 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Gahl, W.A. et al. The National Institutes of Health Undiagnosed Diseases Program: insights into rare diseases. Genet. Med. 14, 51–59 (2012).

    CAS  PubMed  Google Scholar 

  50. Adams, D.R. et al. Analysis of DNA sequence variants detected by high-throughput sequencing. Hum. Mutat. 33, 599–608 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Bone, W.P. et al. Computational evaluation of exome sequence data using human and model organism phenotypes improves diagnostic efficiency. Genet. Med. 18, 608–617 (2016).

    CAS  PubMed  Google Scholar 

  52. de Ligt, J. et al. Diagnostic exome sequencing in persons with severe intellectual disability. N. Engl. J. Med. 367, 1921–1929 (2012).

    CAS  PubMed  Google Scholar 

  53. Wang, K., Li, M. & Hakonarson, H. ANNOVAR: functional annotation of genetic variants from high-throughput sequencing data. Nucleic Acids Res. 38, e164 (2010).

    PubMed  PubMed Central  Google Scholar 

  54. Akawi, N. et al. Discovery of four recessive developmental disorders using probabilistic genotype and phenotype matching among 4,125 families. Nat. Genet. 47, 1363–1369 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Deciphering Developmental Disorders Study. Large-scale discovery of novel genetic causes of developmental disorders. Nature 519, 223–228 (2015).

  56. Untergasser, A. et al. Primer3—new capabilities and interfaces. Nucleic Acids Res. 40, e115 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Koressaar, T. & Remm, M. Enhancements and modifications of primer design program Primer3. Bioinformatics 23, 1289–1291 (2007).

    CAS  PubMed  Google Scholar 

  58. Samocha, K.E. et al. A framework for the interpretation of de novo mutation in human disease. Nat. Genet. 46, 944–950 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  59. Hyland, K. et al. Cerebrospinal fluid concentrations of pterins and metabolites of serotonin and dopamine in a pediatric reference population. Pediatr. Res. 34, 10–14 (1993).

    CAS  PubMed  Google Scholar 

  60. Ormazabal, A. et al. HPLC with electrochemical and fluorescence detection procedures for the diagnosis of inborn errors of biogenic amines and pterins. J. Neurosci. Methods 142, 153–158 (2005).

    CAS  PubMed  Google Scholar 

  61. Finn, R.D. et al. The Pfam protein families database: towards a more sustainable future. Nucleic Acids Res. 44, D1, D279–D285 (2016).

    Google Scholar 

  62. Landau, M. et al. ConSurf 2005: the projection of evolutionary conservation scores of residues on protein structures. Nucleic Acids Res. 33, W299–W302 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Söding, J., Biegert, A. & Lupas, A.N. The HHpred interactive server for protein homology detection and structure prediction. Nucleic Acids Res. 33, W244–W248 (2005).

    PubMed  PubMed Central  Google Scholar 

  64. Sali, A. & Blundell, T.L. Comparative protein modelling by satisfaction of spatial restraints. J. Mol. Biol. 234, 779–815 (1993).

    CAS  PubMed  Google Scholar 

  65. Shen, M.Y. & Sali, A. Statistical potential for assessment and prediction of protein structures. Protein Sci. 15, 2507–2524 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  66. Pettersen, E.F. et al. UCSF Chimera—a visualization system for exploratory research and analysis. J. Comput. Chem. 25, 1605–1612 (2004).

    CAS  PubMed  Google Scholar 

  67. Shechter, D., Dormann, H.L., Allis, C.D. & Hake, S.B. Extraction, purification and analysis of histones. Nat. Protoc. 2, 1445–1457 (2007).

    CAS  PubMed  Google Scholar 

  68. Fey, P., Kowal, A.S., Gaudet, P., Pilcher, K.E. & Chisholm, R.L. Protocols for growth and development of Dictyostelium discoideum. Nat. Protoc. 2, 1307–1316 (2007).

    CAS  PubMed  Google Scholar 

  69. Fey, P., Dodson, R.J., Basu, S. & Chisholm, R.L. One stop shop for everything Dictyostelium: dictyBase and the Dicty Stock Center in 2012. Methods Mol. Biol. 983, 59–92 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  70. Faix, J., Gerisch, G. & Noegel, A.A. Overexpression of the csA cell adhesion molecule under its own cAMP-regulated promoter impairs morphogenesis in Dictyostelium. J. Cell Sci. 102, 203–214 (1992).

    CAS  PubMed  Google Scholar 

  71. Fusaki, N., Ban, H., Nishiyama, A., Saeki, K. & Hasegawa, M. Efficient induction of transgene-free human pluripotent stem cells using a vector based on Sendai virus, an RNA virus that does not integrate into the host genome. Proc. Jpn. Acad., Ser. B, Phys. Biol. Sci. 85, 348–362 (2009).

    CAS  Google Scholar 

  72. Hartfield, E.M. et al. Physiological characterisation of human iPS-derived dopaminergic neurons. PLoS One 9, e87388 (2014).

    PubMed  PubMed Central  Google Scholar 

  73. Kirkeby, A. et al. Generation of regionally specified neural progenitors and functional neurons from human embryonic stem cells under defined conditions. Cell Rep. 1, 703–714 (2012).

    CAS  PubMed  Google Scholar 

  74. Frost, J.M. et al. Evaluation of allelic expression of imprinted genes in adult human blood. PLoS One 5, e13556 (2010).

    PubMed  PubMed Central  Google Scholar 

  75. Livak, K.J. & Schmittgen, T.D. Analysis of relative gene expression data using real-time quantitative PCR and the 2−ΔΔCT method. Methods 25, 402–408 (2001).

    CAS  PubMed  Google Scholar 

  76. Schindelin, J. et al. Fiji: an open-source platform for biological-image analysis. Nat. Methods 9, 676–682 (2012).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

We thank all our patients and their families for taking part in this study and encouraging international collaboration to seek out similar cases. Thanks are given to M. Ishida (GOS-ICH) for kindly providing the fetal cDNA, K. Tuschl (GOS-ICH) for kindly providing the human cDNA panel, L. Bassioni (Great Ormond Street Hospital, GOSH) for kindly selecting DaTSCAN images for the supplementary manuscript, M. Adams (National Hospital for Neurology and Neurosurgery, NHNN) for reviewing the imaging of the patients at NHNN and R. Meijer for helping with the sequencing analysis at the Department of Human Genetics (Nijmegen). We thank G. Moore (GOS-ICH) and P. Stanier (GOS-ICH) for proofreading the manuscript. Many thanks are given to A. Panahian-Jand (GOSH) for excellent administrative support. M.A.K. has a Wellcome Intermediate Clinical Fellowship (WT098524MA). E.M. and M.A.K. received funding from the Rosetrees Trust, the Great Ormond Street Hospital Children's Charity and the Gracious Heart Foundation. N.E.M. receives support from the UK Department of Health's NIHR Biomedical Research Centers funding streams. A. Papandreou has a joint Action Medical Research/British Paediatric Neurology Association Research Training Fellowship. J. Ng has an MRC Research Training Fellowship. A.N. has an Action Medical Research Training Fellowship. H.B.-P. has a DBS training travel grant from the Daniel Turnberg Trust Fund. H.H. is funded by the MRC and Wellcome Trust (Synaptopathies award). D.A. is supported by the Prusiner-Abramsky Award. H.P. has received grant support from the Dystonia Society (UK). K.J.P. has an Academy of Medical Sciences Clinical Starter Grant. B.P.-D. received funding from grants 20143130-La Marató de TV3 and PI15/00287-Ministerio Español de Economia y Competitividad. J.-P.L. has been supported by Guy's and St Thomas' Charity New Services and Innovation Grant G060708, the Dystonia Society (UK), grants 01/2011 and 07/2013 and an Action Medical Research, AMR-GN2097. This research was supported by the NIHR Biomedical Research Centre at Great Ormond Street Hospital for Children NHS Foundation Trust, University College London and University of Cambridge and by funding from the NIHR for the BioResource for Rare Diseases (grant RG65966). This study makes use of data generated by the DECIPHER community. A full list of centers contributing to the generation of the data is available from http://decipher.sanger.ac.uk/ and via e-mail from decipher@sanger.ac.uk. Funding for the project was provided by the Wellcome Trust for UK10K (WT091310) and the DDD study. The DDD study presents independent research commissioned by the Health Innovation Challenge Fund (grant HICF-1009-003); see http://www.ddduk.org/access.html for full acknowledgment. This work was supported in part by the Intramural Research Program of the National Human Genome Research Institute and the Common Fund, NIH Office of the Director. This work was supported in part by the German Ministry of Research and Education (grants 01GS08160 and 01GS08167; German Mental Retardation Network) as part of the National Genome Research Network to A.R. and D.W., and by the Deutsche Forschungsgemeinschaft (AB393/2-2) to A.R. Brain expression data were provided by the UK Human Brain Expression Consortium (UKBEC), which comprises John A. Hardy, Mina Ryten, Michael Weale, Daniah Trabzuni, Adaikalavan Ramasamy, Colin Smith and Robert Walker, affiliated with the UCL Institute of Neurology (J.A.H., M.R. and D.T.), King's College London (M.R., M.W. and A.R.) and the University of Edinburgh (C.S. and R.W.).

Author information

Authors and Affiliations

Authors

Consortia

Contributions

E.M., K.J.C., J.M.E.N., J.R.C., F.L.R. and M.A.K. conceived and designed experiments. J.R., N.E.M., A. Papandreou, J. Ng, H.B.-P., M.A.W., D.A., A. Barnicoat, H.B., S. Bhate, N.D., N.F., N.G., A.H., H.H., J.A.H., Z.I., M.K., P.L., D.L., S. McKee, S. Misra, S.S.M., V.N., J. Nicolai, M.N., H.P., K.J.P., G.B.P., P.P., M.S.R., P.R., R.S., M. Sinnema, M. Smith, P.T., S.M.W., D.W., B.T.W., G.W., the UK10K Consortium, the DDD study, the NIHRBR-RD Consortium, L.J.C., B.P.-D., J.-P.L., A.R., W.A.G., C.T., K.P.B., N.W.W., E.-J.K., P.G., R.C.D., F.L.R. and M.A.K. ascertained patients and contributed clinical information, photographs, videos and neuroimaging studies. M.A.K. performed phenotypic characterization of all patients. W.K.C. and M.A.K. reviewed patient neuroimaging. A. Papandreou and M.A.K. edited patient videos. A. Boys, C.W. and D.M. undertook chromosomal microarray analysis. E.M., K.J.C., D.G., N.E.M., S.W., A. Pittman, the UK10K Consortium, the DDD Study, the NIHRBR-RD Consortium, A.R., W.A.G., C.T., E.-J.K. and M.A.K. carried out whole-exome and whole-genome sequencing analysis. E.M. and A.N. performed variant validation by direct Sanger sequencing. K.J.C. performed enrichment analysis (and corresponding statistical analysis). S.P. and S.J.H.H. analyzed CSF neurotransmitters. A.P.J. and M.T. undertook comparative homology modeling. J.M.E.N. and J.R.C. undertook the histone methylation assays (and corresponding statistical analysis) and cloning of the Set1 point substitution in Dictyostelium. S. Barral generated dopaminergic neurons, collected RNA and cDNA samples, and undertook qRT–PCR experiments. E.M. maintained fibroblast cultures, collected RNA, cDNA and protein samples, and performed fibroblast immunoblotting analysis (and corresponding statistical analysis) and CSF immunoblotting (and corresponding statistical analysis). J. Ng carried out CSF immunoblotting analysis. P.G. and F.L.R. contributed critical suggestions for experimental work. E.M. and M.A.K. wrote the manuscript. K.J.C., J.R., J.M.E.N., D.G., A.P.J., N.E.M., A.R., W.A.G., C.T., E.-J.K., W.K.C., M.T., J.R.C. and F.L.R. contributed written sections for the manuscript. M.A.K. oversaw the overall project. All authors critically reviewed manuscript.

Corresponding author

Correspondence to Manju A Kurian.

Ethics declarations

Competing interests

H.P. has unrestricted support for educational activity from Medtronic.

Additional information

A list of members and affiliations appears in the Supplementary Note.

A list of members and affiliations appears in the Supplementary Note.

A list of members and affiliations appears in the Supplementary Note.

Integrated supplementary information

Supplementary Figure 1 Schematic representation of the cohorts screened for KMT2B variants.

Thirty-four patients with childhood-onset dystonia were identified in the GOS-ICH Dystonia Cohort for molecular genetic investigations. Chromosomal microarray, whole-exome sequencing and whole-genome sequencing were undertaken. Collaborations with other centers facilitated identification of further individuals with KMT2B variants.

Supplementary Figure 2 Sanger confirmation and segregation of the identified KMT2B variants.

Sequence chromatograms of patients and their parents (where available) are shown. For patients 11–19, 21, 23–25 and 26b, the sequence chromatograms show the de novo status of the respective KMT2B variant. For patients 22, 26a and 27, the sequencing chromatograms show that the respective KMT2B variant is maternally inherited.

Supplementary Figure 3 Effect of age on clinical presentation.

Age of clinical presentation for patients with KMT2B dystonia. Twenty-seven patients with variants in KMT2B were identified. Seven patients had missense variants (black squares), and the remainder had microdeletions (Del), frameshift mutations (Fs), nonsense and splice-site mutations (black circles). The age of clinical presentation for those with missense variants (n = 7; mean = 6.4 ± 2.07 (s.d.) years) was significantly higher than for those with mutations predicted to truncate or elongate the protein (n = 19; mean = 4.1 ± 2.18 (s.d.) years). Data were analyzed by two-tailed unpaired t test: *P = 0.0223 (t = 2.443, d.f. = 24). No significant difference in variances between the groups was detected by F test. Individual data points are plotted with the center bar showing the mean and error bars showing s.d.

Supplementary Figure 4 Additional clinical features in KMT2B dystonia.

(a,b) Examples of handwriting from patient 11, age 7 years, showing deterioration over a 4-week period, with reduced legibility and increased letter size due to worsening upper-limb dystonia. (c) Patient 3, age 8 years, with evidence of scalp cutis aplasia.

Supplementary Figure 5 Effect of age on neuroradiological findings.

(a) Age of neuroimaging for patients with KMT2B dystonia. Twenty-two of 27 patients had neuroimaging, specifically T2/T2*, diffusion and susceptibility weighted sequences of sufficient quality for assessment of hypointensity in the globus pallidus (Supplementary Table 7). The mean age of 17 patients with an abnormal scan (n = 21; mean = 11.7 ± 4.31 (s.d.) years; black circles) was significantly lower than that of the 5 patients with normal scans (n = 7; 19.0 ± 11.12 (s.d.) years; black squares). Data were analyzed by two-tailed unpaired t test: *P = 0.0167 (t = 2.559, d.f. = 26). A significant difference in variances between the groups was detected by F test (**P = 0.0011). Individual data points are plotted with the center bar showing the mean and error bars showing s.d. (b,c) Serial magnetic resonance imaging (MRI) in patient 22. MRI with diffusion imaging data sets with b value of zero in patient 22 at 13 years, 1 month (b) and 17 years (c) showing reduction in the hypointense appearance of the globus pallidus over time. (d) Normal DaTSCAN imaging on patient 14, age 18 years.

Supplementary Figure 6 Uncut agarose gel of PCR analysis for expression of KMT2B.

The top row of this uncut agarose gel (relating to Fig. 5a) shows KMT2B-specific PCR bands of human fetal and adult cDNA samples amplified by the intron-spanning primer pair KMT2B-ex29/30-F and KMT2B-ex29/30-R (Supplementary Table 9). The bottom row displays the GAPDH-specific PCR bands for the same cDNA samples amplified by the intron-spanning primer pair GAPDH-F and GAPDH-R (Supplementary Table 9). The last sample in both rows represents a negative control in which no DNA was included in the PCR reaction.

Supplementary Figure 7 Histone methylation assays.

(a) Example immunoblots of methylated and pan histone H3 extracted from control and patient-derived fibroblasts (relating to Fig. 5d). (b) Example immunoblots of H3K4me3, pan histone H3 and GFP in Dictyostelium cell lysates (relating to Fig. 5e). (c) Quantification of anti-GFP immunoblotting, showing expression of GFP-DdSet1(I1447T) as a percentage of GFP-DdSet1 expression. Individual data points are plotted with the center bar showing the mean and error bars showing s.d. Three independently transformed GFP-DdSet1(I1447T) cell lines were tested. n = 3 technical replicates prepared and blotted on different days (normalized expression level (by anti-GFP): set1 GFP-DdSet1(I1447T) 1 = 34.39 ± 19.52 (s.d.), set1 GFP-DdSet1(I1447T) 2 = 68.32 ± 9.191 (s.d.), set1 GFP-DdSet1(I1447T) 3 = 52 ± 5.575 (s.d.)). Dilutions of GFP-DdSet1-containing lysates were tested to ensure that quantification by densitometry was conducted in the linear range of assay sensitivity.

Supplementary Figure 8 Conservation of the amino acids altered by the identified KMT2B missense variants.

(af) The KMT2B amino acid sequences from seven different species and the human paralog KMT2A (another member of the MLL protein family) are aligned. (a) Additionally, the amino acid sequence of the human PHF6 protein, used to model the PHD-like domain, is shown. Gly1652 (in red) is highly conserved in all listed amino acid sequences, while the Phe1662 residue (in red) is either conserved or tolerates replacement by the similar amino acid tyrosine (Supplementary Fig. 12). (b) Arg1705 (in red) is conserved to zebrafish and in human KMT2A. (c) Arg1762 is fully conserved throughout the species. (d) Leu1781 (in red) is conserved in all listed mammalian homologs of KMT2B. (e) Arg2517 (in red) is conserved to zebrafish and in human KMT2A. (f) Additionally, the amino acid sequences of SET domain–containing proteins (human SETD1A and SETD1B, yeast Set1 and Dictyostelium discoideum Set1) are shown. Ile2674 (in red) is highly conserved in all listed amino acid sequences. Human KMT2B (NP_055542.1); dog KMT2B (XP_003432729.3); cow KMT2B (XP_003587289.1); mouse Kmt2B (NP_083550.2); zebrafish kmt2ba (XP_689347.5); zebrafish kmt2bb (XP_002664717.2); fruit fly trx (NP_599109.1); human KMT2A (NP_001184033.1); human PHF6 (NP_115834.1); yeast Set1 (NP_011987.1); Dicty Set1 (XP_636258.1); human SETD1A (NP_055527.1); human SETD1B (NP_055863.1).

Supplementary Figure 9 H3K4 methylation profiles.

(ae) The publically available ENCODE (Encyclopedia of DNA Elements) data (Nucleic Acids Res. 41, D56–D63, 2013) on the UCSC genome browser (https://genome.ucsc.edu/) were interrogated for the histone H3 lysine 4 trimethylation (H3K4me3) pattern in different cell lines. The layered H3K4me3 lane includes data from seven cell lines (GM12878, H1-hESC, HSMM, HUVEC, K562, NHEK and NHLF). (a) The H3K4 methylation profile from THAP1 showed a very strong, sharp methylation peak around exon 1 in all investigated cell lines. (b) A similar but slightly less sharp H3K4 methylation peak around exon 1 of TOR1A was observed. (c) The identified weaker H3K4 methylation in GNAL did not coincide with exon 1 and was located in a large intron. (d,e) For the two investigated markers of the dopaminergic system, D2R and TH, no H3K4 methylation peak was seen. NH-A, cell line from brain tissue containing astrocytes; NHDF-Ad, adult dermal fibroblast cell line based on skin tissue; NHEK, cell line of epidermal keratinocytes of skin tissue.

Supplementary Figure 10 Box plots of THAP1 mRNA expression levels in ten adult brain regions.

Expression levels are based on exon array experiments and are plotted on a log2 scale (y axis). As previously described, this data set was generated using Affymetrix Exon 1.0 ST arrays and brain tissue originating from 134 control individuals, collected by the Medical Research Council (MRC) Sudden Death Brain and Tissue Bank, Edinburgh, UK, and the Sun Health Research Institute (SHRI), an affiliate of Sun Health Corporation, USA (J. Neurochem. 119, 275–282, 2011). This plot shows that THAP1 is ubiquitously expressed across all ten brain regions analyzed; as for KMT2B, expression is higher in the cerebellum than in any other region. Putamen (PUTM), frontal cortex (FCTX), temporal cortex (TCTX), occipital cortex (OCTX), hippocampus (HIPP), substantia nigra (SNIG), medulla (specifically inferior olivary nucleus, MEDU), intralobular white matter (WHMT), thalamus (THAL), and cerebellar cortex (CRBL). N indicates the number of brain samples analyzed to generate the results for each brain region. Ranges extend from the box to 1.5 times the interquartile range.

Supplementary Figure 11 Functional investigation of the downstream effects of mutations in KMT2B.

(ac) Investigation of downstream effects in fibroblast cell culture samples (patients, n = 4; controls, n = 2). (a) Quantitative RT–PCR of THAP1 and TOR1A indicates that patients have a reduction of THAP1 and, to a lesser extent, of TOR1A transcripts in comparison to controls (THAP1: controls = 1.04 ± 0.40 (s.d.); patients = 0.55 ± 0.05 (s.d.); TOR1A: controls = 1.00 ± 0.05 (s.d.); patients = 0.79 ± 0.06 (s.d.)). Data were analyzed by two-tailed unpaired t test (THAP1: *P = 0.0498 (t = 2.780, d.f. = 4), TOR1A: *P = 0.0140 (t = 4.170, d.f. = 4)). While a significant difference in variances between the groups was detected by F test (**P = 0.0066) for THAP1, no significant difference in variances between the groups was detected for TOR1A. (b) Immunoblotting studies (for every fibroblast line, n = 3 protein samples were collected and analyzed) indicate a significant reduction in THAP1 levels for patients when compared to controls (controls = 1.57 ± 0.25 (s.d.); patient 2 = 0.83 ± 0.06 (s.d.); patient 13 = 0.77 ± 0.17 (s.d.); patient 14 = 0.53 ± 0.04 (s.d.); patient 16 = 0.54 ± 0.06 (s.d.)). Data were analyzed by one-way ANOVA: ***P < 0.0001 (F = 28.46, d.f. = 4). (c) Immunoblotting studies indicate a statistically reduced level of torsin-1A in patient 14 when compared to controls (controls = 1.10 ± 0.14 (s.d.); patient 14 = 0.64 ± 0.05 (s.d.)) but not for the other patients (patient 2 = 1.09 ± 0.27 (s.d.); patient 13 = 1.39 ± 0.18 (s.d.); patient 16 = 1.13 ± 0.29 (s.d.)). Data were analyzed by one-way ANOVA: **P = 0.0066 (F = 5.800, d.f. = 4). (d) CSF immunoblotting studies on patients 2 and 16 (n = 2 patient CSF samples) show markedly reduced levels of D2R and increased levels of TH when compared to control CSF (n = 4 control CSF samples) (D2R: controls = 1.09 ± 0.21 (s.d.); patients = 0.64 ± 0.02 (s.d.); TH: controls = 0.52 ± 0.08 (s.d.); patients = 0.90 ± 0.01 (s.d.)). Data were analyzed by two-tailed unpaired t test: D2R: *P = 0.0471 (t = 2.835, d.f. = 4); TH: **P = 0.0036 (t = 6.147, d.f. = 4). No significant difference in variances between the groups was detected by F test. CSF immunoblotting was undertaken once owing to limited volume of CSF. Individual data points are plotted with the center bar showing the mean and error bars showing s.d.

Source data

Supplementary Figure 12 Comparative modeling of p.Phe1662Tyr indicates a tolerated substitution.

In fruit flies and zebrafish, the phenylalanine is replaced by a tyrosine at position 1662 of KMT2B. Based on the similar side chains for these amino acids, tyrosine can be accommodated instead of phenylalanine at position 1662 in KMT2B without disruption of the structure of the protein. Furthermore, the different prediction programs propose this substitution to be non-deleterious (PolyPhen-2, benign with a score of 0.005; SIFT, tolerated with a score of 1.000; PROVEAN, neutral with a score of 1.16).

Supplementary Figure 13 Exclusion of KMT2B imprinting.

(a,b) The chromosomal microarray of patients 2 and 10 harboring KMT2B microdeletions and their parents indicated that the inherited SNPs were only of maternal origin, implying that the paternal KMT2B allele was deleted. While in at least these two cases the KMT2B mutation occurs on the paternal allele, we have shown that KMT2B changes in patients 22, 26a and 27 are maternally inherited. This proves that KMT2B variants are not restricted to either the maternal or parental allele. (ce) Sanger sequencing of cDNA samples from control fibroblasts, patient fibroblasts and generated dopaminergic neurons showed biallelic expression of the common KMT2B SNP rs231591 (c.7091A>G). (f) Although there are nine imprinted regions reported in the literature (http://igc.otago.ac.nz/home.html), none of them coincide with the chromosomal location of KMT2B (hg19: chromosome 19: 36,208,921–36,229,779).

Supplementary Figure 14 Evaluation of the differentiation efficiency of midbrain dopaminergic neurons.

Immunofluorescence of generated midbrain dopaminergic neurons at day 50 of differentiation using MAP2 (green) as a neuronal marker and TH (red) as a specific dopaminergic marker showed cells with colocalization (merge) of both markers indicating the successful generation of midbrain dopaminergic neurons. Nuclei were stained with DAPI (blue). Quantification of the two markers from three independent differentiations (n = 3; 600 cells per experiment) yielded 50.3 ± 17.2% of cells as MAP2 positive, of which 39.1 ± 11.0% were also TH positive.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–14, Supplementary Tables 1–11 and Supplementary Note. (PDF 4367 kb)

Supplementary Video 1

Lower-limb dystonia and gait disturbance in patients with KMT2B variants. (MPG 9702 kb)

Supplementary Video 2

Upper-limb dystonia in patients with KMT2B variants. (MPG 7260 kb)

Supplementary Video 3

Progression to generalized dystonia in patients with KMT2B variants. (MPG 9312 kb)

Supplementary Video 4

Cranial, cervical and laryngeal features in patients with KMT2B variants. (MPG 19576 kb)

Supplementary Video 5

Motor overflow secondary to dystonia in patients with KMT2B variants. (MPG 6358 kb)

Supplementary Video 6

Dystonic crisis in a patient with a KMT2B mutation. (MPG 3298 kb)

Supplementary Video 7

Myoclonus dystonia in a patient with a KMT2B mutation (MPG 2234 kb)

Supplementary Video 8

Response to deep brain stimulation in patient 9. (MPG 1946 kb)

Supplementary Video 9

Response to deep brain stimulation in patient 17. (MPG 3862 kb)

Supplementary Video 10

Response to deep brain stimulation in patient 21. (MPG 3402 kb)

Supplementary Video 11

Response to deep brain stimulation in patient 22. (MPG 10030 kb)

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Meyer, E., Carss, K., Rankin, J. et al. Mutations in the histone methyltransferase gene KMT2B cause complex early-onset dystonia. Nat Genet 49, 223–237 (2017). https://doi.org/10.1038/ng.3740

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/ng.3740

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