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:

3D mapping and accelerated super-resolution imaging of the human genome using in situ sequencing

Abstract

There is a need for methods that can image chromosomes with genome-wide coverage, as well as greater genomic and optical resolution. We introduce OligoFISSEQ, a suite of three methods that leverage fluorescence in situ sequencing (FISSEQ) of barcoded Oligopaint probes to enable the rapid visualization of many targeted genomic regions. Applying OligoFISSEQ to human diploid fibroblast cells, we show how four rounds of sequencing are sufficient to produce 3D maps of 36 genomic targets across six chromosomes in hundreds to thousands of cells, implying a potential to image thousands of targets in only five to eight rounds of sequencing. We also use OligoFISSEQ to trace chromosomes at finer resolution, following the path of the X chromosome through 46 regions, with separate studies showing compatibility of OligoFISSEQ with immunocytochemistry. Finally, we combined OligoFISSEQ with OligoSTORM, laying the foundation for accelerated single-molecule super-resolution imaging of large swaths of, if not entire, human genomes.

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: Using OligoFISSEQ to sequence barcoded Oligopaints in situ.
Fig. 2: OligoFISSEQ-LIT on 36plex-5K.
Fig. 3: Every-pixel analysis pipeline on 36plex-5K.
Fig. 4: Improving O-LIT by using JEB.
Fig. 5: Tracing 46 regions along the X chromosome.
Fig. 6: OligoFISSEQ extensions and applications.

Similar content being viewed by others

Data availability

All data are available in the main text or the supplementary materials, and materials are available upon request. Information regarding all datasets (for example, cells, replicates and filters) can be found in Supplementary Table 9. Source data are provided with this paper.

Code availability

All code is available at https://github.com/3DGenomes/OligoFISSEQ/.

References

  1. Hu, Q., Maurais, E. G. & Ly, P. Cellular and genomic approaches for exploring structural chromosomal rearrangements. Chromosome Res. 28, 19–30 (2020).

    CAS  PubMed  PubMed Central  Google Scholar 

  2. Bolzer, A. et al. Three-dimensional maps of all chromosomes in human male fibroblast nuclei and prometaphase rosettes. PLoS Biol. 3, e157 (2005).

    PubMed  PubMed Central  Google Scholar 

  3. Cremer, T. & Cremer, M. Chromosome territories. Cold Spring Harb. Perspect. Biol. 2, a003889 (2010).

  4. Beliveau, B. J. et al. Versatile design and synthesis platform for visualizing genomes with Oligopaint FISH probes. Proc. Natl Acad. Sci. USA 109, 21301–21306 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  5. Beliveau, B. J. et al. Single-molecule super-resolution imaging of chromosomes and in situ haplotype visualization using Oligopaint FISH probes. Nat. Commun. 6, 7147 (2015).

    CAS  PubMed  Google Scholar 

  6. Chen, K. H., Boettiger, A. N., Moffitt, J. R., Wang, S. & Zhuang, X. Spatially resolved, highly multiplexed RNA profiling in single cells. Science 348, aaa6090 (2015).

  7. Boettiger, A. N. et al. Super-resolution imaging reveals distinct chromatin folding for different epigenetic states. Nature 529, 418–422 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Shah, S., Lubeck, E., Zhou, W. & Cai, L. In situ transcription profiling of single cells reveals spatial organization of cells in the mouse hippocampus. Neuron 92, 342–357 (2016).

  9. Wang, S. et al. Spatial organization of chromatin domains and compartments in single chromosomes. Science 353, 598–602 (2016).

  10. Beliveau, B. J. et al. In situ super-resolution imaging of genomic DNA with OligoSTORM and OligoDNA-PAINT. Methods Mol. Biol. 1663, 231–252 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  11. Cattoni, D. I. et al. Single-cell absolute contact probability detection reveals chromosomes are organized by multiple low-frequency yet specific interactions. Nat. Commun. 8, 1753 (2017).

    PubMed  PubMed Central  Google Scholar 

  12. Eng, C.-H. L., Shah, S., Thomassie, J. & Cai, L. Profiling the transcriptome with RNA SPOTs. Nat. Methods 14, 1153–1155 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  13. Bintu, B. et al. Super-resolution chromatin tracing reveals domains and cooperative interactions in single cells. Science 362, eaau1783 (2018).

    PubMed  PubMed Central  Google Scholar 

  14. Nir, G. et al. Walking along chromosomes with super-resolution imaging, contact maps and integrative modeling. PLoS Genet. 14, e1007872 (2018).

    PubMed  PubMed Central  Google Scholar 

  15. Rosin, L. F., Nguyen, S. C. & Joyce, E. F. Condensin II drives large-scale folding and spatial partitioning of interphase chromosomes in Drosophila nuclei. PLoS Genet. 14, e1007393 (2018).

    PubMed  PubMed Central  Google Scholar 

  16. Szabo, Q. et al. TADs are 3D structural units of higher-order chromosome organization in Drosophila. Sci. Adv. 4, eaar8082 (2018).

  17. Cardozo Gizzi, A. M. et al. Microscopy-based chromosome conformation capture enables simultaneous visualization of genome organization and transcription in intact organisms. Mol. Cell 74, 212–222 (2019).

    CAS  PubMed  Google Scholar 

  18. Eng, C.-H. L. et al. Transcriptome-scale super-resolved imaging in tissues by RNA seqFISH. Nature 568, 235–239 (2019).

  19. Fields, B. D., Nguyen, S. C., Nir, G. & Kennedy, S. A multiplexed DNA FISH strategy for assessing genome architecture in Caenorhabditis elegans. eLife 8, e42823 (2019).

    PubMed  PubMed Central  Google Scholar 

  20. Mateo, L. J. et al. Visualizing DNA folding and RNA in embryos at single-cell resolution. Nature 568, 49–54 (2019).

  21. Sawh, A. N. et al. Lamina-dependent stretching and unconventional chromosome compartments in early C. elegans embryos. Mol. Cell 78, 96–111 (2020).

  22. Ke, R. et al. In situ sequencing for RNA analysis in preserved tissue and cells. Nat. Methods 10, 857–860 (2013).

    CAS  PubMed  Google Scholar 

  23. Lee, J. H. et al. Highly multiplexed subcellular RNA sequencing in situ. Science 343, 1360–1363 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Metzker, M. L. Sequencing technologies—the next generation. Nat. Rev. Genet. 11, 31–46 (2010).

    CAS  PubMed  Google Scholar 

  25. Shendure, J. et al. DNA sequencing at 40: past, present and future. Nature 550, 345–353 (2017).

    CAS  PubMed  Google Scholar 

  26. Player, A. N., Shen, L.-P., Kenny, D., Antao, V. P. & Kolberg, J. A. Single-copy gene detection using branched DNA (bDNA) in situ hybridization. J. Histochem. Cytochem. 49, 603–612 (2001).

    CAS  PubMed  Google Scholar 

  27. Heride, C. et al. Distance between homologous chromosomes results from chromosome positioning constraints. J. Cell Sci. 123, 4063–4075 (2010).

    CAS  PubMed  Google Scholar 

  28. Mayer, W., Smith, A., Fundele, R. & Haaf, T. Spatial separation of parental genomes in preimplantation mouse embryos. J. Cell Biol. 148, 629–634 (2000).

    CAS  PubMed  PubMed Central  Google Scholar 

  29. Hua, L. L. & Mikawa, T. Mitotic antipairing of homologous and sex chromosomes via spatial restriction of two haploid sets. Proc. Natl Acad. Sci. USA 115, E12235–E12244 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  30. Reichmann, J. et al. Dual-spindle formation in zygotes keeps parental genomes apart in early mammalian embryos. Science 361, 189–193 (2018).

    CAS  PubMed  Google Scholar 

  31. Joyce, E. F., Erceg, J. & Wu, C.-t. Pairing and anti-pairing: a balancing act in the diploid genome. Curr. Opin. Genet. Dev. 37, 119–128 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Watkins, N. E. & SantaLucia, J. Jr. Nearest-neighbor thermodynamics of deoxyinosine pairs in DNA duplexes. Nucleic Acids Res. 33, 6258–6267 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Yunis, J. J. & Prakash, O. The origin of man: a chromosomal pictorial legacy. Science 215, 1525–1530 (1982).

    CAS  PubMed  Google Scholar 

  34. Tjong, H. et al. Population-based 3D genome structure analysis reveals driving forces in spatial genome organization. Proc. Natl Acad. Sci. USA 113, E1663–E1672 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Kishi, J. Y. et al. SABER amplifies FISH: enhanced multiplexed imaging of RNA and DNA in cells and tissues. Nat. Methods 16, 533–544 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  36. Rouhanifard, S. H. et al. ClampFISH detects individual nucleic acid molecules using click chemistry–based amplification. Nat. Biotechnol. 37, 84–89 (2019).

    CAS  Google Scholar 

  37. Nora, E. P. et al. Spatial partitioning of the regulatory landscape of the X-inactivation centre. Nature 485, 381–385 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  38. Rao, S. S. P. et al. A 3D map of the human genome at kilobase resolution reveals principles of chromatin looping. Cell 159, 1665–1680 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  39. Deng, X. et al. Bipartite structure of the inactive mouse X chromosome. Genome Biol. 16, 152 (2015).

    PubMed  PubMed Central  Google Scholar 

  40. Minajigi, A. et al. A comprehensive Xist interactome reveals cohesin repulsion and an RNA-directed chromosome conformation. Science 349, aab2276 (2015).

  41. Darrow, E. M. et al. Deletion of DXZ4 on the human inactive X chromosome alters higher-order genome architecture. Proc. Natl Acad. Sci. USA 113, E4504–E4512 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Giorgetti, L. et al. Structural organization of the inactive X chromosome in the mouse. Nature 535, 575–579 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  43. Wang, C.-Y., Jégu, T., Chu, H.-P., Oh, H. J. & Lee, J. T. SMCHD1 merges chromosome compartments and assists formation of super-structures on the inactive X. Cell 174, 406–421 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  44. Rust, M. J., Bates, M. & Zhuang, X. W. Sub-diffraction-limit imaging by stochastic optical reconstruction microscopy (STORM). Nat. Methods 3, 793–795 (2006).

    CAS  PubMed  PubMed Central  Google Scholar 

  45. Chen, F., Tillberg, P. W. & Boyden, E. S. Expansion microscopy. Science 347, 543–548 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  46. Sage, D. et al. Super-resolution fight club: assessment of 2D and 3D single-molecule localization microscopy software. Nat. Methods 16, 387–395 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  47. Kohman, R. E. & Church, G. M. Fluorescent in situ sequencing of DNA barcoded antibodies. Preprint at bioRxiv https://doi.org/10.1101/2020.04.27.060624 (2020).

  48. Finn, E. H. et al. Extensive heterogeneity and intrinsic variation in spatial genome organization. Cell 176, 1502–1515 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Finn, E. H. & Misteli, T. Molecular basis and biological function of variability in spatial genome organization. Science 365, eaaw9498 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  50. Lorenz, E. N. Deterministic nonperiodic flow. J. Atmos. Sci. 20, 130–141 (1963).

    Google Scholar 

  51. Beliveau, B. J. et al. OligoMiner provides a rapid, flexible environment for the design of genome-scale oligonucleotide in situ hybridization probes. Proc. Natl Acad. Sci. USA 115, E2183–E2192 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  52. Ball, M. P. et al. A public resource facilitating clinical use of genomes. Proc. Natl Acad. Sci. USA 109, 11920–11927 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  53. Zhang, K. et al. Digital RNA allelotyping reveals tissue-specific and allele-specific gene expression in human. Nat. Methods 6, 613–618 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Pardue, M. L. et al. Molecular hybridization of radioactive DNA to the DNA of cytological preparations. Proc. Natl Acad. Sci. USA 64, 600–604 (1969).

    CAS  PubMed  PubMed Central  Google Scholar 

  55. Bauman, J. G., Wiegant, J., Borst, P. & van Duijn, P. A new method for fluorescence microscopical localization of specific DNA sequences by in situ hybridization of fluorochrome-labelled RNA. Exp. Cell Res. 128, 485–490 (1980).

    CAS  PubMed  Google Scholar 

  56. Shendure, J. et al. Accurate multiplex polony sequencing of an evolved bacterial genome. Science 309, 1728–1732 (2005).

    CAS  PubMed  Google Scholar 

  57. Valouev, A. et al. A high-resolution, nucleosome position map of C. elegans reveals a lack of universal sequence-dictated positioning. Genome Res. 18, 1051–1063 (2008).

  58. Guo, J. et al. Four-color DNA sequencing with 3′-O-modified nucleotide reversible terminators and chemically cleavable fluorescent dideoxynucleotides. Proc. Natl Acad. Sci. USA 105, 9145–9150 (2008).

  59. Lubeck, E., Coskun, A. F., Zhiyentayev, T., Ahmad, M. & Cai, L. Single-cell in situ RNA profiling by sequential hybridization. Nat. Methods 11, 360–361 (2014).

    CAS  PubMed  PubMed Central  Google Scholar 

  60. Moffitt, J. R. et al. High-performance multiplexed fluorescence in situ hybridization in culture and tissue with matrix imprinting and clearing. Proc. Natl Acad. Sci. USA 113, 14456–14461 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

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

    CAS  PubMed  Google Scholar 

  62. Linkert, M. et al. Metadata matters: access to image data in the real world. J. Cell Biol. 189, 777–782 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  63. Schmid, B., Schindelin, J., Cardona, A., Longair, M. & Heisenberg, M. A high-level 3D visualization API for Java and ImageJ. BMC Bioinformatics 11, 274 (2010).

    PubMed  PubMed Central  Google Scholar 

  64. Anand, L. ChromoMap: an R package for interactive visualization and annotation of chromosomes. Preprint at bioRxiv https://doi.org/10.1101/605600 (2019).

  65. 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 

  66. Richardson, W. H. Bayesian-based iterative method of image restoration. J. Opt. Soc. Am. 62, 55–59 (1972).

    Google Scholar 

  67. Parslow, A., Cardona, A. & Bryson-Richardson, R. J. Sample drift correction following 4D confocal time-lapse imaging. J. Vis. Exp. 51086 (2014).

  68. Tinevez, J.-Y. et al. TrackMate: an open and extensible platform for single-particle tracking. Methods 115, 80–90 (2017).

    CAS  PubMed  Google Scholar 

  69. Wagstaff, K., Cardie, C., Rogers, S. & Schroedl, S. Constrained k-means clustering with background knowledge. In Proc. 18th International Conference on Machine Learning, 577–584 (2001).

  70. Russel, D. et al. Putting the pieces together: integrative modeling platform software for structure determination of macromolecular assemblies. PLoS Biol 10, e1001244 (2012).

Download references

Acknowledgements

We acknowledge members of the Marti-Renom and Wu laboratories for technical and conceptual support, especially T. Ryu, A. Lioutas and S. Aufmkolk as well as J. AlHaj Abed, S. D. Lee, J. Erceg and T. Hatkevich; B. Beliveau, H. Sasaki, J. Horrell, L. Cai, J. Kishi and P. Soler-Vila for discussion; D. Barclay, R. Kohman, E. Iyer, K. Rodgers, A. Skrynnyk, J. Tam and R. Terry for discussion about FISSEQ and sequencing reagents; S. Alon, F. Chen, Z. Chiang, D. Goodwin, A. Payne, A. Sinha and O. Wassie for discussion about FISSEQ; C. Ebeling, J. Rosenberg and J. Stuckey for discussion and technical assistance; F. Pan and A. Hutchinson for assistance in procuring SOLiD reagents; P. Montero-Llopis and the MicRoN imaging core at Harvard Medical School; the ImageJ discussion forum; and StackOverflow. This work was supported by a Damon Runyon Dale F. Frey Breakthrough Award (to B.J.B.) to support B.J.B. and E.A.H., awards from the NSERC of Canada (PGS D) to P.L.R., the NIH (HG005550 and HG008525) and NSF (DGE1144152) to E.R.D., the European Research Council under the Seventh Framework Program (FP7/2007–2013 609989), the European Union’s Horizon 2020 Research and Innovation Program (676556) and the Spanish Ministerio de Ciencia, Innovación y Universidades (BFU2017-85926-P) to M.A.M.-R., the Centro de Excelencia Severo Ochoa 2013–2017 (SEV-2012-0208) and the CERCA Programme/Generalitat de Catalunya to the CRG, from the NIH to GMC (RM1HG008525-03) and the NIH (DP1GM106412, R01HD091797 and R01GM123289) to C.-t.W.

Author information

Authors and Affiliations

Authors

Contributions

H.Q.N., S.C., D.C., S.C.N., G.M.C., E.R.D., M.A.M.R. and C.-t.W. conceived the study with the original conceptualization of OligoFISSEQ contributed by S.C.N and E.R.D.; G.N., A.L. and N.M.C.M. provided guidance for barcode design and angle analysis; A.L., E.A.H. and B.J.B. provided guidance for Oligopaint sequences and barcode design. P.L.R. supported early protocol development; M.H. provided technical support; H.Q.N. and S.C. designed and performed the experiments. H.Q.N., S.C., D.C., G.M.C., M.A.M.R. and C.-t.W. analyzed the data; H.Q.N. wrote the manuscript with S.C., D.C., M.A.M.R. and C.-t.W. with input from all authors; C.-t.W. oversaw the project.

Corresponding authors

Correspondence to Marc A. Marti-Renom or C.-ting Wu.

Ethics declarations

Competing interests

Harvard University has filed patent applications on behalf of C.-t.W., H.Q.N. and S.C., pertaining to Oligopaints and related oligonucleotide-based methods for genome imaging. E.R.D. is currently an employee of ReadCoor and has an equity interest in ReadCoor. Potential conflicts of interest for G.M.C. are listed on http://arep.med.harvard.edu/gmc/tech.html/. C.-t.W. has an equity interest in ReadCoor and an active research collaboration with Bruker Nano in her laboratory at Harvard Medical School.

Additional information

Peer review information Lei Tang was the primary editor on this article and managed its editorial process and peer review in collaboration with the rest of the editorial team.

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

Extended data

Extended Data Fig. 1 Chr19-20K and 36plex-5K-O-LIT optimization.

a, Chr19-20K targets 18,536 Oligopaint oligos to human chromosome 19. Right, Chr19-20K detection with secondary oligo (red) in PGP1f cells representative of 5 replicates. b, Signal is completely removed in each OligoFISSEQ method after cleavage. Images showing two rounds of sequencing with a cleavage step (C) and representative of 4 replicates. c, 36plex-5K O-LIT off of both Mainstreet and Backstreet (MSBS; bottom, red) produces stronger signal than off of Mainstreet (MS; top, blue). Cy5 channel from first round of O-LIT. n = 1. d, O-LIT off of both streets produces stronger signal than off of MS. Grey intensity value measurements from yellow lines in panel c. n = 1. e, Raw, non-deconvolved field of view of cell from Figs. 2c, d and 3a–c. Maximum z-projection. n = 1. f, Manual decoding of cell from panel c and Figs. 2c, d and 3a–c yields 100% target recovery. n = 1. g, Tier1 detection efficiency after 36plex-5K O-LIT off of both streets and detected with TrackMate (blue, 29.93 ± 4.9%) or Every-pixel (orange, 62.8% ± 4.8%). n = 111 cells from 3 replicates. Detection efficiency from individual replicates are plotted. Error bars represent 95% bootstrap confidence interval of the mean.

Source data

Extended Data Fig. 2 Detection efficiency after 36plex-5K O-LIT.

a, Detection efficiency without filtering after 36plex-5K O-LIT off of both streets. 95 ± 5.15% of targeted regions are detected (n = 611 from 15 replicates). Detection efficiency from individual replicates are plotted. Error bars represent 95% bootstrap confidence interval of the mean. b, False positive (FP) discovery rate from panel a. FP discovery rate from individual replicates are plotted. Error bars represent 95% bootstrap confidence interval of the mean. c, Tier 1 detection efficiency after 36plex-5K O-LIT off of Mainstreet (orange, 61.93 ± 12%, n = 53 from 2 replicates) versus off of both streets (blue, 62.17% ± 6.68%, n = 611 cells from 15 replicates). Detection efficiency from individual replicates are plotted. Error bars represent 95% bootstrap confidence interval of the mean. d, FP discovery rate from panel c. Using Mainstreet = 8.64% and using both streets = 5.29%. FP discovery rate from individual replicates are plotted. Error bars represent 95% bootstrap confidence interval of the mean. e, Tier 2 detection efficiency after 36plex-5K off of Mainstreet (orange, 92.3% ± 3.42% from 53 cells from 2 replicates) versus off of both streets (blue, 80.19 ± 7.29%, n = 611 cells from 15 replicates). Detection efficiency from individual replicates are plotted. Error bars represent 95% bootstrap confidence interval of the mean. f, Detection efficiency after 36plex-5K O-LIT off of both streets for individual cells from 15 replicates in panel e. g, Percentage of cells displaying a range of efficiencies of barcode detection after 36plex-5K O-LIT off of both streets. Data taken from panel e. h, Principal component analysis showing lack of batch effect in 36plex datasets (n = 1171 cells from 15 36plex-5K O-LIT replicates and 8 36plex-1K O-eLIT replicates).

Source data

Extended Data Fig. 3 O-LIT with 36plex-5K to interrogate genome organization.

a, Chromosome traces of Cell 611 after Tier 2 detection of cell 611 after four rounds of O-LIT 36plex-5K off of both streets. 59/66 (89%) of 36plex-5K targets were detected. Image is from the first round of O-LIT with target identities. n = 1. b, Ball and stick of Cell 611. Colored spheres represent chromosomal targets, while black spheres represent targets that were not detected and, thus, were placed by calculating the median proportionate distance between flanking detected targets. Beginning of chromosome (for example 2pR1) marked by an asterisk. c, Single-cell pairwise spatial distance matrix after Tier 1 (top) and Tier 2 (bottom) detection of the nucleus in Fig. 3. Targets are represented on the x-axis with homologs separately displayed. Undetected targets are represented by grey lines. d, Single-cell pairwise spatial distance matrix after Tier 1 (top) and Tier 2 (bottom) detection of Cell 611. Targets are represented on the x-axis with homologs separately displayed. Undetected targets are represented by grey lines. e, 36plex-5K population pairwise spatial distances (top, from Fig. 3f). Average pairwise spatial distances from cell population after Tier 1 detection (n = 611 from 15 replicates). (Spearman’s rank correlation 0.705, two-sided p-value for a hypothesis test whose null hypothesis is that two sets of data are uncorrelated = 1.77e-174). Measurements from homologous targets were combined. Bottom, Hi-C data of 36plex-5K targets obtained from (Nir et al. 2018). f, Average distances between the nuclear membrane and the closest of the six targets imaged for each chromosome. (n = 686, 668, 364, 586, 760, and 494 for Chr2, 3, 5, 16, 19, and X, respectively.) The thick line in each violin plot represents the Interquartile range (IQR), the white dot marks the median and the thin lines extend 1.5 times the IQR.

Source data

Extended Data Fig. 4 O-LIT with 36plex-5K to interrogate homolog organization.

a, Minimum distances between heterologous and homologous chromosomes. All measurements represent distances between the geometric centers of chromosomes for which all six targets were imaged. Distances between a chromosome and a heterologous chromosome is the shorter of the two distances between that chromosome and the two homologous copies of the heterologous chromosome (n = 686, 668, 364, 586, 760, and 494 for Chr2, 3, 5, 16, 19, and X, respectively). Inter-homolog distances for Chr16 and 19 are less than those for Chr2, 3, and 5 (independent-samples t-test p = 4.28 ×10-37). Boxes represent the IQR (25th, 50th and 75th percentiles) and whiskers extend 1.5 times the IQR. b, Number of cells with varying numbers of homologs split by K-means clustering. The K-means algorithm was applied to 258 nuclei, individually, to cluster chromosomes into two groups based on proximity and then report the number of homolog pairs that were split by the clustering. A value of “5” indicates that the homologs from each five pairs of imaged autosomes in a single nucleus clustered into two spatially separate groups. Observed, PGP1f cells. Directed random, raw positions in Observed but with the chromosome identities of all positions randomized, with the larger chromosomes (2, 3, 5) biased towards the nuclear periphery and smaller chromosomes (16 and 19) biased towards the nuclear interior. Completely random category, randomization of the chromosome identities carried out with no spatial bias. The significance of each pair was evaluated from a two proportion z-test with n = 258 for each category with a null hypothesis of equal proportion and a significance level of 0.05. c, Density plots of homolog positions. Built by using Kernel density estimation (KDE) of nuclei projected and aligned along the x-y plane of the position of the chromosomes. d, Pie charts of total number of cells with homologs split by a virtual line along the y-axis. e, Number of aligned cells with homologs split by a virtual line parallel to the y-axis at different distances from the origin, that is, number of autosomes with one of their homologs on the left of the line and the other on the right (n = 258 for each category). Boxes represent the IQR (25th, 50th and 75th percentiles) and whiskers extend 1.5 times the IQR.

Source data

Extended Data Fig. 5 O-eLIT with JEB.

a, Chr19-9K. One round of O-LIT (SOLiD) or O-eLIT (JEB) off of Mainstreet. Maximum z-projections representative of 2 replicates. b, Chr19-9K signal over nuclear background measurements after one round of O-LIT (orange; n = 113 puncta from 55 cells from 2 replicates) or O-eLIT (blue; n = 136 puncta from 57 cells from 2 replicates). Bar is the mean and SD. c, Tier 1 detection of 36plex-1K after five rounds of O-LIT with SOLiD reagents (orange; average of 51.75%, n = 41) or O-eLIT with JEB (blue; average of 61.2 ± 10.2%, n = 440 from 9 replicates). Detection efficiency from individual replicates are plotted. Error bars represent 95% bootstrap confidence interval of the mean. 36plex-1K library shares first 1,000 Oligopaint oligos of each target in 36plex-5K. For example, for target 2pR1, 36plex-5K spans the chromosomal region from nt position 1,002,895 to 1,660,898 (~658 kb), whereas 36plex-1K spans the region from nt 1,002,895 to 1,147,495 (~144 kb). d, FP discovery rate from panel c. SOLiD = 7.49% and JEB = 8.95%. FP discovery rate from individual replicates are plotted. Error bars represent 95% bootstrap confidence interval of the mean. e, Chromosome traces and ball and stick of Fig. 4c cell after Tier 2 detection and five rounds of O-eLIT 36plex-1K. 63/66 (95%) targets were detected. Asterisks, beginning of chromosomes. n = 1. f, Single-cell pairwise spatial distance matrices of panel C cell. g, 36plex-1K population pairwise spatial distance measurements (top, from Fig. 3f). Average pairwise spatial distance from cell population after Tier 1 detection (n = 440 from 9 replicates). Measurements from homologous targets were combined. Bottom, Hi-C data of 36plex-5K targets obtained from (Nir et al. 2018). h, 36plex-1K detection rate for individual cells from 9 replicates. i, Percentage of cells displaying a range of efficiencies of barcode detection after 36plex-1K O-eLIT off of Mainstreet.

Source data

Extended Data Fig. 6 O-eLIT with ChrX-46plex-2K.

a, ChrX-46plex-2K O-eLIT Tier 1 detection off of one street and off of both streets combined (52.86 ± 5.78% from 177 cells from 7 replicates). Detection efficiency from individual replicates are plotted. Error bars represent 95% bootstrap confidence interval of the mean. b, FP discovery rate from panel a. Error bars represent 95% bootstrap confidence interval of the mean. c, Single-cell pairwise spatial distance matrix after Tier 1 (top) and Tier 2 (bottom) detection of Cell 1 from Fig. 5b. Undetected targets are represented by grey lines. d, Chromosome traces (top) and ball and stick representation (bottom) of Cell 177 after Tier 2 detection and interpolation and five rounds of O-eLIT on ChrX-46plex-2K off of both streets. Image is from the first round of O-eLIT with target identities. n = 1. e, Single-cell pairwise spatial distance matrix after Tier 1 (top), Tier 2 (bottom) of Cell 177 (left), and Tier 2 (top) and interpolation (bottom) of same cell (right). Undetected targets are represented by grey lines. f, ChrX-46plex-2K population pairwise spatial distances (top). Average pairwise spatial distances from cell population after Tier 1 detection (n = 177 from 7 replicates). Bottom, Hi-C (Nir et al. 2018) data of ChrX-46plex-2K targets. (Spearman’s rank correlation 0.641, two-sided p-value for a hypothesis test whose null hypothesis is that two sets of data are uncorrelated = 7.074e-245). g, ChrX-46plex-2K detection rate for individual cells from 7 replicates. h, Percentage of cells displaying a range of efficiencies of barcode detection after ChrX-46plex-2K O-eLIT. i, Mean spatial distance versus Interaction frequency of Hi-C (Nir et al. 2018) of ChrX-46plex-2K targets. Pearson correlation coefficient (r = −0.84) and p-value = 5.08E-275 (two-sided, using slope = 0 for null hypothesis and Wald Test with t-distribution as test statistic) of the linear least-squares regression. j, Mean spatial distance versus genomic distance for all pairwise ChrX-46plex-2K targets (n = 177 from 7 replicates).

Source data

Extended Data Fig. 7 O-eLIT identifies clusters after ChrX-46plex O-eLIT.

a, Hierarchical clustering based on structure of ChrX traces from ChrX-46plex after 5 rounds of O-eLIT and Tier 2 detection yielded two clusters (Cluster 1 = 20; Cluster 2 = 156). See Methods for more details. b, ChrX representative models (existing traces that are closer to the virtual centroid) of the two clusters obtained after Hierarchical clustering in panel a. c, ChrX-46plex-2K population contact matrix of two clusters derived after Hierarchical clustering in panel a where pairwise spatial distances are considered to be in contact if less than 2 µm apart. d, Radius of gyration for the two clusters (Cluster 1 = 20; Cluster 2 = 156) derived after the hierarchical clustering shown in panel a. The thick line in each violin plot represents the Interquartile range (IQR), the white dot marks the median and the thin lines extend 1.5 times the IQR.

Source data

Extended Data Fig. 8 Angles from 36plex.

a, Measurements of angles formed by three points along the p arm (left), q arm (right), and intersection of vectors formed by pR2-pR3 and qR1-qR2 (middle) for each chromosome. Measurements were obtained by combining data from 36plex-5K and 36plex-1K analyses and selecting chromosomes that had all six targets identified. Chr2: n = 686, Chr3: n = 668, Chr5: n = 363, Chr16: n = 586, Chr19: n = 760, ChrX: n = 493 (n = 1,051 cells from 24 replicates; for 36plex-5K, n = 611 from 15 replicates; for 36plex-1K, n = 440 from 9 replicates). b, Distribution of angles formed by segments in panel a. The thick line in each violin plot represents the Interquartile range (IQR), the white dot marks the median and the thin lines extend 1.5 times the IQR. c, Box plots comparing p and q arm angles. Two-sided student’s t-test with null hypothesis of equal mean was performed to compare arms, ns p > 0.05, * p ≤ 0.05, ** p ≤ 0.01, *** p ≤ 0.001, **** p ≤ 0.0001. Boxes represent the IQR (25th, 50th and 75th percentiles) and whiskers extend 1.5 times the IQR. Sample size information in a). Exact p-values for each chromosome: Chr.2 = 4.149e-16, Chr.3 = 0.004, Chr.5 = 0.093, Chr.16 = 1.357e-14, Chr.19 = 3.325e-11, Chr.X = 0.101. d, Linear least-squares regression between arm angle and arm length with Pearson correlation coefficient r = 0.26 and p-value = 0.42 (two-sided, using slope = 0 for null hypothesis and Wald Test with t-distribution as test statistic).

Source data

Extended Data Fig. 9 O-eLIT comparison of X chromosomes in female IMR-90 cells after ChrX-46plex-2K O-eLIT off of both streets.

a, First round of O-eLIT sequencing. MacroH2A.1 immunostaining after five rounds of O-eLIT marks the Xi. n = 1. b, c, Xi and Xa traces (b) and ball and stick (c) of panel a nucleus after Tier 2 analysis and interpolation of missing targets. Sphere color corresponds to chromosome cartoon. n = 1. d, Single-cell pairwise spatial distances after interpolation of missing targets in panel a. e, Tier 2 target detection efficiency after five rounds of O-eLIT. 38.57% of targeted regions are detected in 71 cells. Detection efficiency from individual replicates are plotted. Error bars represent 95% bootstrap confidence interval of the mean. f, Population pairwise spatial distances after Tier 1 detection (n = 71 cells) and Hi-C data of IMR-90 cells (Rao et al. 2014). g, Population contact maps (top) where two targets are considered to be in contact if less than 2 µm apart (n = 315 chromosomes). Bottom, Hi-C data as in panel f. (Spearman’s rank correlation with the Hi-C matrix is r = 0.733, two-sided p-value for a hypothesis test whose null hypothesis is that two sets of data are uncorrelated = 2.564 ×10e-175). h, Radius of gyration for the Xi (n = 40 chromosomes) and Xa (n = 31 chromosomes). The thick line in each violin plot represents the Interquartile range (IQR), the white dot marks the median and the thin lines extend 1.5 times the IQR. P-value = 7.08 ×10e-6 (two-sided t-test whose null hypothesis is equal means). i, j, Linear plot of the mean spatial distance versus the genomic distance for all pairwise targets for Xi (n = 40 chromosomes) and Xa (n = 31 chromosomes). k–l, Population contact maps for Xi (n = 40 chromosomes) and Xa (n = 31 chromosomes) with eigenvector analysis used to identify different domains. X1-X18 (white) and X19-X46 (grey) targets p and q arms, respectively.

Source data

Extended Data Fig. 10 OligoFISSEQ applications.

a, O-eLIT and immunofluorescence (IF). 36plex-1K was sequenced 5 rounds with O-eLIT off Mainstreet. Then, the same sample was prepared for IF and stained with antibodies. Samples were counterstained with wheat germ agglutinin (WGA) to stain membranes. Images are from deconvolved, maximum z-projections representative of 2 replicates. b, Chromosomal regions imaged with OligoSTORM from Fig. 6d enlarged and displayed separately. Orientation may differ from Fig. 6d. n = 1. c, 8 rounds of O-LIT sequencing of Chr19-9K off of Mainstreet. Images are maximum z-projections. Signal is detectable in all rounds even though the imaging was conducted without the advantage of eLIT, suggesting that 8 rounds of O-eLIT will produce even stronger signals. Images are representative of 2 replicates. d, O-LIT is compatible with gel embedding and target amplification via rolling circle amplification (RCA). Chr19-9K was hybridized to PGP1f cells, after which the sample was embedded in a hydrogel and then cleared of cellular background with proteinase. Next, a molecular inversion probe (MIP) was hybridized to a Chr19-9K specific barcode on Backstreet as well as a fluorophore labeled (purple) secondary oligo to Mainstreet to visualize Chr19-9K Oligopaint oligos. MIPs were circularized via ligation and RCA, after which the first digit of the barcode was sequenced using O-LIT (green). Images are representative of two replicates. e, Comparison of secondary fluorophore signal (2o) versus first round sequencing signal (LIT) from puncta in panel b images. Center values are mean values (3.4 for 2o and 4.9 for O-LIT) with SD.

Source data

Supplementary information

Supplementary Information

Supplementary Figs. 1–4, Supplementary Notes 1 and 2 and Supplementary Protocols 1–5.

Reporting Summary

Supplementary Tables 1, 7 and 8

Lists of reagents, oligonucleotides and O-HIT secondary oligonucleotides used in this study.

Supplementary Tables 2–6 and 16

Oligopaint library sequences for all libraries used in this study.

Supplementary Table 9–11

Image dataset information.

Supplementary Table 12

Genomic and probe information for all libraries used in this study.

Supplementary Table 13

List of Spearman’s rank correlations and two-sided P values for a hypothesis test whose null hypothesis is that two sets of data are uncorrelated.

Supplementary Table 14

Every-pixel analysis intensity thresholds.

Supplementary Table 15

Mean and s.d. of the distance from the center of chromosomes to the nuclear envelope.

Source data

Source Data Fig. 1

O-LIT, O-SIT and O-HIT barcode detection for Chr19-20K library.

Source Data Fig. 3

36plex-5K data.

Source Data Fig. 4

36plex-1K data.

Source Data Fig. 5

ChrX-46plex-2K data.

Source Data Fig. 6

Six-gene library detection data.

Source Data Extended Data Fig. 1

36plex-5K: street sequencing, manual tracing and every-pixel comparison with TrackMate.

Source Data Extended Data Fig. 2

36plex-5K detection data.

Source Data Extended Data Fig. 3

36plex-5K extra cell traces and genome organization comparison.

Source Data Extended Data Fig. 4

36plex chromosome clustering data.

Source Data Extended Data Fig. 5

36plex-1K extra data.

Source Data Extended Data Fig. 6

ChrX-46plex-2K extra data.

Source Data Extended Data Fig. 7

ChrX cluster data.

Source Data Extended Data Fig. 8

36plex angle data.

Source Data Extended Data Fig. 9

ChrX-46plex-2K in IMR-90 data.

Source Data Extended Data Fig. 10

Intensity measurements after rolling circle amplification of Chr19-9K.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Nguyen, H.Q., Chattoraj, S., Castillo, D. et al. 3D mapping and accelerated super-resolution imaging of the human genome using in situ sequencing. Nat Methods 17, 822–832 (2020). https://doi.org/10.1038/s41592-020-0890-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41592-020-0890-0

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