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.

  • Letter
  • Published:

Variable chromatin structure revealed by in situ spatially correlated DNA cleavage mapping

Abstract

Chromatin structure at the length scale encompassing local nucleosome–nucleosome interactions is thought to play a crucial role in regulating transcription and access to DNA1,2,3. However, this secondary structure of chromatin remains poorly understood compared with the primary structure of single nucleosomes or the tertiary structure of long-range looping interactions4. Here we report the first genome-wide map of chromatin conformation in human cells at the 1–3 nucleosome (50–500 bp) scale, obtained using ionizing radiation-induced spatially correlated cleavage of DNA with sequencing (RICC-seq) to identify DNA–DNA contacts that are spatially proximal. Unbiased analysis of RICC-seq signal reveals regional enrichment of DNA fragments characteristic of alternating rather than adjacent nucleosome interactions in tri-nucleosome units, particularly in H3K9me3-marked heterochromatin. We infer differences in the likelihood of nucleosome–nucleosome contacts among open chromatin, H3K27me3-marked, and H3K9me3-marked repressed chromatin regions. After calibrating RICC-seq signal to three-dimensional distances, we show that compact two-start helical fibre structures with stacked alternating nucleosomes are consistent with RICC-seq fragmentation patterns from H3K9me3-marked chromatin, while non-compact structures and solenoid structures are consistent with open chromatin. Our data support a model of chromatin architecture in intact interphase nuclei consistent with variable longitudinal compaction of two-start helical fibres.

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

Access options

Rent or buy this article

Prices vary by article type

from$1.95

to$39.95

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

Figure 1: RICC-seq principle and assay.
Figure 2: The RICC-seq FLD reveals variable chromatin compaction states.
Figure 3: Spatially correlated break frequency scales exponentially with 3D distance allowing calibration of signal.
Figure 4: RICC-seq identifies distinct classes of predicted fibre models correlated with epigenetic features.

Similar content being viewed by others

Accession codes

Primary accessions

Gene Expression Omnibus

References

  1. Li, G. & Reinberg, D. Chromatin higher-order structures and gene regulation. Curr. Opin. Genet. Dev. 21, 175–186 (2011)

    Article  CAS  Google Scholar 

  2. Luger, K., Dechassa, M. L. & Tremethick, D. J. New insights into nucleosome and chromatin structure: an ordered state or a disordered affair? Nature Rev. Mol. Cell Biol. 13, 436–447 (2012)

    Article  CAS  Google Scholar 

  3. Gilbert, N. et al. Chromatin architecture of the human genome: gene-rich domains are enriched in open chromatin fibers. Cell 118, 555–566 (2004)

    Article  CAS  Google Scholar 

  4. Risca, V. I. & Greenleaf, W. J. Unraveling the 3D genome: genomics tools for multiscale exploration. Trends Genet. 31, 357–372 (2015)

    Article  CAS  Google Scholar 

  5. Bednar, J. et al. Nucleosomes, linker DNA, and linker histone form a unique structural motif that directs the higher-order folding and compaction of chromatin. Proc. Natl Acad. Sci. USA 95, 14173–14178 (1998)

    Article  ADS  CAS  Google Scholar 

  6. Dorigo, B. et al. Nucleosome arrays reveal the two-start organization of the chromatin fiber. Science 306, 1571–1573 (2004)

    Article  ADS  CAS  Google Scholar 

  7. Horowitz, R. A., Agard, D. A., Sedat, J. W. & Woodcock, C. L. The three-dimensional architecture of chromatin in situ: electron tomography reveals fibers composed of a continuously variable zig-zag nucleosomal ribbon. J. Cell Biol. 125, 1–10 (1994)

    Article  CAS  Google Scholar 

  8. Schalch, T., Duda, S., Sargent, D. F. & Richmond, T. J. X-ray structure of a tetranucleosome and its implications for the chromatin fibre. Nature 436, 138–141 (2005)

    Article  ADS  CAS  Google Scholar 

  9. Finch, J. T. & Klug, A. Solenoidal model for superstructure in chromatin. Proc. Natl Acad. Sci. USA 73, 1897–1901 (1976)

    Article  ADS  CAS  Google Scholar 

  10. Robinson, P. J., Fairall, L., Huynh, V. A. & Rhodes, D. EM measurements define the dimensions of the “30-nm” chromatin fiber: evidence for a compact, interdigitated structure. Proc. Natl Acad. Sci. USA 103, 6506–6511 (2006)

    Article  ADS  CAS  Google Scholar 

  11. Widom, J. & Klug, A. Structure of the 300A chromatin filament: X-ray diffraction from oriented samples. Cell 43, 207–213 (1985)

    Article  CAS  Google Scholar 

  12. Routh, A., Sandin, S. & Rhodes, D. Nucleosome repeat length and linker histone stoichiometry determine chromatin fiber structure. Proc. Natl Acad. Sci. USA 105, 8872–8877 (2008)

    Article  ADS  CAS  Google Scholar 

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

    Article  ADS  CAS  Google Scholar 

  14. Hsieh, T.-H. S. et al. Mapping nucleosome resolution chromosome folding in yeast by micro-C. Cell 162, 108–119 (2015)

    Article  CAS  Google Scholar 

  15. Rydberg, B., Holley, W. R., Mian, I. S. & Chatterjee, A. Chromatin conformation in living cells: support for a zig-zag model of the 30 nm chromatin fiber. J. Mol. Biol. 284, 71–84 (1998)

    Article  CAS  Google Scholar 

  16. Scheffer, M. P., Eltsov, M. & Frangakis, A. S. Evidence for short-range helical order in the 30-nm chromatin fibers of erythrocyte nuclei. Proc. Natl Acad. Sci. USA 108, 16992–16997 (2011)

    Article  ADS  CAS  Google Scholar 

  17. Grigoryev, S. A. et al. Hierarchical looping of zigzag nucleosome chains in metaphase chromosomes. Proc. Natl Acad. Sci. USA 113, 1238–1243 (2016)

    Article  ADS  CAS  Google Scholar 

  18. Eltsov, M., Maclellan, K. M., Maeshima, K., Frangakis, A. S. & Dubochet, J. Analysis of cryo-electron microscopy images does not support the existence of 30-nm chromatin fibers in mitotic chromosomes in situ. Proc. Natl Acad. Sci. USA 105, 19732–19737 (2008)

    Article  ADS  CAS  Google Scholar 

  19. Nishino, Y. et al. Human mitotic chromosomes consist predominantly of irregularly folded nucleosome fibres without a 30-nm chromatin structure. EMBO J. 31, 1644–1653 (2012)

    Article  CAS  Google Scholar 

  20. Alpen, E. L. in Radiation Biophysics, 2nd edn, 104–131 (Academic, 1998)

    Chapter  Google Scholar 

  21. Roots, R., Kraft, G. & Gosschalk, E. The formation of radiation-induced DNA breaks: the ratio of double-strand breaks to single-strand breaks. Int. J. Radiat. Oncol. Biol. Phys. 11, 259–265 (1985)

    Article  CAS  Google Scholar 

  22. Valouev, A. et al. Determinants of nucleosome organization in primary human cells. Nature 474, 516–520 (2011)

    Article  CAS  Google Scholar 

  23. Buenrostro, J. D., Giresi, P. G., Zaba, L. C., Chang, H. Y. & Greenleaf, W. J. Transposition of native chromatin for fast and sensitive epigenomic profiling of open chromatin, DNA-binding proteins and nucleosome position. Nature Methods 10, 1213–1218 (2013)

    Article  CAS  Google Scholar 

  24. Tullius, T. D. & Dombroski, B. A. Hydroxyl radical “footprinting”: high-resolution information about DNA-protein contacts and application to lambda repressor and Cro protein. Proc. Natl Acad. Sci. USA 83, 5469–5473 (1986)

    Article  ADS  CAS  Google Scholar 

  25. Bintu, L. et al. Dynamics of epigenetic regulation at the single-cell level. Science 351, 720–724 (2016)

    Article  ADS  CAS  Google Scholar 

  26. Davey, C. A., Sargent, D. F., Luger, K., Maeder, A. W. & Richmond, T. J. Solvent mediated interactions in the structure of the nucleosome core particle at 1.9 a resolution. J. Mol. Biol. 319, 1097–1113 (2002)

    Article  CAS  Google Scholar 

  27. Koslover, E. F., Fuller, C. J., Straight, A. F. & Spakowitz, A. J. Local geometry and elasticity in compact chromatin structure. Biophys. J. 99, 3941–3950 (2010)

    Article  ADS  CAS  Google Scholar 

  28. Widom, J. A relationship between the helical twist of DNA and the ordered positioning of nucleosomes in all eukaryotic cells. Proc. Natl Acad. Sci. USA 89, 1095–1099 (1992)

    Article  ADS  CAS  Google Scholar 

  29. Song, F. et al. Cryo-EM study of the chromatin fiber reveals a double helix twisted by tetranucleosomal units. Science 344, 376–380 (2014)

    Article  ADS  CAS  Google Scholar 

  30. Fan, J. Y., Rangasamy, D., Luger, K. & Tremethick, D. J. H2A.Z alters the nucleosome surface to promote HP1α-mediated chromatin fiber folding. Mol. Cell 16, 655–661 (2004)

    Article  CAS  Google Scholar 

  31. Zhou, B.-R. et al. Structural mechanisms of nucleosome recognition by linker histones. Mol. Cell 59, 628–638 (2015)

    Article  CAS  Google Scholar 

  32. Grigoryev, S. A., Arya, G., Correll, S., Woodcock, C. L. & Schlick, T. Evidence for heteromorphic chromatin fibers from analysis of nucleosome interactions. Proc. Natl Acad. Sci. USA 106, 13317–13322 (2009)

    Article  ADS  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  34. Sambrook, J. & Russell, D. W. Isolation of DNA fragments from polyacrylamide gels by the crush and soak method. CSH Protoc. 2006, http://dx.doi.org/10.1101/pdb.prot2936 (2006)

  35. Smith, D. J. & Whitehouse, I. Intrinsic coupling of lagging-strand synthesis to chromatin assembly. Nature 483, 434–438 (2012)

    Article  ADS  CAS  Google Scholar 

  36. Corces, M. R. et al. Lineage-specific and single-cell chromatin accessibility charts human hematopoiesis and leukemia evolution. Nature Genet. 48, 1193–1203 (2016)

    Article  CAS  Google Scholar 

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

    Article  Google Scholar 

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

    Article  CAS  Google Scholar 

  39. ENCODE Project Consortium. An integrated encyclopedia of DNA elements in the human genome. Nature 489, 57–74 (2012)

  40. Kheradpour, P. & Kellis, M. Systematic discovery and characterization of regulatory motifs in ENCODE TF binding experiments. Nucleic Acids Res. 42, 2976–2987 (2014)

    Article  CAS  Google Scholar 

  41. Kundaje, A. et al.; Roadmap Epigenomics Consortium et al. Integrative analysis of 111 reference human epigenomes. Nature 518, 317–330 (2015)

    Article  CAS  Google Scholar 

  42. Zhang, Y. et al. Model-based analysis of ChIP-Seq (MACS). Genome Biol. 9, R137 (2008)

    Article  Google Scholar 

  43. Schep, A. N. et al. Structured nucleosome fingerprints enable high-resolution mapping of chromatin architecture within regulatory regions. Genome Res. 25, 1757–1770 (2015)

    Article  CAS  Google Scholar 

  44. Henikoff, J. G., Belsky, J. A., Krassovsky, K., MacAlpine, D. M. & Henikoff, S. Epigenome characterization at single base-pair resolution. Proc. Natl Acad. Sci. USA 108, 18318–18323 (2011)

    Article  ADS  CAS  Google Scholar 

  45. Guse, A., Fuller, C. J. & Straight, A. F. A cell-free system for functional centromere and kinetochore assembly. Nature Protocols 7, 1847–1869 (2012)

    Article  CAS  Google Scholar 

  46. Yoon, S. et al. HiTRACE: high-throughput robust analysis for capillary electrophoresis. Bioinformatics 27, 1798–1805 (2011)

    Article  CAS  Google Scholar 

  47. Eichman, B. F., Vargason, J. M., Mooers, B. H. & Ho, P. S. The Holliday junction in an inverted repeat DNA sequence: sequence effects on the structure of four-way junctions. Proc. Natl Acad. Sci. USA 97, 3971–3976 (2000)

    Article  ADS  CAS  Google Scholar 

  48. Edgar, R., Domrachev, M. & Lash, A. E. Gene Expression Omnibus: NCBI gene expression and hybridization array data repository. Nucleic Acids Res. 30, 207–210 (2002)

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We acknowledge R. Das for conversations and help with capillary sequencing, E. Koslover, A. Spakowitz, and A. Schep for sharing code and sample data, discussions, and reading the manuscript, A. Kathiria for technical assistance, J. Buenrostro and C. Araya for advice and sharing code, I. Whitehouse for sharing a protocol, C. Fuller for nucleosome reconstitution help, P. Zhu and G. Li for sharing their chromatin fibre model, T. Phillips, G. J. Gu, O. Rando and W. Johnson for discussions, and G. Wang for BJ cells. BJ-5ta cells were a gift from J. Cochran. V.I.R. acknowledges the support of the Walter V. and Idun Berry Postdoctoral Fellowship. S.K.D. acknowledges support from a National Institutes of Health (NIH) Predoctoral Molecular Biophysics Training Program grant to Stanford University and from a National Science Foundation Graduate Fellowship. A.F.S. acknowledges support from National Institutes of Health (NIH) grant R01GM106005. W.J.G. acknowledges NIH grants R21HG007726, and P50HG00773501. This work was supported by the Rita Allen Foundation, the Baxter Foundation, and the Human Frontier Science Program.

Author information

Authors and Affiliations

Authors

Contributions

V.I.R., S.K.D., A.F.S., and W.J.G. designed experiments. V.I.R. and S.K.D. performed experiments. V.I.R. and S.K.D. analysed data. V.I.R., A.F.S., and W.J.G. interpreted the results and wrote the paper.

Corresponding author

Correspondence to William J. Greenleaf.

Ethics declarations

Competing interests

Stanford University has filed a patent application on these results with V.I.R., S.J.K.D., and W.J.G. named as co-inventors.

Additional information

Reviewer Information

Nature thanks S. Grigoryev and the other anonymous reviewer(s) for their contribution to the peer review of this work.

Extended data figures and tables

Extended Data Figure 1 RICC-seq library preparation workflow.

Cells lysed in agarose plugs after irradiation are processed within the plugs to avoid mechanical shear of DNA. The 3′ phosphate ends (and all phosphates) are removed by alkaline phosphatase treatment, then 5′ phosphates are added back by polynucleotide kinase (T4 PNK) in preparation for ligation. Short 5′ phosphorylated ssDNA fragments are then passively eluted from agarose plugs after thermal denaturation and incorporated into paired-end sequencing libraries by ligation of random hexamer overhang adapters. Adaptor dimers are removed by size selection before and after PCR amplification of libraries.

Extended Data Figure 2 Reproducibility, sequence bias, robustness to crosslinking, and genome coverage of RICC-seq data sets.

a, Irradiated cell samples (300 Gy and 0 Gy) compared between two technical replicates and 300 Gy fragment size distributions compared after subtraction of scaled 0 Gy histograms. b, Fragment size distributions compared between three biological replicates. Owing to their low random DNA cleavage baselines, experiments 1 and 2 were used for genome-wide fragment length profile analysis and calibration of the exponential cleavage model (Figs 2, 3, 4). Experiment 3 has a higher background of non-specific breaks, probably because of mild nuclease contamination during the plug processing steps of the protocol. Although the low signal-to-background ratio of this experiment led us to exclude it from genome-wide distribution analysis, we did include it in V-plot and footprint generation to obtain higher sequencing coverage of ~78 bp fragments (Fig. 1 and Extended Data Fig. 3). Background subtraction of 0 Gy signal is as in a. c, Fragment end sequence bias for all irradiation conditions in an experiment with machine-mixed random hexamer adaptor oligonucleotides. d, Fragment end sequence bias in biological replicate processed with hand-mixed random hexamer adaptor oligonucleotides. e, Raw fragment size distributions obtained from live and formaldehyde/EGS-crosslinked cells (subsampled to same sequencing depth). f, Normalized expected RICC-seq fragment density from cellular chromatin (red) or dechromatinized DNA (black) estimated by dividing the number of fragments within annotated peaks by the number within size-matched control regions elsewhere in the genome (error bars represent s.e.m. with n = 3 independent experiments; centre value is mean). g, Ratio of cellular fragment density ratio to genomic DNA fragment density ratio from f. h, P values for comparisons in g (Mann–Whitney U test; n = 3 biological replicates).

Extended Data Figure 3 RICC-seq produces characteristic 78 nt fragments around positioned nucleosomes and protection footprints at CTCF-bound sites.

a, V-plot of dechromatinized DNA control for Fig. 1f. b, V-plot of RICC-seq fragments from cells around unoccupied CTCF motifs without a CTCF ChIP-seq peak (see Methods). c, Difference in counts of RICC-seq fragment ends piled up over aggregated CTCF motifs within (bound) or outside (not bound) ChIP peaks. d, V-plot of cell and dechromatinized DNA around active (associated with RNA-CAGE peaks) transcription start sites (see Methods). e, V-plot of cell fragments around inactive TSSs (lacking an ATAC-seq peak in BJ fibroblasts). Grey ovals are cartoon representations of nucleosome positions and phasing. Data for all panels were pooled from three biological replicates.

Extended Data Figure 4 Calculation of FLFE normalizes sequence bias in the RICC-seq fragment size distribution throughout the genome.

a, Workflow for calculating and normalizing FLFE. b, Fragment length distributions in 1 Mb windows sorted by GC content. c, FLFE with respect to lookup table of GC content-binned genome-wide average FLD in equal-sized windows. d, Z-scores of chromatin annotations in same 1 Mb windows. See Methods for annotation details. Data are pooled from two biological replicates.

Extended Data Figure 5 FLFE PC1 and regional profiles are reproducible across biological replicates.

a, PC1 calculated independently from each of two biological replicates. b, Region FLFE profiles calculated for additional annotations. ATAC-seq, H3K9me3-only, and H3K27me3 (excluding H3K9me3 or H3K4me1) are as in Fig. 2. Euchromatin-associated annotations are on a blue background. c, For two annotation regions, reproducibility of regional FLFE profiles between two biological replicates is shown in terms of the raw FLFE profiles (top), the FLFE in de-chromatinized DNA samples that quantifies sequence bias (middle), and the GC baseline normalized FLFE (bottom). Grey shaded areas represent peaks as in Fig. 1.

Extended Data Figure 6 PC1 signal is correlated with ATAC-seq signal and anti-correlated with H3K9me3 ChIP signal.

a, b, Example tracks of RICC-seq PC1 signal for combined biological replicates and each replicate separately in the same region of chromosome 5 as in Fig. 2 (at higher zoom) and in a region of chromosome 3. c, Correlation between replicates for PC1 score. d, Correlation of PC1 score from combined replicates with euchromatic (ATAC-seq) and heterochromatic (H3K9me3 or H3K27me3) annotations in 300 kb windows. ATAC-seq data are shown as insertion density, while H3K9me3 and H3K27me3 signals are shown as −log10(Poisson P value over local background) (see Methods).

Extended Data Figure 7 Fragment length bias correction.

Bias was corrected by comparing fragment size distributions between sequencing data (a) and intensity profiles from 5′ end-labelled denaturing gel electrophoresis of RICC-seq ssDNA fragments before sequencing library preparation (b). c, End-labelled fragments have a length distribution weighted more towards long fragments. d, The ratio of the two distributions (sequencing/gel) was calculated and smoothed, then (e) fitted with a single exponential to obtain an estimate of the length bias. Bias estimate is shown for three biological replicates.

Extended Data Figure 8 RICC-seq signal from positioned nucleosomes reflects known nucleosome structure.

a, V-plot of RICC-seq fragments around ATAC-seq called nucleosome positions. b, Exponential fits of background-subtracted and length bias corrected (Extended Data Fig. 7) fragment counts to the 3D distance between the end positions of each fragment shown for individual replicates and individual strands. c, Example showing the magnitude of predicted FLD changes for a simulated fibre structure (see Methods) between the calibration values obtained from biological replicate experiments. d, Cleavage locations (fragment ends) were mapped around nucleosome dyad centres called by ATAC-seq (see Methods). Background-subtracted (0 Gy), sequencing depth, and cell-number-scaled fragment end counts are shown for the positive and negative reference genome strand. Blue trace shows cleavage frequency from in vitro hydroxyl-radical footprinted reconstituted nucleosome (scaled to comparable amplitude).

Extended Data Figure 9 Clustering fibre structures into five classes on the basis of helical symmetry.

a, Example fibre structure illustrating nucleosome and fibre coordinates. b, Energy (log10(kBT)) per nucleosome for each minimum-energy structure. c, Distribution of linker DNA crossing distance for each minimum-energy structure at every linker length and compaction level combination. d, e, Maps of nucleosome tilt angles (nucleosome axis relative to fibre axis) (d) and rotation angles around fibre for adjacent nucleosomes (e). f, g, Distance between centres of second-nearest-neighbour (f, N to N + 2) and third-nearest-neighbour (g, N to N + 3) nucleosomes in each fibre model. h, Hierarchical clustering of the 969 candidate structures into five classes on the basis of the distribution of nucleosome centre distances for intervals between N to N + 1 (adjacent nucleosomes) and N to N + 7. The matrix was row-standardized before clustering to measure the relative intensity of contact distances for each structure. Classes are coloured as in Fig. 4b. i, Mean with bands (one s.d.) of predicted FLFE plotted for each class of structure. j, Mean energy of the structures in each class. k, Comparison of predicted FLD between a solenoid structure and an extended zig-zag structure without length bias correction for 51–2,000 bp fragments. l, Measured RICC-seq FLD beyond 500 nt before (left) and after (right) subtraction of an exponential baseline. k, Comparison of predicted FLD between a solenoid structure and an extended zig-zag structure without length bias correction for 51–2,000 bp fragments. m, Example FLFEs of structure models with varying linker length. n, Example FLFEs of structure models with varying compaction level.

Extended Data Figure 10 Two-start helix structures (class 1) are consistent with genome-wide average chromatin structure, facultative heterochromatin, and a reconstituted fibre containing linker histone, while bent linkers are consistent with open chromatin.

a, Map of similarity scores between the predicted FLFEs of simulated fibre structures and the genome-wide experimental FLFE, shown in c, measured as the ratio of the genome-wide FLD to the genome-wide dechromatinized DNA FLD. b, Mean similarity scores as calculated in a for each structure class. c, Genome-wide experimental FLFE calculated with respect to the dechromatinized DNA FLD after a one-parameter fit to obtain equivalent scaling of read counts. d, Top view and side view of reconstituted chromatin fibre with NRL = 187 bp (40 bp linker length)29. e, Similarity score between predicted FLDs for each simulated fibre structure and the FLFE of the reconstituted fibre structure. Linker length of the reconstituted fibre is indicated. f, Mean similarity scores from e, for each structure class. g, Predicted FLD for the reconstituted fibre and genome-wide experimental FLD. h, Fragments with characteristic lengths (red) labelled on the tetranucleosome unit of the reconstituted fibre. i, Similarity score between predicted FLFE for the fibre structure candidates and the experimental FLFE from H3K27me3 peaks with no H3K9me3 or H3K4 me1 ChIP peak overlap, with (j) averages of similarity score within clusters as in Fig. 4b. Class 1 regions (Fig. 4b) are outlined in black in a, e, and i. k, Comparison of FLFEs from H3K27me3 ChIP peaks with H3K9me3 FLFE and predicted FLFE from a high-similarity-score structure. l, Mononucleosome models built by adding 10 bp straight DNA linker pieces to achieve 30 bp linker lengths. Structures are based on the core nucleosome crystal structure, chromatosome crystal structure with linker histone H5, and a chromatin fibre with linker histone H1 (see Methods). The predicted FLD and FLFE with respect to the genome-wide aggregate FLD is shown for each structure. m, The FLFE and FLD in three chromatin states shown for comparison. n, Predicted chromatosome FLDs for extended zig-zag and solenoid structures. Curves in ln are scaled by the maximum of the 78 bp peak.

Supplementary information

Supplementary Information

This file contains Supplementary Tables 1-2. (PDF 46 kb)

PowerPoint slides

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Risca, V., Denny, S., Straight, A. et al. Variable chromatin structure revealed by in situ spatially correlated DNA cleavage mapping. Nature 541, 237–241 (2017). https://doi.org/10.1038/nature20781

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature20781

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

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