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:

Revealing nanostructures in brain tissue via protein decrowding by iterative expansion microscopy

Abstract

Many crowded biomolecular structures in cells and tissues are inaccessible to labelling antibodies. To understand how proteins within these structures are arranged with nanoscale precision therefore requires that these structures be decrowded before labelling. Here we show that an iterative variant of expansion microscopy (the permeation of cells and tissues by a swellable hydrogel followed by isotropic hydrogel expansion, to allow for enhanced imaging resolution with ordinary microscopes) enables the imaging of nanostructures in expanded yet otherwise intact tissues at a resolution of about 20 nm. The method, which we named ‘expansion revealing’ and validated with DNA-probe-based super-resolution microscopy, involves gel-anchoring reagents and the embedding, expansion and re-embedding of the sample in homogeneous swellable hydrogels. Expansion revealing enabled us to use confocal microscopy to image the alignment of pre-synaptic calcium channels with post-synaptic scaffolding proteins in intact brain circuits, and to uncover periodic amyloid nanoclusters containing ion-channel proteins in brain tissue from a mouse model of Alzheimer’s disease. Expansion revealing will enable the further discovery of previously unseen nanostructures within cells and tissues.

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: ExR, a technology for decrowding of proteins through isotropic protein separation.
Fig. 2: Validation of ExR using synapses and comparison with DNA-PAINT.
Fig. 3: Validation of ExR enhancement and effective resolution in synapses of mouse cortex.
Fig. 4: ExR reveals how calcium channel distributions participate in trans-synaptic nanoarchitecture.
Fig. 5: ExR reveals periodic nanoclusters of Aβ42 peptide in the fornix of Alzheimer’s model 5xFAD mice.
Fig. 6: ExR reveals co-localized clusters of Aβ42 peptide with potassium and sodium ion channels in the fornix of Alzheimer’s model 5xFAD mice.
Fig. 7: Analysis of Aβ42 peptide and potassium ion channel nanoclusters in shapes and their relationships.

Similar content being viewed by others

Data availability

The main data supporting the results in this study are available within the paper and its Supplementary Information. The raw and analysed datasets generated during the study are too large to be publicly shared, yet they are available for research purposes from the corresponding authors on reasonable request. Source data are provided with this paper.

Code availability

The custom code used in this study is available on Zenodo at https://doi.org/10.5281/zenodo.6383293 and on GitHub at https://github.com/schroeme/ExR.

References

  1. Sydor, A. M., Czymmek, K. J., Puchner, E. M. & Mennella, V. Super-resolution microscopy: from single molecules to supramolecular assemblies. Trends Cell Biol. 25, 730–748 (2015).

    Article  CAS  PubMed  Google Scholar 

  2. Wassie, A. T., Zhao, Y. & Boyden, E. S. Expansion microscopy: principles and uses in biological research. Nat. Methods 16, 33–41 (2019).

    Article  CAS  PubMed  Google Scholar 

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

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  4. Chang, J. B. et al. Iterative expansion microscopy. Nat. Methods 14, 593–599 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  5. Chen, F. et al. Nanoscale imaging of RNA with expansion microscopy. Nat. Methods 13, 679–684 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Tillberg, P. W. et al. Protein-retention expansion microscopy of cells and tissues labeled using standard fluorescent proteins and antibodies. Nat. Biotechnol. 34, 987–992 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Zhao, Y. et al. Nanoscale imaging of clinical specimens using pathology-optimized expansion microscopy. Nat. Biotechnol. 35, 757–764 (2017).

    CAS  Google Scholar 

  8. Dani, A., Huang, B., Bergan, J., Dulac, C. & Zhuang, X. Superresolution imaging of chemical synapses in the brain. Neuron 68, 843–856 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Monteiro, P. & Feng, G. SHANK proteins: roles at the synapse and in autism spectrum disorder. Nat. Rev. Neurosci. 18, 147–157 (2017).

    Article  CAS  PubMed  Google Scholar 

  10. Dolphin, A. C. & Lee, A. Presynaptic calcium channels: specialized control of synaptic neurotransmitter release. Nat. Rev. Neurosci. 21, 213–229 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Xiao, B., Cheng, Tu,J. & Worley, P. F. Homer: a link between neural activity and glutamate receptor function. Curr. Opin. Neurobiol. 10, 370–374 (2000).

    Article  CAS  PubMed  Google Scholar 

  12. Szumlinski, K. K., Kalivas, P. W. & Worley, P. F. Homer proteins: implications for neuropsychiatric disorders. Curr. Opin. Neurobiol. 16, 251–257 (2006).

    Article  CAS  PubMed  Google Scholar 

  13. Zhu, F. et al. Architecture of the mouse brain synaptome. Neuron 99, 781–799.e10 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  14. Vazquez, L. E., Chen, H. J., Sokolova, I., Knuesel, I. & Kennedy, M. B. SynGAP regulates spine formation. J. Neurosci. 24, 8862–8872 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Davydova, D. et al. Bassoon specifically controls presynaptic P/Q-type Ca2+ channels via RIM-binding protein. Neuron 82, 181–194 (2014).

    Article  CAS  PubMed  Google Scholar 

  16. Peça, J. et al. Shank3 mutant mice display autistic-like behaviours and striatal dysfunction. Nature 472, 437–442 (2011).

    Article  PubMed  PubMed Central  Google Scholar 

  17. Graf, E. R. et al. RIM promotes calcium channel accumulation at active zones of the Drosophila neuromuscular junction. J. Neurosci. 32, 16586–16596 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Kiyonaka, S. et al. RIM1 confers sustained activity and neurotransmitter vesicle anchoring to presynaptic Ca2+ channels. Nat. Neurosci. 10, 691–701 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  19. Frank, T. et al. Bassoon and the synaptic ribbon organize Ca2+ channels and vesicles to add release sites and promote refilling. Neuron 68, 724–738 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. El-Husseini, A. E., Schnell, E., Chetkovich, D. M., Nicoll, R. A. & Bredt, D. S. PSD-95 involvement in maturation of excitatory synapses. Science 290, 1364–1368 (2000).

    Article  CAS  PubMed  Google Scholar 

  21. Migaud, M. et al. Enhanced long-term potentiation and impaired learning in mice with mutant postsynaptic density-95 protein. Nature 396, 433–439 (1998).

    Article  CAS  PubMed  Google Scholar 

  22. Hayashi, M. K. et al. The postsynaptic density proteins Homer and Shank form a polymeric network structure. Cell 137, 159–171 (2009).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Tang, A. H. et al. A trans-synaptic nanocolumn aligns neurotransmitter release to receptors. Nature 536, 210–214 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Heine, M. & Holcman, D. Asymmetry between pre- and postsynaptic transient nanodomains shapes neuronal communication. Trends Neurosci. 43, 182–196 (2020).

    Article  CAS  PubMed  Google Scholar 

  25. Hruska, M., Henderson, N., Le Marchand, S. J., Jafri, H. & Dalva, M. B. Synaptic nanomodules underlie the organization and plasticity of spine synapses. Nat. Neurosci. 21, 671–682 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Rebola, N. et al. Distinct nanoscale calcium channel and synaptic vesicle topographies contribute to the diversity of synaptic function. Neuron 104, 693–710.e9 (2019).

    Article  CAS  PubMed  Google Scholar 

  27. Brockmann, M. M. et al. RIM-BP2 primes synaptic vesicles via recruitment of Munc13-1 at hippocampal mossy fiber synapses. eLife 8, e43243 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  28. Holderith, N. et al. Release probability of hippocampal glutamatergic terminals scales with the size of the active zone. Nat. Neurosci. 15, 988–997 (2012).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Eggermann, E., Bucurenciu, I., Goswami, S. P. & Jonas, P. Nanodomain coupling between Ca2+ channels and sensors of exocytosis at fast mammalian synapses. Nat. Rev. Neurosci. 13, 7–21 (2012).

    Article  CAS  Google Scholar 

  30. Canter, R. G., Penney, J. & Tsai, L. H. The road to restoring neural circuits for the treatment of Alzheimer’s disease. Nature 539, 187–196 (2016).

    Article  PubMed  Google Scholar 

  31. Oakley, H. et al. Intraneuronal β-amyloid aggregates, neurodegeneration, and neuron loss in transgenic mice with five familial Alzheimer’s disease mutations: potential factors in amyloid plaque formation. J. Neurosci. 26, 10129–10140 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Chao, L. L. et al. Associations between white matter hyperintensities and β amyloid on integrity of projection, association, and limbic fiber tracts measured with diffusion tensor MRI. PLoS ONE 8, e65175 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Song, S. K., Kim, J. H., Lin, S. J., Brendza, R. P. & Holtzman, D. M. Diffusion tensor imaging detects age-dependent white matter changes in a transgenic mouse model with amyloid deposition. Neurobiol. Dis. 15, 640–647 (2004).

    Article  CAS  PubMed  Google Scholar 

  34. Dong, J. W. et al. Diffusion MRI biomarkers of white matter microstructure vary nonmonotonically with increasing cerebral amyloid deposition. Neurobiol. Aging 89, 118–128 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  35. Gail Canter, R. et al. 3D mapping reveals network-specific amyloid progression and subcortical susceptibility in mice. Commun. Biol. 2, 1–12 (2019).

    Article  CAS  Google Scholar 

  36. Dunn, A. R. & Kaczorowski, C. C. Regulation of intrinsic excitability: roles for learning and memory, aging and Alzheimer’s disease, and genetic diversity. Neurobiol. Learn. Mem. 164, 107069 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Chong, S. Y. C. et al. Neurite outgrowth inhibitor Nogo-A establishes spatial segregation and extent of oligodendrocyte myelination. Proc. Natl Acad. Sci. USA 109, 1299–1304 (2012).

    Article  CAS  PubMed  Google Scholar 

  38. Brohawn, S. G. et al. The mechanosensitive ion channel traak is localized to the mammalian node of Ranvier. eLife 8, 1–22 (2019).

    Article  Google Scholar 

  39. Dupree, J. L. et al. Oligodendrocytes assist in the maintenance of sodium channel clusters independent of the myelin sheath. Neuron Glia Biol. 1, 179–192 (2004).

    Article  PubMed  PubMed Central  Google Scholar 

  40. Lubetzki, C., Sol-Foulon, N. & Desmazières, A. Nodes of Ranvier during development and repair in the CNS. Nat. Rev. Neurol. 16, 426–439 (2020).

    Article  PubMed  Google Scholar 

  41. Shah, N. H. & Aizenman, E. Voltage-gated potassium channels at the crossroads of neuronal function, ischemic tolerance, and neurodegeneration. Transl. Stroke Res. 5, 38–58 (2014).

    Article  CAS  PubMed  Google Scholar 

  42. Hessler, S. et al. β-secretase BACE1 regulates hippocampal and reconstituted M-currents in a β-subunit-like fashion. J. Neurosci. 35, 3298–3311 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  43. Ciccone, R. et al. Amyloid β-induced upregulation of Nav1.6 underlies neuronal hyperactivity in Tg2576 Alzheimer’s disease mouse model. Sci. Rep. 9, 1–18 (2019).

    Article  CAS  Google Scholar 

  44. Ghatak, S. et al. Mechanisms of hyperexcitability in Alzheimer’s disease hiPSC-derived neurons and cerebral organoids vs. isogenic control. eLife 8, e50333 (2019).

    Article  PubMed  PubMed Central  Google Scholar 

  45. Lim, C. J., Lee, S. Y., Teramoto, J., Ishihama, A. & Yan, J. The nucleoid-associated protein Dan organizes chromosomal DNA through rigid nucleoprotein filament formation in E. coli during anoxia. Nucleic Acids Res. 41, 746–753 (2013).

    Article  CAS  PubMed  Google Scholar 

  46. Xu, K., Zhong, G. & Zhuang, X. Actin, spectrin, and associated proteins form a periodic cytoskeletal structure in axons. Science 339, 452–456 (2013).

    Article  CAS  PubMed  Google Scholar 

  47. Winardhi, R. S., Castang, S., Dove, S. L. & Yan, J. Single-molecule study on histone-like nucleoid-structuring protein (H-NS) paralogue in Pseudomonas aeruginosa: MvaU bears DNA organization mode similarities to MvaT. PLoS ONE 9, e112246 (2014).

    Article  PubMed  PubMed Central  Google Scholar 

  48. Leterrier, C. et al. Nanoscale architecture of the axon initial segment reveals an organized and robust scaffold. Cell Rep. 13, 2781–2793 (2015).

    Article  CAS  PubMed  Google Scholar 

  49. Chiang, Y. L. et al. Atomic force microscopy characterization of protein fibrils formed by the amyloidogenic region of the bacterial protein MinE on mica and a supported lipid bilayer. PLoS ONE 10, e0142506 (2015).

    Article  PubMed  PubMed Central  Google Scholar 

  50. Prakash, K. et al. Superresolution imaging reveals structurally distinct periodic patterns of chromatin along pachytene chromosomes. Proc. Natl Acad. Sci. USA 112, 14635–14640 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Makky, A., Bousset, L., Polesel-Maris, J. & Melki, R. Nanomechanical properties of distinct fibrillar polymorphs of the protein α-synuclein. Sci. Rep. 6, 37970 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  52. D’Este, E., Kamin, D., Göttfert, F., El-Hady, A. & Hell, S. W. STED nanoscopy reveals the ubiquity of subcortical cytoskeleton periodicity in living neurons. Cell Rep. 10, 1246–1251 (2015).

    Article  PubMed  Google Scholar 

  53. D’Este, E. et al. Subcortical cytoskeleton periodicity throughout the nervous system. Sci. Rep. 6, (2016).

  54. Qu, Y., Hahn, I., Webb, S. E. D., Pearce, S. P. & Prokop, A. Periodic actin structures in neuronal axons are required to maintain microtubules. Mol. Biol. Cell 28, 296–308 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  55. Bose, K., Lech, C. J., Heddi, B. & Phan, A. T. High-resolution AFM structure of DNA G-wires in aqueous solution. Nat. Commun. 9, 1959 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  56. Ku, T. et al. Multiplexed and scalable super-resolution imaging of three-dimensional protein localization in size-adjustable tissues. Nat. Biotechnol. 34, 973–981 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  57. M’Saad, O. & Bewersdorf, J. Light microscopy of proteins in their ultrastructural context. Nat. Commun. https://doi.org/10.1038/s41467-020-17523-8 (2020).

  58. Gambarotto, D. et al. Imaging cellular ultrastructures using expansion microscopy (U-ExM). Nat. Methods 16, 71–74 (2019).

    Article  CAS  PubMed  Google Scholar 

  59. Zwettler, F. U. et al. Molecular resolution imaging by post-labeling expansion single-molecule localization microscopy (Ex-SMLM). Nat. Commun. 11, 1–11 (2020).

    Article  Google Scholar 

  60. Chen, F. et al. Nanoscale imaging of RNA with expansion microscopy. Nat. Methods 13, 679–684 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  61. Klapoetke, N. C. et al. Independent optical excitation of distinct neural populations. Nat. Methods 11, 338–346 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  62. Saka, S. K. et al. Immuno-SABER enables highly multiplexed and amplified protein imaging in tissues. Nat. Biotechnol. 37, 1080–1090 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  63. Schnitzbauer, J. et al. Super-resolution microscopy with DNA-PAINT. Nat. Protoc. 12, 1198–1228 (2017).

    Article  CAS  PubMed  Google Scholar 

  64. Savtchenko, L. P. & Rusakov, D. A. The optimal height of the synaptic cleft. Proc. Natl Acad. Sci. USA 104, 1823–1828 (2007).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Motulsky, H. J. & Brown, R. E. Detecting outliers when fitting data with nonlinear regression - a new method based on robust nonlinear regression and the false discovery rate. BMC Bioinformatics 7, 123 (2006).

    Article  PubMed  PubMed Central  Google Scholar 

  66. Chen, J. H., Blanpied, T. A. & Tang, A. H. Quantification of trans-synaptic protein alignment: a data analysis case for single-molecule localization microscopy. Methods 174, 72–80 (2020).

    Article  CAS  PubMed  Google Scholar 

  67. McQuin, C. et al. CellProfiler 3.0: next-generation image processing for biology. PLoS Biol. 16, e2005970 (2018).

    Article  PubMed  PubMed Central  Google Scholar 

  68. Mangan, A. P. & Whitaker, R. T. Partitioning 3D surface meshes using watershed segmentation. IEEE Trans. Vis. Comput. Graph. 5, 308–321 (1999).

    Article  Google Scholar 

Download references

Acknowledgements

We thank T. Biederer, C. Zhang and Y. Liu for antibody advice, S. Alon and K. Piatkevich for trainings and discussions; B. Kang for decrowding analysis advice; and D. Goodwin for instructions and scripts for manual image registration. D.S. acknowledges funding from NIH K99/R00 Pathway to independence Award (K99GM126277 and R00GM126277); M.E.S was supported by the National Science Foundation Graduate Research Fellowship under Grant No. 1745302 and the MathWorks Science Fellowship; A.-H.T. acknowledges funding from NARSAD; E.D.N. acknowledges funding from Alana Down Syndrome Center and Robert A. and Renee E. Belfer Foundation; P.Y. acknowledges funding from NIH RF1 MH124606 and NIH Director’s Pioneer Award DP1 GM133052; L.-H.T. acknowledges funding from Robert A and Renee E Belfer Family Foundation, Alana Down Syndrome Center, Ludwig Family Foundation, the JPB Foundation, NIH Grants RO1 NS102730 and RF1 AG054321; T.A.B acknowledges funding from NIMH R37MH080046 and R01MH119826; E.S.B. acknowledges funding from the Ludwig Family Foundation, the Open Philanthropy Project, John Doerr, Lisa Yang and the Tan-Yang Center for Autism Research at MIT, the US Army Research Laboratory and the US Army Research Office under contract/grant number W911NF1510548; Charles Hieken, Tom Stocky, Kathleen Octavio, Lore McGovern, Good Ventures; NIH Director’s Pioneer Award 1DP1NS087724, NIH R01MH110932, NIH R01EB024261, NIH U24NS109113, NIH R37MH08004613, NIH 1RM1HG008525, NIH 1R01AG070831, NIH 1R56AG069192 and NIH R01MH124606; funding from the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (grant agreement No 835102); and HHMI. We also thank B. Chow and E. Betzig, as well as many current and past members of the Synthetic Neurobiology group, for discussions. The funders had no role in the study design, data collection and analysis, decision to publish or preparation of the manuscript.

Author information

Authors and Affiliations

Authors

Contributions

D.S. initiated work, developed the ExR technology, contributed key ideas, designed and performed experiments and interpreted data for all projects, and wrote and edited the manuscript. J.K. contributed key ideas, designed and performed experiments for all projects, performed analysis for decrowding, and wrote and edited the manuscript. A.T.W. co-developed approaches to measuring decrowding, contributed key ideas to ExR technology development and decrowding analysis, designed and performed experiments, analysed data for all projects, and wrote and edited the manuscript. M.E.S contributed key ideas, designed and implemented analysis, visualization and statistical tests for all projects, created the schematic in Fig. 1 with input from others, and wrote and edited the manuscript. Z.P. contributed key ideas, designed and performed experiments, and conducted analysis for the Alzheimer’s project. T.B.T. contributed key ideas, designed experiments, created software, analysed and interpreted data related to synapses, and helped in writing and editing the manuscript. A.-H.T. created software and analysed data related to synapses. E.D.N. designed experiments, interpreted experimental data for the Alzheimer’s project, and helped in writing and editing the manuscript. J.Z.Y. oversaw the Alzheimer’s project and designed experiments for it. H.S. performed DNA-PAINT validation experiments. D.P. prepared cultured neurons. P.Y. contributed key ideas and designed experiments for DNA-PAINT validation. L.-H.T. conceptualized and initiated the Alzheimer’s project, provided oversight and funding, contributed key ideas and designed experiments, and helped in writing and editing the manuscript. T.A.B. contributed key ideas, designed experiments and interpreted data for the synapse project, and helped in writing and editing the manuscript. E.S.B. supervised the project, initiated work, contributed key ideas, designed experiments, helped with data analysis and interpretation, and wrote and edited the manuscript.

Corresponding authors

Correspondence to Li-Huei Tsai, Thomas A. Blanpied or Edward S. Boyden.

Ethics declarations

Competing interests

D.S., A.T.W., J.K. and E.S.B. are co-inventors on a patent application for ExR (US 2020/0271556 A1). E.S.B. is co-founder of a company seeking to deploy applications of ExM-related technologies. The other authors declare no competing interests.

Peer review

Peer review information

Nature Biomedical Engineering thanks Laurent Nguyen, Markus Sauer and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Additional information

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

Extended data

Extended Data Fig. 1 Comparison of synaptic nanostructures imaged using DNA-PAINT and ExR in cultured neurons.

(a-c) Three representative fields of view imaged using DNA-PAINT (i) and ExR (ii) after rigid body registration to DNA-PAINT images. Scale bar = 5 μm. (iii-iv) Processed, binary versions of (i) and (ii) used to automatically count synaptic puncta number and pairwise distances. Scale bar, 5μm. (d) Representative manually-cropped matched synaptic ROIs for DNA-PAINT (top row) and ExR (bottom row), used for the distortion analysis shown in Fig. 2h,i (scale bar = 250 nm). (e) Pixel-wise correlation between min-max normalized ExR and DNA-PAINT channels as a function of shift distance in x- and y-directions for two randomly selected synaptic ROIs. (f) Pixel-wise autocorrelation between min-max normalized DNA-PAINT (PAINT-PAINT, magenta), ExR (ExR-ExR, yellow), and pixel-wise correlation between DNA-PAINT and ExR (PAINT-ExR, black) as a function of shift distance in x- and y-directions for the synaptic ROIs shown in (e). (g) Representative manually-cropped pairs of synaptic puncta used to generate the data shown in Fig. 2j and panel i. Shown is an overlay of DNA-PAINT (green) and ExR (magenta) binary masks (scale bar = 250 nm). (h) Histogram of difference in number of synaptic puncta counted after thresholding pairs of synaptic puncta. (i) Difference in radial distance between pairs of synaptic puncta, DNA-PAINT – ExR (mean = -0.008854, 95% CI [-0.05419, 0.03649]). (j) Total number of synaptic puncta for the five fields of view imaged using DNA-PAINT and ExR (two-sided paired t-test, p = 0.9271, t = 0.09735, df = 4). All data are from 5 ROIs from 3 wells of one cultured neuron batch. Shown are images from one representative experiment from two independent replicates.

Source data

Extended Data Fig. 2 Analysis of the ExR decrowding effect.

(a, b) Quantification of decrowding in a set of manually identified synapses. Statistical significance was determined using Sidak’s multiple comparisons test on two-sided t-tests following ANOVA (*P ≤ 0.05, **P ≤ 0.01, ***P < 0.001, ****P < 0.0001, here and throughout the paper, and plotted is the mean, with error bars representing standard error of the mean (SEM), here and throughout the paper). (a) Mean signal intensity inside and outside of (that is, nearby to) dilated reference ROIs for pre- and post-expansion stained manually-identified synapses (Supplementary Table 2 for numbers of technical and biological replicates)). Data points represent the mean across all synapses from a single field of view. (b) Total volume (in voxels; 1 voxel = 17.16 × 17.15 × 40, or 11,779, nm3) of signals inside and outside of dilated reference ROIs, in both cases within cropped images containing one visually identified synapse (Supplementary Fig. 3h), for pre- and post-expansion stained manually-identified synapses (Supplementary Table 2 for numbers of biological and technical replicates). Data points represent the mean across all synapses for 3 fields of view from 3 biological replicates (n = 9 fields of view total per protein). (c) Mean voxel size and (d) mean signal-to-noise (SNR) ratio of pre- and post-expansion immunostaining showing 7 proteins in somatosensory cortex regions L1, L2/3, and L4 of 3 mice. Plotted is mean and SEM. To compare the 3D voxel size and SNR of pre- and post-expansion stained synapses for each of the seven proteins, three two-sided t-tests (one for each layer) were run (n = 49–70 puncta per layer from 3 mice; Supplementary Table 2 for exact n values). Statistical significance was determined using multiple t-tests corrected using the Holm-Sidak method, with alpha = 0.05. (e) Population distribution (violin plot of density, with a dashed line at the median and dotted lines at the quartiles) of the fractional difference in the number of synaptic puncta between post- and pre-expansion staining channels for Homer1 and Shank3 (n = 480 synapses from 3 mice). (f) Population distribution of the difference in distance (in nm) between the shift at which the correlation is half maximal half-maximal shift for pre-pre autocorrelation and post-pre correlation (calculated pixel-wise between intensity values normalized to the minimum and maximum of the image, see Methods) for x-, y-, and z-directions (x- and y-directions being transverse, z-direction being axial) for Homer1 and Shank3 (n = 458 synapses, 3 directions each, from 3 mice). (g) Same as (f), for post-post autocorrelation and pre-post correlation.

Source data

Extended Data Fig. 3 Analysis of the distortion caused by post-expansion staining, as compared to classical pre-expansion staining.

(a-b) Representative background-subtracted and binary images of Homer1 (a) and Shank3 (b) in pre- and post-expansion staining channels (top row (yellow): pre-expansion channel, second row (black/white): binary pre-expansion channel, third row (magenta): post-expansion channel, bottom row (black/white): binary post-expansion channel). (c) Number of synaptic puncta for pre- and post-expansion staining channels, after filtration, for the images in (a-b). (d) Population distribution (violin plot of density, with a dashed line at the median and dotted lines at the quartiles) of the number of synaptic puncta in the pre-expansion staining channel for Homer1 and Shank3 (see Supplementary Table 5 for statistics for this figure). (e) Population distribution of the number of synaptic puncta in the post-expansion staining channel for Homer1 and Shank3. (f) Difference in the number of synaptic puncta between post- and pre-expansion staining channels normalized to the number of synaptic puncta in the pre-expansion staining channel. (g-h) Pixel-wise autocorrelation between pre-expansion (pre-pre, yellow), post-expansion (post-post, magenta), and pixel-wise correlation between pre- and post-expansion (pre-post, black) as a function of shift distance in x- (left column), y- (middle column), and z- (right column) directions for Homer1 (g) and Shank3 (h). The mean across all synapses is shown in the top row, and representative synapses are shown in the second through fourth rows. These values were used to calculate the linearized error measure shown in Extended Data Fig. 2f, g. (i-j) Pixel-wise correlation between mean-normalized, masked pre- and post-expansion channels as a function of shift distance in x- and y-directions (z = 1) for Homer1 (i) and Shank3 (j). The mean across all synapses is shown in the top left, standard deviation across all synapses shown in second from the top left, and representative synapses are shown in the remaining plots. (k-l) Mutually overlapped volume between pre- and post-expansion stained synaptic puncta, normalized to total puncta volume in the pre-expansion staining channel, as a function of shift distance in x- and y-directions (z = 1) for Homer1 (k) and Shank3 (l). The mean across all synapses is shown in the top left, standard deviation across all synapses shown in second from the top left, and representative synapses are shown in the remaining plots. Analysis was conducted on 480 (before exclusion based on size) synapses for Shank3 and Homer1 from 3 mice (see Supplementary Table 5 for exact numbers).

Source data

Extended Data Fig. 4 ExR and unexpanded tissue confocal images showing immunolabeling of Aβ42.

ExR confocal images (single z-slices) showing immunolabeling of Aβ42 with two different monoclonal antibodies (a) D54D2 + 6E10 and (b) D54D2 + 12F4 with SMI co-staining in the fornix of 5xFAD mouse (n = 3 fields of view of 2 slices from 2 mice). Scale bar, 10 μm (top row), 1 μm (i, ii panels). (c) Unexpanded tissue confocal image, a single z-slice, showing pre-expansion Aβ42 (yellow) and SMI (cyan) staining in the fornix of WT (upper panel) and 5xFAD mice (lower panel) (n = 3 fields of view of 1 slice from 1 mouse per WT and 5xFAD, respectively). Scale bar, 30 μm (left panel) and 6 μm (panels i, ii)).

Extended Data Fig. 5 ExR confocal images showing immunolabeling of PLP, SMI, and 12F4 in 5xFAD and WT fornix.

ExR reveals relative localization of Aβ42 peptide and myelin in the fornix of Alzheimer’s model 5xFAD and WT mice (n = 2 fields of view of 1 slice from 2 mice per WT and 5xFAD, respectively). (a) ExR confocal image (max intensity projections, 900–1000 nm thickness) showing post-expansion Aβ42 (magenta), SMI (cyan) and PLP (green) staining in the fornix of 5xFAD mice. Leftmost panel, merged low-magnification image; right images show individual channels. Insets (i-iii) show close-up views of the boxed regions highlighted in the upper left image. (b) ExR confocal image (max intensity projections, 1.72 μm thickness) showing post-expansion Aβ42 (magenta), SMI (cyan) and PLP (green) staining in the fornix of wild-type mice. Leftmost panel, merged image. All images were subtracted background using Fiji’s Rolling Ball algorithm with radius 50 pixels, and adjusted with auto-contrast. Scale bar = 500 nm. (c) Comparison of 12F4, SMI, and PLP intensity levels along axons in 5xFAD fornix with and without 12F4. To analyze axonal amyloid beta deposition with myelination and SMI intensity, we measure the (i) 12F4, (ii) SMI and (iii) PLP intensity levels along 10 axons with (12F4+) and without 12F4 (12F4-) from the same field of view (n = 3 fields of view of 2 slices from 2 5xFAD). All images were subtracted background with 50 pixels, and adjusted with auto-contrast for analysis by ImageJ. On each axon, three lines were drawn cross-sectionally across each axon in Image J and averaged intensity levels of PLP, 12F4 and SMI from different channels were measured respectively along these lines. For 12F4 + axons, each line was drawn across the centroid of amyloid beta deposition. For 12F4- axons, lines were positioned randomly along the axon. We then compared PLP, 12F4 and SMI312 intensity levels between 12F4 + and 12F4- axons. Plotted is the mean, with error bars representing standard error of the mean (SEM). Two-sided paired t-test, (i) ****p < 0.0001, t = 6.112, df = 18, (ii) p = 0.0595, t = 2.012, df = 18, (iii) p = 0.6580, t = 0.4502, df = 18.

Source data

Supplementary information

Supplementary Information

Supplementary figures and tables.

Reporting Summary

Source data

Rights and permissions

Springer Nature or its licensor holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Sarkar, D., Kang, J., Wassie, A.T. et al. Revealing nanostructures in brain tissue via protein decrowding by iterative expansion microscopy. Nat. Biomed. Eng 6, 1057–1073 (2022). https://doi.org/10.1038/s41551-022-00912-3

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41551-022-00912-3

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