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:

The complete sequence of a human Y chromosome

Abstract

The human Y chromosome has been notoriously difficult to sequence and assemble because of its complex repeat structure that includes long palindromes, tandem repeats and segmental duplications1,2,3. As a result, more than half of the Y chromosome is missing from the GRCh38 reference sequence and it remains the last human chromosome to be finished4,5. Here, the Telomere-to-Telomere (T2T) consortium presents the complete 62,460,029-base-pair sequence of a human Y chromosome from the HG002 genome (T2T-Y) that corrects multiple errors in GRCh38-Y and adds over 30 million base pairs of sequence to the reference, showing the complete ampliconic structures of gene families TSPY, DAZ and RBMY; 41 additional protein-coding genes, mostly from the TSPY family; and an alternating pattern of human satellite 1 and 3 blocks in the heterochromatic Yq12 region. We have combined T2T-Y with a previous assembly of the CHM13 genome4 and mapped available population variation, clinical variants and functional genomics data to produce a complete and comprehensive reference sequence for all 24 human chromosomes.

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: Structure of a complete Y chromosome.
Fig. 2: Ampliconic genes forming composite repeats.
Fig. 3: Structure of the T2T-Y centromere.
Fig. 4: Comparison of the palindromic structure of the P1–P3 region.
Fig. 5: Heterochromatic region of Yq12.
Fig. 6: Short-read mappability and variant-calling improvements on T2T-Y.

Similar content being viewed by others

Data availability

The T2T-CHM13v2.0 (T2T-CHM13+Y) assembly, reference analysis set, complete list of resources—including gene annotation, repeat annotation, epigenetic profiles, variant-calling results from 1KGP and SGDP, gnomAD, ClinVar, GWAS and dbSNP datasets—are available for download at https://github.com/marbl/CHM13. The assembly is also available from NCBI and EBI with GenBank accession GCA_009914755.4. Annotation and associated resources are also browsable as ‘hs1’ from the UCSC Genome Browser (http://genome.ucsc.edu/cgi-bin/hgTracks?db=hub_3671779_hs1), the Ensembl Genome Browser (https://projects.ensembl.org/hprc/) (assembly name T2T-CHM13v2.0) and NCBI data-hub (https://www.ncbi.nlm.nih.gov/data-hub/genome/GCF_009914755.1/). Potential assembly issues are listed and can be tracked at https://github.com/marbl/CHM13-issues. 1KGP and SGDP short-read alignments and variant calls are available within AnVIL at https://anvil.terra.bio/#workspaces/anvil-datastorage/AnVIL_T2T_CHRY. Original data from the Gerton lab underlying this manuscript can be accessed from the Stowers Original Data Repository at http://www.stowers.org/research/publications/libpb-2358. Sequencing data used in this study are listed in Supplementary Table 1.

Code availability

Custom codes developed for data analysis and visualization are available at https://github.com/arangrhie/T2T-HG002Y, https://github.com/snurk/sg_sandbox and https://github.com/schatzlab/t2t-chm13-chry and are deposited with Zenodo159. Software and parameters used are stated in the Supplementary Methods with further details.

References

  1. Skaletsky, H. et al. The male-specific region of the human Y chromosome is a mosaic of discrete sequence classes. Nature 423, 825–837 (2003).

    Article  ADS  CAS  PubMed  Google Scholar 

  2. Miga, K. H. et al. Centromere reference models for human chromosomes X and Y satellite arrays. Genome Res. 24, 697–707 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Vollger, M. R. et al. Segmental duplications and their variation in a complete human genome. Science 376, eabj6965 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Nurk, S. et al. The complete sequence of a human genome. Science 376, 44–53 (2022).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  5. Schneider, V. A. et al. Evaluation of GRCh38 and de novo haploid genome assemblies demonstrates the enduring quality of the reference assembly. Genome Res. 27, 849–864 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Gustafson, M. L. & Donahoe, P. K. Male sex determination: current concepts of male sexual differentiation. Annu. Rev. Med. 45, 505–524 (1994).

    Article  CAS  PubMed  Google Scholar 

  7. Vog, P. H. et al. Human Y chromosome azoospermia factors (AZF) mapped to different subregions in Yq11. Hum. Mol. Genet. 5, 933–943 (1996).

    Article  Google Scholar 

  8. Miga, K. H. et al. Telomere-to-telomere assembly of a complete human X chromosome. Nature 585, 79–84 (2020).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  9. Logsdon, G. A. et al. The structure, function and evolution of a complete human chromosome 8. Nature 593, 101–107 (2021).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  10. Wenger, A. M. et al. Accurate circular consensus long-read sequencing improves variant detection and assembly of a human genome. Nat. Biotechnol. 37, 1155–1162 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Jain, M. et al. Nanopore sequencing and assembly of a human genome with ultra-long reads. Nat. Biotechnol. 36, 338–345 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  12. Nurk, S. et al. HiCanu: accurate assembly of segmental duplications, satellites, and allelic variants from high-fidelity long reads. Genome Res. 30, 1291–1305 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Rautiainen, M. & Marschall, T. GraphAligner: rapid and versatile sequence-to-graph alignment. Genome Biol. 21, 253 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  14. Formenti, G. et al. Merfin: improved variant filtering, assembly evaluation and polishing via k-mer validation. Nat. Methods 19, 696–704 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Kirsche, M. et al. Jasmine and Iris: population-scale structural variant comparison and analysis. Nat. Methods 20, 408–417 (2023).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Jain, C., Rhie, A., Hansen, N. F., Koren, S. & Phillippy, A. M. Long-read mapping to repetitive reference sequences using Winnowmap2. Nat. Methods 19, 705–710 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Mc Cartney, A. M. et al. Chasing perfection: validation and polishing strategies for telomere-to-telomere genome assemblies. Nat. Methods 19, 687–695 (2022).

    Article  Google Scholar 

  18. Rhie, A., Walenz, B. P., Koren, S. & Phillippy, A. M. Merqury: reference-free quality, completeness, and phasing assessment for genome assemblies. Genome Biol. 21, 245 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Wang, T. et al. The Human Pangenome Project: a global resource to map genomic diversity. Nature 604, 437–446 (2022).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  20. Jarvis, E. D. et al. Semi-automated assembly of high-quality diploid human reference genomes. Nature 611, 519–531 (2022).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  21. Shumate, A. et al. Assembly and annotation of an Ashkenazi human reference genome. Genome Biol. 21, 129 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Zook, J. M. et al. Extensive sequencing of seven human genomes to characterize benchmark reference materials. Sci. Data 3, 160025 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Landrum, M. J. et al. ClinVar: improvements to accessing data. Nucleic Acids Res. 48, D835–D844 (2020).

    Article  CAS  PubMed  Google Scholar 

  24. Welter, D. et al. The NHGRI GWAS Catalog, a curated resource of SNP-trait associations. Nucleic Acids Res. 42, D1001–D1006 (2014).

    Article  CAS  PubMed  Google Scholar 

  25. Smigielski, E. M., Sirotkin, K., Ward, M. & Sherry, S. T. dbSNP: a database of single nucleotide polymorphisms. Nucleic Acids Res. 28, 352–355 (2000).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Karczewski, K. J. et al. The mutational constraint spectrum quantified from variation in 141,456 humans. Nature 581, 434–443 (2020).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  27. Byrska-Bishop, M. et al. High-coverage whole-genome sequencing of the expanded 1000 Genomes Project cohort including 602 trios. Cell 185, 3426–3440 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  29. Dunham, I. et al. An integrated encyclopedia of DNA elements in the human genome. Nature 489, 57–74 (2012).

    Article  ADS  CAS  Google Scholar 

  30. Ebert, P. et al. Haplotype-resolved diverse human genomes and integrated analysis of structural variation. Science 372, eabf7117 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  31. Sanders, A. D. et al. Single-cell analysis of structural variations and complex rearrangements with tri-channel processing. Nat. Biotechnol. 38, 343–354 (2020).

    Article  CAS  PubMed  Google Scholar 

  32. Hallast, P. et al. Assembly of 43 human Y chromosomes reveals extensive complexity and variation. Nature https://doi.org/10.1038/s41586-023-06425-6 (2023).

  33. Hammer, M. F. et al. Extended Y chromosome haplotypes resolve multiple and unique lineages of the Jewish priesthood. Hum. Genet. 126, 707 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  34. Poznik, G. D. et al. Punctuated bursts in human male demography inferred from 1,244 worldwide Y-chromosome sequences. Nat. Genet. 48, 593–599 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Vegesna, R., Tomaszkiewicz, M., Medvedev, P. & Makova, K. D. Dosage regulation, and variation in gene expression and copy number of human Y chromosome ampliconic genes. PLoS Genet. 15, e1008369 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. NCBI RefSeq v110 Browser. Homo sapiens isolate NA24385 chromosome Y, alternate assembly T2T-CHM13v2.0. https://tinyurl.com/bdfudexn (2022).

  37. Hoyt, S. J. et al. From telomere to telomere: the transcriptional and epigenetic state of human repeat elements. Science 376, eabk3112 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  38. Warburton, P. E. et al. Analysis of the largest tandemly repeated DNA families in the human genome. BMC Genomics 9, 533 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  39. Halabian, R. & Makałowski, W. A map of 3′ DNA transduction variants mediated by non-LTR retroelements on 3202 human genomes. Biology 11, 1032 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Weissensteiner, M. H. et al. Accurate sequencing of DNA motifs able to form alternative (non-B) structures. Genome Res. 33, 907-922 (2023).

  41. Tyler-Smith, C., Taylor, L. & Müller, U. Structure of a hypervariable tandemly repeated DNA sequence on the short arm of the human Y chromosome. J. Mol. Biol. 203, 837–848 (1988).

    Article  CAS  PubMed  Google Scholar 

  42. Xue, Y. & Tyler-Smith, C. An exceptional gene: evolution of the TSPY gene family in humans and other great apes. Genes 2, 36–47 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Saxena, R. et al. Four DAZ genes in two clusters found in the AZFc region of the human Y chromosome. Genomics 67, 256–267 (2000).

    Article  CAS  PubMed  Google Scholar 

  44. Altemose, N. et al. Complete genomic and epigenetic maps of human centromeres. Science 376, eabl4178 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  45. Jain, M. et al. Linear assembly of a human centromere on the Y chromosome. Nat. Biotechnol. 36, 321–323 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Gershman, A. et al. Epigenetic patterns in a complete human genome. Science 376, eabj5089 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  47. Kasinathan, S. & Henikoff, S. Non-B-form DNA is enriched at centromeres. Mol. Biol. Evol. 35, 949–962 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Nailwal, M. & Chauhan, J. B. Azoospermia factor C subregion of the Y chromosome. J. Hum. Reprod. Sci. 10, 256 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. Kuroda-Kawaguchi, T. et al. The AZFc region of the Y chromosome features massive palindromes and uniform recurrent deletions in infertile men. Nat. Genet. 29, 279–286 (2001).

    Article  CAS  PubMed  Google Scholar 

  50. Repping, S. et al. A family of human Y chromosomes has dispersed throughout northern Eurasia despite a 1.8-Mb deletion in the azoospermia factor c region. Genomics 83, 1046–1052 (2004).

    Article  CAS  PubMed  Google Scholar 

  51. Porubsky, D. et al. Recurrent inversion polymorphisms in humans associate with genetic instability and genomic disorders. Cell 185, 1986–2005 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. Teitz, L. S., Pyntikova, T., Skaletsky, H. & Page, D. C. Selection has countered high mutability to preserve the ancestral copy number of Y chromosome amplicons in diverse human lineages. Am. J. Hum. Genet. 103, 261–275 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  53. Jobling, M. A. Copy number variation on the human Y chromosome. Cytogenet. Genome Res. 123, 253–262 (2008).

    Article  CAS  PubMed  Google Scholar 

  54. Navarro-Costa, P., Plancha, C. E. & Gonçalves, J. Genetic dissection of the AZF regions of the human Y chromosome: thriller or filler for male (in)fertility? Biomed Res. Int. 2010, e936569 (2010).

    Google Scholar 

  55. Evans, H. J., Gosden, J. R., Mitchell, A. R. & Buckland, R. A. Location of human satellite DNAs on the Y chromosome. Nature 251, 346–347 (1974).

    Article  ADS  CAS  Google Scholar 

  56. Schmid, M., Guttenbach, M., Nanda, I., Studer, R. & Epplen, J. T. Organization of DYZ2 repetitive DNA on the human Y chromosome. Genomics 6, 212–218 (1990).

    Article  CAS  PubMed  Google Scholar 

  57. Manz, E., Alkan, M., Bühler, E. & Schmidtke, J. Arrangement of DYZ1 and DYZ2 repeats on the human Y-chromosome: a case with presence of DYZ1 and absence of DYZ2. Mol. Cell. Probes 6, 257–259 (1992).

    Article  CAS  PubMed  Google Scholar 

  58. Altemose, N. A classical revival: human satellite DNAs enter the genomics era. Semin. Cell Dev. Biol. 128, 2–14 (2022).

    Article  CAS  PubMed  Google Scholar 

  59. Gripenberg, U. Size variation and orientation of the human Y chromosome. Chromosoma 15, 618–629 (1964).

    Article  CAS  PubMed  Google Scholar 

  60. Mathias, N., Bayés, M. & Tyler-Smith, C. Highly informative compound haplotypes for the human Y chromosome. Hum. Mol. Genet. 3, 115–123 (1994).

    Article  CAS  PubMed  Google Scholar 

  61. Altemose, N., Miga, K. H., Maggioni, M. & Willard, H. F. Genomic characterization of large heterochromatic gaps in the human genome assembly. PLoS Comput. Biol. 10, e1003628 (2014).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  62. Cooke, H. Repeated sequence specific to human males. Nature 262, 182–186 (1976).

    Article  ADS  CAS  PubMed  Google Scholar 

  63. Frommer, M., Prosser, J. & Vincent, P. C. Human satellite I sequences include a male specific 2.47 kb tandemly repeated unit containing one Alu family member per repeat. Nucleic Acids Res. 12, 2887–2900 (1984).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Babcock, M., Yatsenko, S., Stankiewicz, P., Lupski, J. R. & Morrow, B. E. AT-rich repeats associated with chromosome 22q11.2 rearrangement disorders shape human genome architecture on Yq12. Genome Res. 17, 451–460 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Webster, T. H. et al. Identifying, understanding, and correcting technical artifacts on the sex chromosomes in next-generation sequencing data. GigaScience 8, giz074 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  66. Aganezov, S. et al. A complete reference genome improves analysis of human genetic variation. Science 376, eabl3533 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Bekritsky, M. A., Colombo, C. & Eberle, M. A. Identifying genomic regions with high quality single nucleotide variant calling. Illumina https://www.illumina.com/content/illumina-marketing/amr/en_US/science/genomics-research/articles/identifying-genomic-regions-with-high-quality-single-nucleotide-.html (2023).

  68. Breitwieser, F. P., Pertea, M., Zimin, A. V. & Salzberg, S. L. Human contamination in bacterial genomes has created thousands of spurious proteins. Genome Res. 29, 954–960 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  69. Steinegger, M. & Salzberg, S. L. Terminating contamination: large-scale search identifies more than 2,000,000 contaminated entries in GenBank. Genome Biol. 21, 115 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Chrisman, B. et al. The human “contaminome”: bacterial, viral, and computational contamination in whole genome sequences from 1000 families. Sci. Rep. 12, 9863 (2022).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  71. Kent, W. J. et al. The Human Genome Browser at UCSC. Genome Res. 12, 996–1006 (2002).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  72. Rautiainen, M. et al. Telomere-to-telomere assembly of diploid chromosomes with Verkko. Nat. Biotechnol. https://doi.org/10.1038/s41587-023-01662-6 (2023).

  73. Liao, W.-W. et al. A draft human pangenome reference. Nature 617, 312–324 (2023).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  74. Jiang, Z., Hubley, R., Smit, A. & Eichler, E. E. DupMasker: a tool for annotating primate segmental duplications. Genome Res. 18, 1362–1368 (2008).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  75. Vollger, M. R., Kerpedjiev, P., Phillippy, A. M. & Eichler, E. E. StainedGlass: interactive visualization of massive tandem repeat structures with identity heatmaps. Bioinformatics 38, 2049–2051 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  76. Skene, P. J. & Henikoff, S. An efficient targeted nuclease strategy for high-resolution mapping of DNA binding sites. eLife 6, e21856 (2017).

    Article  PubMed  PubMed Central  Google Scholar 

  77. Robinson, J. T. et al. Integrative genomics viewer. Nat. Biotechnol. 29, 24–26 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  78. Shafin, K. et al. Nanopore sequencing and the Shasta toolkit enable efficient de novo assembly of eleven human genomes. Nat. Biotechnol. 38, 1044–1053 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  79. Koren, S. et al. De novo assembly of haplotype-resolved genomes with trio binning. Nat. Biotechnol. 36, 1174–1182 (2018).

    Article  CAS  Google Scholar 

  80. Kolmogorov, M., Yuan, J., Lin, Y. & Pevzner, P. A. Assembly of long, error-prone reads using repeat graphs. Nat. Biotechnol. 37, 540–546 (2019).

    Article  CAS  PubMed  Google Scholar 

  81. Poplin, R. et al. A universal SNP and small-indel variant caller using deep neural networks. Nat. Biotechnol. 36, 983–987 (2018).

    Article  CAS  PubMed  Google Scholar 

  82. Shafin, K. et al. Haplotype-aware variant calling with PEPPER-Margin-DeepVariant enables high accuracy in nanopore long-reads. Nat. Methods 18, 1322–1332 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  83. Sedlazeck, F. J. et al. Accurate detection of complex structural variations using single molecule sequencing. Nat. Methods 15, 461–468 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  84. Jiang, T. et al. Long-read-based human genomic structural variation detection with cuteSV. Genome Biol. 21, 189 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Bzikadze, A. V., Mikheenko, A. & Pevzner, P. A. Fast and accurate mapping of long reads to complete genome assemblies with VerityMap. Genome Res. 32, 2107–2118 (2022).

    Article  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  87. Porubsky, D. et al. breakpointR: an R/Bioconductor package to localize strand state changes in Strand-seq data. Bioinformatics 36, 1260–1261 (2020).

    Article  CAS  PubMed  Google Scholar 

  88. PacBio Revio WGS Dataset. Homo sapiens – GIAB trio HG002-4. https://downloads.pacbcloud.com/public/revio/2022Q4/ (2022).

  89. Poznik, D. yhaplo | Identifying Y-chromosome haplogroups. GitHub https://github.com/23andMe/yhaplo (2022).

  90. Tseng, B. et al. Y-SNP Haplogroup Hierarchy Finder: a web tool for Y-SNP haplogroup assignment. J. Hum. Genet. 67, 487–493 (2022).

    Article  CAS  PubMed  Google Scholar 

  91. Li, H. Minimap2: pairwise alignment for nucleotide sequences. Bioinformatics 34, 3094–3100 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  92. Li, H. Identifying centromeric satellites with dna-brnn. Bioinformatics 35, 4408–4410 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  93. Harris, R. S. Improved Pairwise Alignmnet of Genomic DNA (Pennsylvania State Univ., 2007).

  94. Morgulis, A., Gertz, E. M., Schäffer, A. A. & Agarwala, R. WindowMasker: window-based masker for sequenced genomes. Bioinformatics 22, 134–141 (2006).

    Article  CAS  PubMed  Google Scholar 

  95. Chin, C.-S. et al. Multiscale analysis of pangenomes enables improved representation of genomic diversity for repetitive and clinically relevant genes. Nat. Methods https://doi.org/10.1038/s41592-023-01914-y (2023).

  96. Frankish, A. et al. GENCODE 2021. Nucleic Acids Res. 49, D916–D923 (2021).

    Article  CAS  PubMed  Google Scholar 

  97. Armstrong, J. et al. Progressive Cactus is a multiple-genome aligner for the thousand-genome era. Nature 587, 246–251 (2020).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  98. Kovaka, S. et al. Transcriptome assembly from long-read RNA-seq alignments with StringTie2. Genome Biol. 20, 278 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  99. Stanke, M., Diekhans, M., Baertsch, R. & Haussler, D. Using native and syntenically mapped cDNA alignments to improve de novo gene finding. Bioinformatics 24, 637–644 (2008).

    Article  CAS  PubMed  Google Scholar 

  100. Fiddes, I. T. et al. Comparative Annotation Toolkit (CAT)—simultaneous clade and personal genome annotation. Genome Res. 28, 1029–1038 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Shumate, A. & Salzberg, S. L. Liftoff: accurate mapping of gene annotations. Bioinformatics 37, 1639–1643 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  102. Dale, R. K., Pedersen, B. S. & Quinlan, A. R. Pybedtools: a flexible Python library for manipulating genomic datasets and annotations. Bioinformatics 27, 3423–3424 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Rhie, A. et al. Towards complete and error-free genome assemblies of all vertebrate species. Nature 592, 737–746 (2021).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

  104. Pruitt, K. D. et al. RefSeq: an update on mammalian reference sequences. Nucleic Acids Res. 42, D756–D763 (2014).

    Article  CAS  PubMed  Google Scholar 

  105. Kapustin, Y., Souvorov, A., Tatusova, T. & Lipman, D. Splign: algorithms for computing spliced alignments with identification of paralogs. Biol. Direct 3, 20 (2008).

    Article  PubMed  PubMed Central  Google Scholar 

  106. Katoh, K. & Standley, D. M. MAFFT multiple sequence alignment software version 7: improvements in performance and usability. Mol Biol Evol. 30, 772-80 (2013).

  107. Slater, G. S. C. & Birney, E. Automated generation of heuristics for biological sequence comparison. BMC Bioinformatics 6, 31 (2005).

    Article  PubMed  PubMed Central  Google Scholar 

  108. Zook, J. M. et al. Integrating human sequence data sets provides a resource of benchmark SNP and indel genotype calls. Nat. Biotechnol. 32, 246–251 (2014).

    Article  CAS  PubMed  Google Scholar 

  109. Numanagić, I. et al. Fast characterization of segmental duplications in genome assemblies. Bioinformatics 34, i706–i714 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  110. Benson, G. Tandem repeats finder: a program to analyze DNA sequences. Nucleic Acids Res. 27, 573–580 (1999).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  111. Arian, F. A. S., Hubley, R. & Green, P. RepeatMasker Open-4.0 2013-2015. http://www.repeatmasker.org (2015).

  112. Storer, J., Hubley, R., Rosen, J., Wheeler, T. J. & Smit, A. F. The Dfam community resource of transposable element families, sequence models, and genome annotations. Mob. DNA 12, 2 (2021).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  113. Olson, D. & Wheeler, T. ULTRA: a model based tool to detect tandem repeats. ACM BCB 2018, 37–46 (2018)

  114. Quinlan, A. R. & Hall, I. M. BEDTools: a flexible suite of utilities for comparing genomic features. Bioinformatics 26, 841–842 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  115. Storer, J. M., Hubley, R., Rosen, J. & Smit, A. F. A. Curation guidelines for de novo generated transposable element families. Curr. Protoc. 1, e154 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  116. Kent, W. J. BLAT—the BLAST-like alignment tool. Genome Res. 12, 656–664 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  117. Szak, S. T. et al. Molecular archeology of L1 insertions in the human genome. Genome Biol. 3, research0052.1 (2002).

    Article  Google Scholar 

  118. Altschul, S. F., Gish, W., Miller, W., Myers, E. W. & Lipman, D. J. Basic local alignment search tool. J. Mol. Biol. 215, 403–410 (1990).

    Article  CAS  PubMed  Google Scholar 

  119. Cer, R. Z. et al. Searching for non-B DNA-forming motifs using nBMST (non-B DNA motif search tool). Curr. Protoc. Hum. Genet. 73, 18.7.1–18.7.22 (2012).

    Google Scholar 

  120. Zou, X. et al. Short inverted repeats contribute to localized mutability in human somatic cells. Nucleic Acids Res. 45, 11213–11221 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  121. Svetec Miklenić, M. et al. Size-dependent antirecombinogenic effect of short spacers on palindrome recombinogenicity. DNA Repair 90, 102848 (2020).

    Article  PubMed  Google Scholar 

  122. Sahakyan, A. B. et al. Machine learning model for sequence-driven DNA G-quadruplex formation. Sci. Rep. 7, 14535 (2017).

    Article  ADS  PubMed  PubMed Central  Google Scholar 

  123. Hao, Z. et al. RIdeogram: drawing SVG graphics to visualize and map genome-wide data on the idiograms. PeerJ Comput. Sci. 6, e251 (2020).

    Article  PubMed  PubMed Central  Google Scholar 

  124. Dotmatics. GraphPad Prism v.9.1.0 for Windows. https://www.graphpad.com (16 March 2021).

  125. Vollger, M. R. SafFire. GitHub https://github.com/mrvollger/SafFire (2022).

  126. Pendleton, A. L. et al. Comparison of village dog and wolf genomes highlights the role of the neural crest in dog domestication. BMC Biol. 16, 64 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  127. Hach, F. et al. mrsFAST: a cache-oblivious algorithm for short-read mapping. Nat. Methods 7, 576–577 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  128. Escalona, M. et al. Whole-genome sequence and assembly of the Javan gibbon (Hylobates moloch). J. Hered. 114, 35–43 (2023).

    Article  PubMed  Google Scholar 

  129. Cortez, D. et al. Origins and functional evolution of Y chromosomes across mammals. Nature 508, 488–493 (2014).

    Article  ADS  CAS  PubMed  Google Scholar 

  130. Stamatakis, A. RAxML version 8: a tool for phylogenetic analysis and post-analysis of large phylogenies. Bioinformatics 30, 1312–1313 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  131. Dotmatics. Geneious v2019.2.3. https://www.geneious.com/ (2019).

  132. Rambaut et al. FigTree v1.4.4. http://tree.bio.ed.ac.uk/software/figtree/ (2018).

  133. Tyler-Smith, C. & Brown, W. R. A. Structure of the major block of alphoid satellite DNA on the human Y chromosome. J. Mol. Biol. 195, 457–470 (1987).

    Article  CAS  PubMed  Google Scholar 

  134. Shepelev, V. A. et al. Annotation of suprachromosomal families reveals uncommon types of alpha satellite organization in pericentromeric regions of hg38 human genome assembly. Genomics Data 5, 139–146 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Lee, I. et al. Simultaneous profiling of chromatin accessibility and methylation on human cell lines with nanopore sequencing. Nat. Methods 17, 1191–1199 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  136. Krumsiek, J., Arnold, R. & Rattei, T. Gepard: a rapid and sensitive tool for creating dotplots on genome scale. Bioinformatics 23, 1026–1028 (2007).

    Article  CAS  PubMed  Google Scholar 

  137. Rice, P., Longden, I. & Bleasby, A. EMBOSS: The European Molecular Biology Open Software Suite. Trends Genet. 16, 276–277 (2000).

    Article  CAS  PubMed  Google Scholar 

  138. Sun, C. et al. Deletion of azoospermia factor a (AZFa) region of human Y chromosome caused by recombination between HERV15 proviruses. Hum. Mol. Genet. 9, 2291–2296 (2000).

    Article  CAS  PubMed  Google Scholar 

  139. Lassmann, T. Kalign 3: multiple sequence alignment of large datasets. Bioinformatics 36, 1928–1929 (2020).

    Article  CAS  Google Scholar 

  140. Wheeler, T. J. & Eddy, S. R. nhmmer: DNA homology search with profile HMMs. Bioinformatics 29, 2487–2489 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Stephens, Z. D. et al. Simulating next-generation sequencing datasets from empirical mutation and sequencing models. PLoS ONE 11, e0167047 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  142. Bushnell, B. BBMap: a fast, accurate, splice-aware aligner. OSTI.gov https://www.osti.gov/biblio/1241166 (2017).

  143. Aken, B. L. et al. Ensembl 2017. Nucleic Acids Res. 45, D635–D642 (2017).

    Article  CAS  PubMed  Google Scholar 

  144. Poznik, G. D. et al. Sequencing Y chromosomes resolves discrepancy in time to common ancestor of males versus females. Science 341, 562–565 (2013).

    Article  ADS  CAS  PubMed  PubMed Central  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  146. Schatz, M. C. et al. Inverting the model of genomics data sharing with the NHGRI Genomic Data Science Analysis, Visualization, and Informatics Lab-space. Cell Genomics 2, 100085 (2022).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  147. Danecek, P. et al. Twelve years of SAMtools and BCFtools. GigaScience 10, giab008 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  148. Talenti, A. & Prendergast, J. nf-LO: a scalable, containerized workflow for genome-to-genome lift over. Genome Biol. Evol. 13, evab183 (2021).

    Article  PubMed  PubMed Central  Google Scholar 

  149. Guarracino, A., Mwaniki, N., Marco-Sola, S., & Garrison, E. wfmash: whole-chromosome pairwise alignment using the hierarchical wavefront algorithm. GitHub https://github.com/ekg/wfmash (2021).

  150. Sherry, S. T., Ward, M. & Sirotkin, K. dbSNP—database for single nucleotide polymorphisms and other classes of minor genetic variation. Genome Res. 9, 677–679 (1999).

    Article  CAS  PubMed  Google Scholar 

  151. Landrum, M. J. et al. ClinVar: improving access to variant interpretations and supporting evidence. Nucleic Acids Res. 46, D1062–D1067 (2018).

    Article  CAS  PubMed  Google Scholar 

  152. Buniello, A. et al. The NHGRI-EBI GWAS Catalog of published genome-wide association studies, targeted arrays and summary statistics 2019. Nucleic Acids Res. 47, D1005–D1012 (2019).

    Article  CAS  PubMed  Google Scholar 

  153. Van der Auwera G. A. & O’Connor B. D. Genomics in the Cloud: Using Docker, GATK, and WDL in Terra (O’Reilly Media, 2020).

  154. Langmead, B. & Salzberg, S. L. Fast gapped-read alignment with Bowtie 2. Nat. Methods 9, 357–359 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  155. Ramírez, F. et al. deepTools2: a next generation web server for deep-sequencing data analysis. Nucleic Acids Res. 44, W160–W165 (2016).

    Article  PubMed  PubMed Central  Google Scholar 

  156. Zhao, H. et al. CrossMap: a versatile tool for coordinate conversion between genome assemblies. Bioinformatics 30, 1006–1007 (2014).

    Article  PubMed  Google Scholar 

  157. Marçais, G. et al. MUMmer4: a fast and versatile genome alignment system. PLoS Comput. Biol. 14, e1005944 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  158. Ondov, B. D., Bergman, N. H. & Phillippy, A. M. Interactive metagenomic visualization in a Web browser. BMC Bioinformatics 12, 385 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  159. Rhie, A. Repositories for the analysis of T2T-Y and T2T-CHM13v2.0. Zenodo https://doi.org/10.5281/zenodo.8136598 (2023).

  160. Falconer, E. et al. DNA template strand sequencing of single-cells maps genomic rearrangements at high resolution. Nat. Methods 9, 1107–1112 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

Download references

Acknowledgements

We thank P. Hallast, M. C. Loftus, M. K. Konkel, P. Ebert, T. Marschall and C. Lee for coordination and discussions, J.C.-I. Lee for sharing the GRCh38-Y coordinates used in Y-Finder and members of the Telomere-to-Telomere consortium and HPRC for constructive feedback. This work utilized the computational resources of the National Institutes of Health (NIH) HPC Biowulf cluster (https://hpc.nih.gov). Computational resources were partially provided by the e-INFRA CZ project (no. 90140), supported by the Ministry of Education, Youth and Sports of the Czech Republic and Computational Biology Core, Institute for Systems Genomics, University of Connecticut. Certain commercial equipment, instruments and materials are identified to specify adequately experimental conditions or reported results. Such identification does not imply recommendation or endorsement by the NIST, nor does it imply that the equipment, instruments or materials identified are necessarily the best available for the purpose. We thank the Intramural Research Program of NHGRI, NIH no. HG200398 (A.R., S.N., S.K., M.R., A.M.M., B.P.W. and A.M.P.); NIH no. GM123312 (S.J.H., P.G.S.G., G.A.H. and R.J.O.); NIH no. GM130691 (P.M., M.H.W. and K.D.M.); HHMI Hanna Gray Fellowship (N.A.); NIH no. CA266339 (J.G. and T.P.); NIH no. GM147352 (G.A.L.); NIH nos. HG002939 and HG010136 (R.M.H. and J.M.S.); NIH no. HG009190 (P.W.H., A. Gershman and W.T.); NIH nos. HG010263, HG006620 and CA253481 and NSF no. DBI-1627442 (M.C.S.); NIH no. GM136684 (K.D.M.); NIH nos. HG011274 and HG010548 (K.H.M.); NIH nos. HG010961 and HG010040 (H.L.); NIH no. HG007234 (M.D.); NIH no. HG011758 (F.J.S.); NIH no. DA047638 (E.G.); NIH no. GM124827 (M.A.W.); NIH no. GM133747 (R.C.M.); NIH no. CA240199 (R.J.O.); NIH nos. HG002385, HG010169 and HG010971 (E.E.E.); Stowers Institute for Medical Research (J.L.G. and T.P.); National Center for Biotechnology Information of the National Library of Medicine, NIH (F.T.-N. and T.D.M.); intramural funding at NIST (J.M.Z.); NIST no. 70NANB20H206 (M.J.); and NIH nos. HG010972 and WT222155/Z/20/Z and the European Molecular Biology Laboratory (J.A., P.F., C.G.G., L.H., T.H., S.E.H., F.J.M. and L.S.). RNA generation was supported by NIST no. 70NANB21H101 and NIH no. 1S10OD028587; the Ministry of Science and Higher Education of the Russian Federation, St. Petersburg State University, no. PURE 73023672 (I.A.A.); the Computation, Bioinformatics, and Statistics Predoctoral Training Program awarded to Penn State by the NIH (A.C.W.); and Achievement Rewards for College Scientists Foundation, The Graduate College at Arizona State University (A.M.T.O.). E.E.E. is an investigator for HHMI.

Author information

Authors and Affiliations

Authors

Contributions

V.A.S. is retired from the Institute of Molecular Genetics. Assembly was carried out by S.N., S.K. and M.R. Validation was performed by A.R., S.K., M.A., A.V.B., G.F., A.F., A.M.M., J.M., A.M., L.F.P., D.P., F.J.S., K.S., P.M., J.M.Z. and K.D.M. ChrY haplogroups were determined by A.R. and A.C.W. Alignment was done by C.-S.C., M.D., R. Harris, M.R.V. and K.D.M. Satellite annotation was performed by N.A., I.A.A., G.A.L., F.R., V.A.S. and K.H.M. N.A., J.G. and T.P. carried out FISH. Repeat annotation was done by S.J.H., P.G.S.G., G.A.H., R.M.H., J.M.S. and R.J.O. Retro-elements were dealt with by R. Halabian and W.M. Non-B DNA was dealt with by M.H.W. and K.D.M. Gene annotation was undertaken by A.R., M.D., P.F., C.G.G., L.H., M.H., J.H., T.H., F.J.M., T.D.M., S.L.S., A.S. and F.T.-N. A.R., R. Harris, W.T.H., P.M., M.T. and K.D.M. dealt with ampliconic genes. Structural annotation was performed by A.R., M.C., H.L., P.M. and K.D.M. Epigenetic analysis was performed by A.R., P.W.H., A. Gershman, W.T. and A.M.W. Mappability was performed by A.M.T.O., M.A.W. and J.M.Z. Non-B DNA was dealt with by M.H.W. and K.D.M. Variants and liftover were carried out by A.R., D.J.T., S.K., J.A., N.-C.C., M.D., E.G., A. Guarracino, N.F.H., W.T.H., S.E.H., S.H., R.C.M., N.D.O., M.E.G.S., L.S., M.R.V., S.Z., J.M.Z., E.E.E. and A.M.P. A.R., S.L.S., B.P.W. and A.M.P. dealt with contamination. Data generation was carried out by M.J., R.K.K., A.P.L., J.K.L., C.M., B.M.M., K.M.M., H.E.O., F.J.S. and Y.Z. Data management was undertaken by A.R., M.D., M.J. and J.K.L. Computational resources were sourced by R.J.O., M.C.S. and A.M.P. A.R., S.N., M.C., S.J.H., D.J.T., N.A., I.A.A., N.-C.C., E.G., J.G., P.G.S.G., A. Guarracino, R. Halabian, W.M., J.M., T.P., F.R., S.L.S., J.M.S., A.M.T.O., A.C.W., M.A.W., S.Z., J.M.Z., E.E.E., R.J.O., M.C.S., K.H.M., K.D.M. and A.M.P. wrote the manuscript draft. A.R. and A.M.P. edited the manuscript, with the assistance of all authors. J.M.Z., E.E.E., R.J.O., M.C.S., K.H.M., K.D.M. and A.M.P. supervised the research. Conceptualization was the responsibility of A.R., S.N., M.C., E.E.E., K.H.M., K.D.M. and A.M.P.

Corresponding author

Correspondence to Adam M. Phillippy.

Ethics declarations

Competing interests

S.N. is now an employee of ONT. S.K. has received travel funding for speaking at events hosted by ONT. A.F. is an employee of DNAnexus. C.-S.C. is an employee of GeneDX Holdings Corp. N.-C.C. is an employee of Exai Bio. L.F.P. receives research support from Genetech. F.J.S. receives research support from Pacific Biosciences, ONT, Illumina and Genetech. K.S. is an employee of Google LLC and owns Alphabet stock as part of the standard compensation package. W.T. has two patents (nos. 8,748,091 and 8,394,584) licensed to ONT. E.E.E. is a scientific advisory board member of Variant Bio, Inc. The remaining authors declare no competing interests.

Peer review

Peer review information

Nature thanks John Lovell, Mikkel Heide Schierup and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Peer reviewer reports are available.

Additional information

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

Extended data figures and tables

Extended Data Fig. 1 Assembling the X and Y chromosomes of HG002.

a. Chromosome X and Y components of the assembly string graph built from HiFi reads, detected based on node sequence alignments to T2T-CHM13 and GRCh38 references. Each node is colored according to the excess of paternal-specific (blue) and maternal-specific (red) k-mers, obtained from parental Illumina reads, indicating if they exclusively belong to chromosome Y or X, respectively. Most complicated tangles are localized within the heterochromatic satellite region on the Y q-arm. The X and Y subgraphs are connected in PAR1 and PAR2. Graph discontinuities are due to a lack of HiFi sequence coverage in these regions caused by contextual sequencing bias, with 9 out of 11 observed breaks falling within PAR1 on either chromosome (5 out of 5 for chromosome Y). Note that for visualization purposes the length of shorter nodes is artificially increased making the extent of the tangles appear larger than reality. b. The effects of manual pruning and semi-automated ONT read integration is illustrated from top to bottom. Top, zoomed in view of a tangle encoding the P1–P3 palindromic region in Y (approx. 22.86–27.08 Mb, see Fig. 4). Middle, corresponding subgraph following the manual pruning and recompaction. Nodes excluded from the curated “single-copy” list for automated ONT-based repeat resolution are shown in yellow. Three hairpin structures are highlighted, which form almost-perfect inverted tandem repeats encompassing the entire P3 and two P2 (red) palindromes. Node outlines in the palindromes are colored according to the palindromic arms as in Fig. 4. Bottom, corresponding subgraph following the repeat resolution using ONT read-to-graph alignments. Remaining ambiguities were resolved by evaluating ONT read alignments to all candidate reconstructions of the corresponding sub-regions. c. PAR1 subgraph labeled with HiFi read coverage on each node. Gaps (green edges) and uneven node coverage estimates indicate biases in HiFi sequencing across the region. Figure 1 shows an enrichment of SINE repeats and non-B DNA motifs in PAR1 that may underlie the sequencing gaps in this region.

Extended Data Fig. 2 Validation and polishing of the T2T-Y.

a. Evaluation and polishing workflow performed on T2T-CHM13v1.1 autosomes + HG002 XY assemblies. b. Venn diagram of the k-mers from the parents and child. On the left, hap-mers18 represent haplotype specific k-mers inherited by the child. The darker outlined circle inside the child k-mers represent single-copy k-mers (k-mers occurring once in the assembly and single-copy in the child’s genome). Right figure shows an example of the paternal specific, “single-copy” and “marker” k-mers. The marker set includes both multi-copy and single-copy k-mers specific to the paternal haplotype that were inherited by the child. Unlike polishing the nearly haploid CHM13 assembly17, both single-copy k-mers and marker k-mers were used for the marker-assisted alignments to HG002 XY. This helped align more reads within repetitive regions to the correct chromosome for evaluation during polishing. Right panel shows counts of the k-mers and coverage of HiFi and ONT reads using the marker-assisted Winnowmap2 alignment, in addition to alignments from VerityMap, which uses locally unique k-mers for anchoring the reads. c. Aggregated Strand-seq coverage profile across all 65 libraries on GRCh38-Y (top) and T2T-Y (bottom). Each bar represents read counts in every 20 kb bin supporting the reference in forward direction (light green) or reverse direction (dark green). Multiple spikes in reverse direction (black asterisks) in GRCh38-Y indicate inversion polymorphisms relative to HG002, likely due to differences between the haplogroups. Such spikes in coverage are not observed on T2T-X and T2T-Y, which confirm the structural and directional accuracy of the HG002 assemblies. A 3 kb inversion of the unique sequence between the P5 palindromic arms was identified as erroneous in T2T-Y (red asterisk), but was confirmed to be polymorphic in the population and left uncorrected in this version of the assembly.

Extended Data Fig. 3 Large structural differences between T2T-Y and previous GRCh Y assemblies.

a-b. Ampliconic genes and X-degenerate sequences revealed from alignments between GRCh38-Y (Y-axis) and T2T-Y (X-axis). a. Dotplot generated using LastZ93 after softmasking with WindowMasker94. b. Identity was computed from matches and mismatches over positions with alignments, excluding gaps. c. Structural differences revealed using PRG-TK95 against GRCh38-Y and GRCh37-Y in the euchromatic region of the Y chromosome.

Extended Data Fig. 4 Repeat discovery and annotation of T2T-Y.

a. Assembly completion allowed for a full assessment of repeats and resulted in the identification of previously unknown satellite arrays (predominantly in the PAR1) and subunit repeats that fall within one of three composite repeat units (TSPY, RBMY, DAZ). b. Ideogram of TE density (per 100 kb bin). This is an extension of Fig. 1 with non-SINEs expanded into separate TE classes (SVA, LTR, LINE, DNA/RC). Density scale ranges from low (white, zero) to high (black, relative to total density) and sequence classes are denoted by color. c. Summary (in terms of base coverage per region) across all five TE classes and two specific families: Alu/SINE and L1/LINE. The satellites in (b) were kept separate as two categories; Cen/Sat as the left satellite block including alpha satellites and DYZ19, while all other categories were combined per sequence classes.

Extended Data Fig. 5 Non-B DNA motifs along the T2T-Y.

HSat3 on the Yq and satellite sequences around the centromere are more enriched with A-phased repeats, direct repeats and STRs, while HSat1B is more enriched with inverted repeats and mirror repeats. Enrichment of non-B DNA sequences were also observed in the PAR region. Notably, the TSPY gene array is enriched for G4 and Z-DNA motifs, as shown in Extended Data Fig. 6b.

Extended Data Fig. 6 Phylogenetic tree analysis of the ampliconic TSPY gene family and pattern of non-B DNA structure.

a. Phylogenetic tree analysis using protein-coding TSPYs from a Sumatran Orangutan (Pongo abelii) and a Silvery gibbon (Hylobates moloch) as outgroups confirmed TSPY2 (distal to the array) and TSPY copies within the array originated from the same branch, distinguished from the rest of the TSPY pseudogenes. Rectangular inset shows a cartoon representation of the simplified tree. Numbers next to the triangles indicate the number of TSPY genes in the same branch. b. G4 and Z-DNA structures predicted for a typical TSPY copy inside the TSPY array. All TSPY copies in the array have the same signature, with one G4 peak present ~500 bases upstream of the TSPY (arrow). Higher Quadron score122 (Q-score) indicates a more stable G4 structure, with scores over 19 considered stable (dotted line).

Extended Data Fig. 7 Recurrent inversions identified with Strand-seq.

a. Five out of 15 individuals have the inverted variant as present in HG002 at the P3 palindrome (white arrow). Although inversions across P1–P2 (yellow and red arrows) are difficult to confirm with Strand-seq because of the high sequence similarity between the palindromic arms, different orientations are observable in these samples. b. Strand states for 65 Strand-seq libraries of HG002. Depending on the mappings of directional Strand-seq reads (+ reads: ‘Crick’, C, – reads: ‘Watson’, W), reference sequence was assigned in three states: WC, WW, and CC. WC, roughly equal mixture of plus and minus reads; WW, all reads mapped in minus orientation; CC, all reads mapped in plus orientation. Changes in strand state along a single chromosome are normally caused by a double-strand-break (DSBs) that occurred during DNA replication160 in a random fashion and we refer to them as sister-chromatid-exchanges (SCEs, yellow thunderbolts). Recurrent change in strand state over the same region in multiple Strand-seq cells indicates misassembly. Similarly, collapsed or incomplete assembly of a certain genomic region will result in a recurrent strand state change as observed for GRCh38-Y (black arrowheads). In contrast, T2T-Y shows strand state changes randomly distributed along each Strand-seq library with no evidence of misassembly or collapse. c. Strand-seq profile of selected libraries over T2T-Y summarized in bins (bin size: 500 kb, step size: 50 kb). Teal, Crick read counts; orange, Watson read counts. As ChrY is haploid, reads are expected to map only in Watson or Crick orientation. Light gray rectangles highlight regions where SCEs were detected in the heterochromatic Yq12 despite a lower coverage of Strand-seq reads. A modified breakpointR parameter was used (windowsize = 500000 minReads = 20) in order to refine detected SCEs presented in panel b and c.

Extended Data Fig. 8 Satellite annotation and recent expansion events in the Yq heterochromatin.

a. A plot showing the top repeat periodicities detected by NTRprism44 in 50 kb blocks tiled across T2T-Y, with centromeric satellite annotations overlaid on the X axis. Large arrays are labeled with their historic nomenclature1, HSat subfamilies61, and predominant repeat periodicities. b. An exact 2000-mer match dotplot of the Yq region (a dot is plotted when an identical 2000 base sequence is found at positions X and Y). The lower triangle has DYZ1/DYZ2 annotations overlaid as yellow and blue bars, respectively. Circled patterns in the upper triangle correspond to recent iterative duplication events, which are illustrated below the X axis. c. A reconstruction of a possible sequence of recent iterative duplications that could explain the observed dotplot patterns. d. A 2000-mer dotplot comparison of two ~800 kb HSat1B sub-arrays that were part of a recent large duplication event, along with self-self comparisons of the same arrays, revealing sites of more recent and smaller-scale deletions and expansions (annotated in yellow and red, with a possible sequence of events illustrated by the schematic on the right).

Extended Data Fig. 9 Genomic similarity in PARs and XTR and improved MAPQ of the PARs through informed sex chromosome complement reference.

a. Dotplots from LASTZ alignments of the CHM13-X, HG002-X, and HG002-Y (T2T-Y) over 96% sequence identity. Dashed gray lines represent the start and end of the approximate PARs or XTR boundaries. Disconnected diagonal lines indicate the presence of genomic diversity between each paired region. More genomic differences are observed in the PAR1 between the HG002-Y and CHM13-X. b-c. Average mapping quality (MAPQ) across GRCh38-X from simulated reads of an XX (b) and XY (c) sample. Top, a default version of GRCh38 (with two copies of identical PARs on XY). Middle, a version of GRCh38 informed on the sex chromosome complement (SCC) of the sample (entire Y hard-masked for the XX sample vs. only PARs on the Y hard-masked for the XY sample). Bottom, the difference in average MAPQ between the SCC and default approaches. MAPQ was averaged in 50 kb windows, sliding 10 kb across the chromosome. A positive value means MAPQ score is higher with SCC reference alignment compared to default alignment.

Extended Data Fig. 10 Number of variants called from 1KGP and SGDP individuals.

a. More variants are called on the X-PARs when using the sex chromosome complement reference approach (calling variants in diploid mode on PARs) than the non-masked approach (calling variants in haploid mode on PARs). The 1KGP results for GRCh38-Y are from Aganezov et al.66, which was performed on CHM13v1.0+GRCh38-Y. b. Num. of variants called from each 1KGP XY sample on chromosome GRCh38-Y and T2T-Y c. Num. of variants called in the syntenic region between the two Ys. A large num. of additional variants are called on each sample attributed to the newly added, non-syntenic sequences on T2T-Y. Within the syntenic regions, a reduction in the number of variants is observed for each population except for samples from R1 haplogroups as shown in Fig. 6c. d. Aggregated total number of variants for the 279 SGDP samples per chromosome. e. SGDP genome-wide counts of variants per-sample (n = 279) demonstrate increased variation in African samples regardless of reference. Each bar in the box plot represents the 1st, 2nd (median), and 3rd quartile of the number of variants in each population. Whiskers are bound to the 1.5 × interquartile range. Data outside of the whisker ranges are shown as dots. For the SGDP samples, variants were called using T2T-CHM13+Y or GRCh38 as the reference. All variants shown in this figure were filtered for “high quality (PASS)”.

Extended Data Fig. 11 Human contaminants in bacterial reference genomes.

a. Number of distinct RefSeq accessions in every 10 kb window containing 64-mers of GRCh38-Y (top), T2T-Y (middle), and in T2T-Y only (bottom). Here, RefSeq sequences with more than 20 64-mers or matching over 10% of the Y chromosome are included. b. Length distribution of the sequences from (a) in log scale. Majority of the shorter (<1 kb) sequences contain 64-mers found in HSat1B or HSat3. c. Number of bacterial RefSeq entries by strain identified to contain sequences of T2T-Y and not GRCh38-Y, visualized with Krona158.

Supplementary information

Supplementary Information

This file contains Supplementary Methods, Figs. 1–19 and Notes 1–4.

Reporting Summary

Peer Review File

Supplementary Tables

This file contains Supplementary Tables 1–32.

Supplementary Data

This file contains Supplementary Data 1–3.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Rhie, A., Nurk, S., Cechova, M. et al. The complete sequence of a human Y chromosome. Nature 621, 344–354 (2023). https://doi.org/10.1038/s41586-023-06457-y

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41586-023-06457-y

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research