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:

Predicting the clinical impact of human mutation with deep neural networks

An Author Correction to this article was published on 17 December 2018

This article has been updated

Abstract

Millions of human genomes and exomes have been sequenced, but their clinical applications remain limited due to the difficulty of distinguishing disease-causing mutations from benign genetic variation. Here we demonstrate that common missense variants in other primate species are largely clinically benign in human, enabling pathogenic mutations to be systematically identified by the process of elimination. Using hundreds of thousands of common variants from population sequencing of six non-human primate species, we train a deep neural network that identifies pathogenic mutations in rare disease patients with 88% accuracy and enables the discovery of 14 new candidate genes in intellectual disability at genome-wide significance. Cataloging common variation from additional primate species would improve interpretation for millions of variants of uncertain significance, further advancing the clinical utility of human genome sequencing.

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

Access options

Buy this article

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

Fig. 1: Missense/synonymous ratios across the human allele frequency spectrum.
Fig. 2: Purifying selection on missense variants identical-by-state with other species.
Fig. 3: Deep learning network for classification of missense variants.
Fig. 4: Classification accuracy within 605 DDD genes with P < 0.05.
Fig. 5: Impact of data used for training on classification accuracy.

Similar content being viewed by others

Change history

  • 17 December 2018

    In the version of this article originally published, the name of author Serafim Batzoglou was misspelled. The error has been corrected in the HTML and PDF versions of the article.

References

  1. MacArthur, D. G. et al. Guidelines for investigating causality of sequence variants in human disease. Nature 508, 469–476 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Rehm, H. L. et al. ClinGen--the Clinical Genome Resource. N. Engl. J. Med. 372, 2235–2242 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  3. Bamshad, M. J. et al. Exome sequencing as a tool for Mendelian disease gene discovery. Nat. Rev. Genet. 12, 745–755 (2011).

    CAS  PubMed  Google Scholar 

  4. Rehm, H. L. Evolving health care through personal genomics. Nat. Rev. Genet. 18, 259–267 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Richards, S. et al. Standards and guidelines for the interpretation of sequence variants: a joint consensus recommendation of the American College of Medical Genetics and Genomics and the Association for Molecular Pathology. Genet. Med. 17, 405–424 (2015).

    PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Mallick, S. et al. The Simons Genome Diversity Project: 300 genomes from 142 diverse populations. Nature 538, 201–206 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Genomes Project Consortium. et al. A global reference for human genetic variation. Nature 526, 68–74 (2015).

    Google Scholar 

  9. Liu, X., Jian, X. & Boerwinkle, E. dbNSFP: a lightweight database of human nonsynonymous SNPs and their functional predictions. Human. Mutat. 32, 894–899 (2011).

    CAS  Google Scholar 

  10. Chimpanzee Sequencing Analysis Consortium. Initial sequence of the chimpanzee genome and comparison with the human genome. Nature 437, 69–87 (2005).

    Google Scholar 

  11. Takahata, N. Allelic genealogy and human evolution. Mol. Biol. Evol. 10, 2–22 (1993).

    CAS  PubMed  Google Scholar 

  12. Asthana, S., Schmidt, S., & Sunyaev, S. A limited role for balancing selection. Trends Genet. 21, 30–32 (2005).

    CAS  PubMed  Google Scholar 

  13. Leffler, E. M. et al. Multiple instances of ancient balancing selection shared between humans and chimpanzees. Science 339, 1578–1582 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. 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 

  15. Ohta, T. Slightly deleterious mutant substitutions in evolution. Nature 246, 96–98 (1973).

    CAS  PubMed  Google Scholar 

  16. Reich, D. E. & Lander, E. S. On the allelic spectrum of human disease. Trends Genet. 17, 502–510 (2001).

    CAS  PubMed  Google Scholar 

  17. Whiffin, N. et al. Using high-resolution variant frequencies to empower clinical genome interpretation. Genet. Med. 19, 1151–1158 (2017).

    PubMed  PubMed Central  Google Scholar 

  18. Prado-Martinez, J. et al. Great ape genome diversity and population history. Nature 499, 471–475 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  19. Klein, J., Satta, Y., O’HUigin, C., & Takahata, N. The molecular descent of the major histocompatibility complex. Annu. Rev. Immunol. 11, 269–295 (1993).

    CAS  PubMed  Google Scholar 

  20. Kimura, M. The Neutral Theory of Molecular Evolution (Cambridge Univ. Press, Canbridge, UK, 1983).

  21. de Manuel, M. et al. Chimpanzee genomic diversity reveals ancient admixture with bonobos. Science 354, 477–481 (2016).

    PubMed  PubMed Central  Google Scholar 

  22. Locke, D. P. et al. Comparative and demographic analysis of orang-utan genomes. Nature 469, 529–533 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  23. Rhesus Macaque Genome Sequencing Analysis Consortium. Evolutionary and biomedical insights from the rhesus macaque genome. Science 316, 222–234 (2007).

    Google Scholar 

  24. Worley, K. C. et al. The common marmoset genome provides insight into primate biology and evolution. Nat. Genet. 46, 850–857 (2014).

    Google Scholar 

  25. Sherry, S. T. et al. dbSNP: the NCBI database of genetic variation. Nucleic Acids Res. 29, 308–311 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Schrago, C. G., & Russo, C. A. Timing the origin of New World monkeys. Mol. Biol. Evol. 20, 1620–1625 (2003).

    CAS  PubMed  Google Scholar 

  27. Landrum, M. J. et al. ClinVar: public archive of interpretations of clinically relevant variants. Nucleic Acids Res. 44, D862–868 (2016).

    CAS  PubMed  Google Scholar 

  28. Brandon, E. P., Idzerda, R. L. & McKnight, G. S. Targeting the mouse genome: a compendium of knockouts (Part II). Curr. Biol. 5, 758–765 (1995).

    CAS  PubMed  Google Scholar 

  29. Lieschke, J. G. & Currie, P. D. Animal models of human disease: zebrafish swim into view. Nat. Rev. Genet. 8, 353–367 (2007).

    CAS  PubMed  Google Scholar 

  30. Sittig, L. J. et al. Genetic background limits generalizability of genotype-phenotype relationships. Neuron 91, 1253–1259 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  31. Bazykin, G. A. et al. Extensive parallelism in protein evolution. Biol. Direct 2, 20 (2007).

    PubMed  PubMed Central  Google Scholar 

  32. Ng, P. C., & Henikoff, S. Predicting deleterious amino acid substitutions. Genome Res. 11, 863–874 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Adzhubei, I. A. et al. A method and server for predicting damaging missense mutations. Nat. Methods 7, 248–249 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Chun, S. & Fay, J. C. Identification of deleterious mutations within three human genomes. Genome Res. 19, 1553–1561 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Schwarz, J. M., Rödelsperger, C., Schuelke, M. & Seelow, D. MutationTaster evaluates disease-causing potential of sequence alterations. Nat. Methods 7, 575–576 (2010).

    CAS  PubMed  Google Scholar 

  36. Reva, B., Antipin, Y., & Sander, C. Predicting the functional impact of protein mutations: application to cancer genomics. Nucleic Acids Res. 39, e118 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  37. Dong, C. et al. Comparison and integration of deleteriousness prediction methods for nonsynonymous SNVs in whole exome sequencing studies. Hum. Mol. Genet. 24, 2125–2137 (2015).

    CAS  PubMed  Google Scholar 

  38. Carter, H., Douville, C., Stenson, P. D., Cooper, D. N., & Karchin, R. Identifying Mendelian disease genes with the variant effect scoring tool. BMC Genom. 14 Suppl 3, S3 (2013).

    Google Scholar 

  39. Choi, Y., Sims, G. E., Murphy, S., Miller, J. R., & Chan, A. P. Predicting the functional effect of amino acid substitutions and indels. PLoS One 7, e46688 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  40. Gulko, B., Hubisz, M. J., Gronau, I., & Siepel, A. A method for calculating probabilities of fitness consequences for point mutations across the human genome. Nat. Genet. 47, 276–283 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Shihab, H. A. et al. An integrative approach to predicting the functional effects of non-coding and coding sequence variation. Bioinformatics 31, 1536–1543 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Quang, D., Chen, Y., & Xie, X. DANN: a deep learning approach for annotating the pathogenicity of genetic variants. Bioinformatics 31, 761–763 (2015).

    CAS  PubMed  Google Scholar 

  43. Bell, C. J. et al. Comprehensive carrier testing for severe childhood recessive diseases by next generation sequencing. Sci. Transl. Med. 3, 65ra64 (2011).

    Google Scholar 

  44. 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 

  45. Smedley, D. et al. A whole-genome analysis framework for effective identification of pathogenic regulatory variants in mendelian disease. Am. J. Hum. Genet. 99, 595–606 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Ioannidis, N. M. et al. REVEL: an ensemble method for predicting the pathogenicity of rare missense variants. Am. J. Hum. Genet. 99, 877–885 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Jagadeesh, K. A. et al. M-CAP eliminates a majority of variants of uncertain significance in clinical exomes at high sensitivity. Nat. Genet. 48, 1581–1586 (2016).

    CAS  PubMed  Google Scholar 

  48. Grimm, D. G. The evaluation of tools used to predict the impact of missense variants is hindered by two types of circularity. Human. Mutat. 36, 513–523 (2015).

    Google Scholar 

  49. He, K., Zhang, X., Ren, S., & Sun, J. Deep residual learning for image recognition. in Proc. IEEE Conf. Comput. Vis. Pattern Recognit. 770–778 (2016).

  50. Heffernan, R. et al. Improving prediction of secondary structure, local backbone angles, and solvent accessible surface area of proteins by iterative deep learning. Sci. Rep. 5, 11476 (2015).

    PubMed  PubMed Central  Google Scholar 

  51. Wang, S., Peng, J., Ma, J. & Xu, J. Protein secondary structure prediction using deep convolutional neural fields. Sci. Rep. 6, 18962–18962 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Harpak, A., Bhaskar, A., & Pritchard, J. K. Mutation rate variation is a primary determinant of the distribution of allele frequencies in humans. PLoS Genet. 12 e1006489 (2016).

  53. Payandeh, J., Scheuer, T., Zheng, N. & Catterall, W. A. The crystal structure of a voltage-gated sodium channel. Nature 475, 353–358 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Shen, H. et al. Structure of a eukaryotic voltage-gated sodium channel at near-atomic resolution. Science 355, eaal4326 (2017).

    PubMed  Google Scholar 

  55. Nakamura, K. et al. Clinical spectrum of SCN2A mutations expanding to Ohtahara syndrome. Neurology 81, 992–998 (2013).

    CAS  PubMed  Google Scholar 

  56. Henikoff, S. & Henikoff, J. G. Amino acid substitution matrices from protein blocks. Proc. Natl. Acad. Sci. USA 89, 10915–10919 (1992).

    CAS  PubMed  PubMed Central  Google Scholar 

  57. Li, W. H., Wu, C. I. & Luo, C. C. Nonrandomness of point mutation as reflected in nucleotide substitutions in pseudogenes and its evolutionary implications. J. Molec. Evol. 21, 58–71 (1984).

    CAS  PubMed  Google Scholar 

  58. Grantham, R. Amino acid difference formula to help explain protein evolution. Science 185, 862–864 (1974).

    CAS  PubMed  Google Scholar 

  59. LeCun, Y., Bottou, L., Bengio, Y., & Haffner, P. Gradient based learning applied to document recognition. Proc. IEEE 86, 2278–2324 (1998).

    Google Scholar 

  60. Vissers, L. E., Gilissen, C., & Veltman, J. A. Genetic studies in intellectual disability and related disorders. Nat. Rev. Genet. 17, 9–18 (2016).

    CAS  PubMed  Google Scholar 

  61. Neale, B. M. et al. Patterns and rates of exonic de novo mutations in autism spectrum disorders. Nature 485, 242–245 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  62. Sanders, S. J. et al. De novo mutations revealed by whole-exome sequencing are strongly associated with autism. Nature 485, 237–241 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  63. De Rubeis, S. et al. Synaptic, transcriptional and chromatin genes disrupted in autism. Nature 515, 209–215 (2014).

    PubMed  PubMed Central  Google Scholar 

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

    Google Scholar 

  65. Deciphering Developmental Disorders Study. Prevalence and architecture of de novo mutations in developmental disorders. Nature 542, 433–438 (2017).

    Google Scholar 

  66. Iossifov, I. et al. The contribution of de novo coding mutations to autism spectrum disorder. Nature 515, 216–221 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  67. Zhu, X., Need, A. C., Petrovski, S. & Goldstein, D. B. One gene, many neuropsychiatric disorders: lessons from Mendelian diseases. Nat. Neurosci. 17, 773–781, https://doi.org/10.1038/nn.3713 (2014).

    Article  CAS  PubMed  Google Scholar 

  68. Leffler, E. M. et al. Revisiting an old riddle: what determines genetic diversity levels within species? PLoS Biol. 10, e1001388 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  69. Estrada, A. et al. Impending extinction crisis of the world’s primates: why primates matter. Sci. Adv. 3, e1600946 (2017).

    PubMed  PubMed Central  Google Scholar 

  70. Kent, W. J. et al. The human genome browser at UCSC. Genome Res. 12, 996–1006 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  71. Tyner, C. et al. The UCSC Genome Browser database: 2017 update. Nucleic Acids Res. 45, D626–D634 (2017).

    CAS  PubMed  Google Scholar 

  72. Kabsch, W., & Sander, C. Dictionary of protein secondary structure: pattern recognition of hydrogen-bonded and geometrical features. Biopolymers 22, 2577–2637 (1983).

    CAS  PubMed  Google Scholar 

  73. Joosten, R. P. et al. A series of PDB related databases for everyday needs. Nucleic Acids Res. 39, D411–419 (2011).

    CAS  PubMed  Google Scholar 

  74. He, K., Zhang, X., Ren, S., & Sun, J. Identity mappings in deep residual networks. in 14th European Conference on Computer Vision – ECCV 2016. ECCV 2016. Lecture Notes in Computer Science, vol 9908; 630–645 (Springer, Cham, Switzerland; 2016).

  75. Ionita-Laza, I., McCallum, K., Xu, B., & Buxbaum, J. D. A spectral approach integrating functional genomic annotations for coding and noncoding variants. Nat. Genet. 48, 214–220 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  76. Li, B. et al. Automated inference of molecular mechanisms of disease from amino acid substitutions. Bioinformatics 25, 2744–2750 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  77. Lu, Q. et al. A statistical framework to predict functional non-coding regions in the human genome through integrated analysis of annotation data. Sci. Rep. 5, 10576 (2015).

    PubMed  PubMed Central  Google Scholar 

  78. Shihab, H. A. et al. Predicting the functional, molecular, and phenotypic consequences of amino acid substitutions using hidden Markov models. Human. Mutat. 34, 57–65 (2013).

    CAS  Google Scholar 

  79. Davydov, E. V. et al. Identifying a high fraction of the human genome to be under selective constraint using GERP++. PLoS Comput. Biol. 6, e1001025 (2010).

    PubMed  PubMed Central  Google Scholar 

  80. Liu, X., Wu, C., Li, C., & Boerwinkle, E. dbNSFPv3.0: a one-stop database of functional predictions and annotations for human nonsynonymous and splice-site SNVs. Human. Mutat. 37, 235–241 (2016).

    Google Scholar 

  81. Jain, S., White, M., Radivojac, P. Recovering true classifier performance in positive-unlabeled learning. in Proceedings Thirty-First AAAI Conference on Artificial Intelligence. 2066–2072 (AAAI Press, San Francisco; 2017).

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

    PubMed  Google Scholar 

  83. Iossifov, I. et al. De novo gene disruptions in children on the autistic spectrum. Neuron 74, 285–299 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  84. O’Roak, B. J. et al. Sporadic autism exomes reveal a highly interconnected protein network of de novo mutations. Nature 485, 246–250 (2012).

    PubMed  PubMed Central  Google Scholar 

  85. Rauch, A. et al. Range of genetic mutations associated with severe non-syndromic sporadic intellectual disability: an exome sequencing study. Lancet 380, 1674–1682 (2012).

    CAS  PubMed  Google Scholar 

  86. Epi, K. C. et al. De novo mutations in epileptic encephalopathies. Nature 501, 217–221 (2013).

    Google Scholar 

  87. EuroEPINOMICS-RES Consortium, Epilepsy Phenome/Genome Project, Epi4K Consortium. De novo mutations in synaptic transmission genes including DNM1 cause epileptic encephalopathies. Am. J. Hum. Genet. 95, 360–370 (2014).

  88. Gilissen, C. et al. Genome sequencing identifies major causes of severe intellectual disability. Nature 511, 344–347 (2014).

    CAS  PubMed  Google Scholar 

  89. Lelieveld, S. H. et al. Meta-analysis of 2,104 trios provides support for 10 new genes for intellectual disability. Nat. Neurosci. 19, 1194–1196 (2016).

    CAS  PubMed  Google Scholar 

  90. Famiglietti, M. L. et al. Genetic variations and diseases in UniProtKB/Swiss-Prot: the ins and outs of expert manual curation. Human. Mutat. 35, 927–935 (2014).

    CAS  Google Scholar 

  91. Horaitis, O., Talbot, C. C.Jr., Phommarinh, M., Phillips, K. M., & Cotton, R. G. A database of locus-specific databases. Nat. Genet. 39, 425 (2007).

    CAS  PubMed  Google Scholar 

  92. Stenson, P. D. et al. The Human Gene Mutation Database: building a comprehensive mutation repository for clinical and molecular genetics, diagnostic testing and personalized genomic medicine. Hum. Genet. 133, 1–9 (2014).

    CAS  PubMed  Google Scholar 

Download references

Acknowledgements

The authors would like to thank J. K. Pritchard, M. E. Hurles, J. W. Belmont, and R. E. Green for insightful discussions. The authors would like to thank the Genome Aggregation Database (gnomAD) and the groups that provided exome and genome variant data to this resource. A full list of contributing groups can be found at http://gnomad.broadinstitute.org/about. The DDD study presents independent research commissioned by the Health Innovation Challenge Fund (grant number HICF-1009-003), a parallel funding partnership between Wellcome and the Department of Health, and the Wellcome Sanger Institute (grant number WT098051). The views expressed in this publication are those of the authors and not necessarily those of Wellcome or the Department of Health. The study has UK Research Ethics Committee approval (10/H0305/83, granted by the Cambridge South REC, and GEN/284/12 granted by the Republic of Ireland REC). The research team acknowledges the support of the National Institute for Health Research, through the Comprehensive Clinical Research Network. L.S., S.R.P., Y.L. and X.L. were partially supported by R01GM110240 from the National Institute of General Medical Sciences and National Science Foundation (grant number CNS-1747783, CNS-1624782, and OAC-1229576).

Author information

Authors and Affiliations

Authors

Contributions

K.K.F., L.S., H.G., S.R.P., and J.F.M. designed the study and wrote the manuscript. L.S., S.R.P., Y.L., N.F., J.H., A.D., J.S., J.X., S.B., X.L., and K.K.F. performed the deep learning analysis. H.G., J.F.M., L.S., S.R.P., J.A.K., and K.K.F. performed the genetics analysis. L.S. and H.G. are co-first authors.

Corresponding author

Correspondence to Kyle Kai-How Farh.

Ethics declarations

Competing interests

Authors with Illumina affiliations were employees of Illumina, Inc., at the time of this work.

Additional information

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

Integrated supplementary information

Supplementary Figure 1 Expected missense:synonymous ratios across the human allele frequency spectrum in the absence of purifying selection.

Shaded gray bars represent the number of synonymous variants, and dark green bars represent the number of missense variants. The dotted line shows the baseline formed by synonymous variants. Missense:synonymous ratios are indicated for each allele frequency category. The expected missense and synonymous counts in each allele frequency category were calculated by taking intronic variants from the ExAC/gnomAD dataset consisting of 123,136 exomes and using them to estimate the fraction of variants expected to fall into each of the four allele frequency categories, based on the trinucleotide context of the variant, which controls for mutational rate and GC bias in gene conversion.

Supplementary Figure 2 Missense:synonymous ratios for CpG and non-CpG variants.

a,b, Missense:synonymous ratios for CpG (a) and non-CpG (b) variants across the human allele frequency spectrum, using all variants from the ExAC/gnomAD exomes. c,d, Missense:synonymous ratios for CpG (c) and non-CpG (d) variants across the human allele frequency spectrum, restricted to only human variants that are identical by state with chimpanzee common polymorphisms.

Supplementary Figure 3 Missense:synonymous ratios of human variants identical by state with six primates.

Patterns of missense:synonymous ratios across the human allele frequency spectrum for ExAC/gnomAD variants that are identical by state with variation present in chimpanzee, bonobo, gorilla, orangutan, rhesus, and marmoset.

Supplementary Figure 4 Schematic illustration of PrimateAI, the deep learning network architecture for pathogenicity classification.

The inputs to the model include 51 amino acids (AA) of flanking sequence for both the reference sequence and the sequence with the variant substituted in, conservation represented by three 51-AA-length position-weighted matrices from primate, mammal, and vertebrate alignments, and the outputs of networks for training secondary structure and solvent accessibility (also 51 AA in length).

Supplementary Figure 5 Detailed illustration of the deep learning network architecture used for predicting secondary structure and solvent accessibility of proteins.

The input to the model is a position-weighted matrix using conservation generated by the RaptorX software (for training on Protein Data Bank sequences) or the 99-vertebrate alignments (for training and inference on human protein sequences). The output of the second to last layer, which is 51 AAs in length, becomes the input for the deep learning network for pathogenicity classification.

Supplementary Figure 6 Correcting for the effect of sequencing coverage on the ascertainment of common primate variants.

The probability of observing a given variant in a nonhuman primate species is inversely correlated with the sequencing depth at that position in the ExAC/gnomAD exome dataset. In contrast, the lower gnomAD read depth did not affect the probability of observing a common human variant at that position (>0.1% allele frequency) because the large number of human exomes sequenced makes ascertainment of common variation almost guaranteed. When picking matched variants for each of the primate variants for training the network, the probability of picking a variant was adjusted for the effects of sequencing depth, in addition to matching for trinucleotide context to control for mutational rate and gene conversion.

Supplementary Figure 7 Recognition of protein motifs by the neural network.

a, To illustrate the neural network’s recognition of protein domains, we show the average PrimateAI scores for variants at each amino acid position in three different protein domains. Top, the collagen strand of COL1A2, with glycine in a repeating GXX motif highlighted. Clinically identified mutations in collagen genes are largely due to missense mutations in glycine in GXX repeats, as these interfere with the normal assembly of collagen and exert strong dominant-negative effects. Middle, the active site of the IDS sulfatase enzyme (highlighted), which contains a cysteine at the active site that is post-translationally modified to formylglycine. Bottom, the bHLHzip domain of the MYC transcription factor. The basic domain contacts DNA via positively charged arginine and lysine residues (highlighted) that interact with the negatively charged sugar-phosphate backbone. The leucine-zipper domain consists of leucine residues spaced seven amino acids apart (highlighted), which are crucial for dimerization. b, Line plot showing the effect of perturbing each position in and around the variant on the predicted DL score for the variant. We systematically zeroed out the inputs at nearby amino acids (positions –25 to + 25) around the variant and measured the change in the neural network’s predicted pathogenicity of the variant. The plot shows the average change in predicted pathogenicity score for perturbations at each nearby amino acid position for 5,000 randomly selected variants.

Supplementary Figure 8 Correlation patterns of weights mimic BLOSUM62 and Grantham score matrices.

Correlation patterns of weights from the first three layers of the secondary-structure DL network show correlations between amino acids that are similar to BLOSUM62 and Grantham score matrices. The left heat map shows the correlation of parameter weights from the first convolutional layer following two initial upsampling layers of the secondary-structure DL network between amino acids encoded using a one-hot representation. The middle heat map shows BLOSUM62 scores between pairs of amino acids. The right heat map shows Grantham distance between amino acids. The Pearson correlation between DL weights and BLOSUM62 scores is 0.63 (P = 3.55 × 10–9). The correlation between DL weights and Grantham scores is –0.59 (P = 4.36 × 10–8). The correlation between BLOSUM62 and Grantham scores is –0.72 (P = 8.09 × 10–13).

Supplementary Figure 9 Performance evaluation of the deep learning network and other classifiers.

a, Accuracy of the deep learning network PrimateAI at predicting a benign consequence for a test set of 10,000 primate variants that were withheld from training and comparison with other classifiers, including SIFT, PolyPhen-2, CADD, REVEL, M-CAP, LRT, MutationTaster, MutationAssessor, FATHMM, PROVEAN, VEST3, MetaSVM, MetaLR, MutPred, DANN, FATHMM-MKL_coding, Eigen, GenoCanyon, integrated_fitCons, and GERP. The y axis represents the percentage of primate variants classified as benign, based on normalizing the threshold for each classifier to its 50th-percentile score using a set of 10,000 randomly selected variants that were matched to the primate variants for trinucleotide context to control for mutational rate and gene conversion. b, Comparison of the performance of the PrimateAI network in separating de novo missense variants in DDD cases versus controls, along with the 20 existing methods listed above. The y axis shows the P values of the Wilcoxon rank-sum test for each classifier. c, Comparison of the performance of the PrimateAI network in separating de novo missense variants in DDD cases versus unaffected controls within 605 disease-associated genes, with the 20 methods listed above. The y axis shows the P values of the Wilcoxon rank-sum test for each classifier.

Supplementary Figure 10 Distribution of the prediction scores of four classifiers.

Histograms of the prediction scores of four classifiers, including SIFT, PolyPhen-2, CADD, and REVEL, for de novo missense variants occurring in DDD cases versus unaffected controls, with corresponding Wilcoxon rank-sum P values.

Supplementary Figure 11 Comparing the accuracy of the PrimateAI network and other classifiers at separating pathogenic and benign variants in 605 disease-associated genes.

a, Scatterplot showing the performance of each of the classifiers on DDD cases versus controls (y axis) and benign prediction accuracy on the withheld primate dataset (x axis). b, Comparison of different classifiers at separating de novo missense variants in cases versus controls within the 605 genes, shown on a receiver operator characteristic (ROC) curve, with area under the curve (AUC) indicated for each classifier. c, Classification accuracy and AUC for the PrimateAI network and the 20 classifiers listed in Supplementary Fig. 9. The classification accuracy shown is the average of the true positive and true negative rates, using the threshold where the classifier would predict the same number of pathogenic and benign variants as expected based on the enrichment in Fig. 4a. The maximum achievable AUC for a perfect classifier is indicated with a dotted line, assuming that de novo missense variants in DDD cases are 67% pathogenic variants and 33% benign, and de novo missense variants in controls are 100% benign.

Supplementary Figure 12 Correlation between classifier performance on human expert-curated ClinVar variants and performance on empirical datasets.

a, Scatterplot showing the classification accuracy (y axis) on ClinVar variants and on the 10,000 withheld primate variants (x axis) for each of the 20 other classifiers and the PrimateAI network trained with human-only or human + primates data. Shown are the Spearman correlation coefficient rho and associated P value. To limit the evaluation to data that were not used for training the classifiers, we only used ClinVar variants that were added between January 2017 and November 2017, and excluded common human variants from ExAC/gnomAD (>0.1% allele frequency). The ClinVar classification accuracy shown is the average of the true positive and true negative rates, using the threshold where the classifier would predict the same number of pathogenic and benign variants as observed in the ClinVar dataset. b, Scatterplot showing the classification accuracy (y axis) on ClinVar variants and the DDD cases versus controls full dataset (x axis) for each of the 20 other classifiers and the PrimateAI network trained with human-only or human + primates data.

Supplementary Figure 13 Simulation showing saturation of new common missense variants discovered by increasing the size of the human cohorts surveyed.

In simulations, the genotypes of each sample were sampled according to gnomAD allele frequencies. The fraction of gnomAD common variants discovered is averaged across 100 simulations for each sample size from 10 to 100,000.

Supplementary Figure 14 Accuracy of PrimateAI across different conservation profiles in the genome.

The x axis represents the percentage alignability of the 51 AA around a sequence with the 99-vertebrate alignments. The y axis represents the classification performance of PrimateAI for variants in each of the conservation bins, benchmarked on the test dataset of 10,000 withheld primate variants.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–14, Supplementary Tables 5, 8–12, 14–17 and 19, Supplementary Data 1–4 and Supplementary Note

Reporting Summary

Supplementary Tables

Supplementary Tables 1–4, 6, 7, 13, 18, 20 and 21

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sundaram, L., Gao, H., Padigepati, S.R. et al. Predicting the clinical impact of human mutation with deep neural networks. Nat Genet 50, 1161–1170 (2018). https://doi.org/10.1038/s41588-018-0167-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41588-018-0167-z

This article is cited by

Search

Quick links

Nature Briefing AI and Robotics

Sign up for the Nature Briefing: AI and Robotics newsletter — what matters in AI and robotics research, free to your inbox weekly.

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