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:

SR-Tesseler: a method to segment and quantify localization-based super-resolution microscopy data

Abstract

Localization-based super-resolution techniques open the door to unprecedented analysis of molecular organization. This task often involves complex image processing adapted to the specific topology and quality of the image to be analyzed. Here we present a segmentation framework based on Voronoï tessellation constructed from the coordinates of localized molecules, implemented in freely available and open-source SR-Tesseler software. This method allows precise, robust and automatic quantification of protein organization at different scales, from the cellular level down to clusters of a few fluorescent markers. We validated our method on simulated data and on various biological experimental data of proteins labeled with genetically encoded fluorescent proteins or organic fluorophores. In addition to providing insight into complex protein organization, this polygon-based method should serve as a reference for the development of new types of quantifications, as well as for the optimization of existing ones.

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

Figure 1: Voronoï-based segmentation.
Figure 2: Automatic segmentation of various proteins and cell types.
Figure 3: Segmentation and quantification of simulations and well-characterized nanotemplates.
Figure 4: Segmentation and quantification of experimental PALM data.
Figure 5: Segmentation and quantification of experimental dSTORM data.

Similar content being viewed by others

References

  1. Betzig, E. et al. Imaging intracellular fluorescent proteins at nanometer resolution. Science 313, 1642–1645 (2006).

    Article  CAS  Google Scholar 

  2. Hess, S.T., Girirajan, T.P. & Mason, M.D. Ultra-high resolution imaging by fluorescence photoactivation localization microscopy. Biophys. J. 91, 4258–4272 (2006).

    Article  CAS  Google Scholar 

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

    Article  CAS  Google Scholar 

  4. Heilemann, M. et al. Subdiffraction-resolution fluorescence imaging with conventional fluorescent probes. Angew. Chem. Int. Edn. Engl. 47, 6172–6176 (2008).

    Article  CAS  Google Scholar 

  5. Folling, J. et al. Fluorescence nanoscopy by ground-state depletion and single-molecule return. Nat. Methods 5, 943–945 (2008).

    Article  Google Scholar 

  6. Klein, T., Proppert, S. & Sauer, M. Eight years of single-molecule localization microscopy. Histochem. Cell Biol. 141, 561–575 (2014).

    Article  CAS  Google Scholar 

  7. Sage, D. et al. Quantitative evaluation of software packages for single-molecule localization microscopy. Nat. Methods 12, 717–724 (2015).

    Article  CAS  Google Scholar 

  8. Ripley, B.D. Modelling spatial patterns. J. R. Stat. Soc. B 39, 172–212 (1977).

    Google Scholar 

  9. Scarselli, M., Annibale, P. & Radenovic, A. Cell type–specific β2-adrenergic receptor clusters identified using photoactivated localization microscopy are not lipid raft related, but depend on actin cytoskeleton integrity. J. Biol. Chem. 287, 16768–16780 (2012).

    Article  CAS  Google Scholar 

  10. Muranyi, W., Malkusch, S., Müller, B., Heilemann, M. & Kräusslich, H.-G. Super-resolution microscopy reveals specific recruitment of HIV-1 envelope proteins to viral assembly sites dependent on the envelope c-terminal tail. PLoS Pathog. 9, e1003198 (2013).

    Article  Google Scholar 

  11. Sengupta, P. et al. Probing protein heterogeneity in the plasma membrane using PALM and pair correlation analysis. Nat. Methods 8, 969–975 (2011).

    Article  CAS  Google Scholar 

  12. Ester, M., Kriegel, H.-P., Jörg, S. & Xu, X. A density-based algorithm for discovering clusters in large spatial databases with noise. Proc. 2nd International Conference on Knowledge Discovery and Data Mining 226–231 (1996).

  13. Endesfelder, U. et al. Multiscale spatial organization of RNA polymerase in Escherichia coli. Biophys. J. 105, 172–181 (2013).

    Article  CAS  Google Scholar 

  14. Nan, X. et al. Single-molecule superresolution imaging allows quantitative analysis of RAF multimer formation and signaling. Proc. Natl. Acad. Sci. USA 110, 18519–18524 (2013).

    Article  CAS  Google Scholar 

  15. Li, W. et al. Single protein molecule detection by glass nanopores. ACS Nano 7, 4129–4134 (2013).

    Article  CAS  Google Scholar 

  16. Deschout, H., Shivanandan, A., Annibale, P., Scarselli, M. & Radenovic, A. Progress in quantitative single-molecule localization microscopy. Histochem. Cell Biol. 142, 5–17 (2014).

    Article  CAS  Google Scholar 

  17. Baddeley, D., Cannell, M.B. & Soeller, C. Visualization of localization microscopy data. Microsc. Microanal. 16, 64–72 (2010).

    Article  CAS  Google Scholar 

  18. Okabe, A., Boots, B. & Sugihara, K. Spatial Tessellations: Concepts and Applications of Voronoi Diagrams (John Wiley & Sons, 1992).

  19. Garrido, S., Moreno, L., Abderrahim, M. & Martin, F. Path planning for mobile robot navigation using Voronoi diagram and fast marching. in Proc. International Conference on Intelligent Robots and Systems 2376–2381 (IEEE, 2006).

  20. Alliez, P., Colin de Verdière, E., Devillers, O. & Isenburg, M. Isotropic surface remeshing. Proc. Shape Modeling International 49–58 (IEEE Computer Society, 2003).

  21. Beheiry, M.E. & Dahan, M. ViSP: representing single-particle localizations in three dimensions. Nat. Methods 10, 689–690 (2013).

    Article  Google Scholar 

  22. Baddeley, D. et al. Optical single-channel resolution imaging of the ryanodine receptor distribution in rat cardiac myocytes. Proc. Natl. Acad. Sci. USA 106, 22275–22280 (2009).

    Article  CAS  Google Scholar 

  23. Fortune, S. A sweepline algorithm for Voronoï diagram. Proc. Second Annual Symposium on Computational Geometry 313–322 (ACM, 1986).

  24. Nair, D. et al. Super-resolution imaging reveals that AMPA receptors inside synapses are dynamically organized in nanodomains regulated by PSD95. J. Neurosci. 33, 13204–13224 (2013).

    Article  CAS  Google Scholar 

  25. Rossier, O. et al. Integrins β1 and β3 exhibit distinct dynamic nanoscale organizations inside focal adhesions. Nat. Cell Biol. 14, 1057–1067 (2012).

    Article  CAS  Google Scholar 

  26. Schneider, C.A., Rasband, W.S. & Eliceiri, K.W. NIH Image to ImageJ: 25 years of image analysis. Nat. Methods 9, 671–675 (2012).

    Article  CAS  Google Scholar 

  27. Williamson, D.J. et al. Pre-existing clusters of the adaptor Lat do not participate in early T cell signaling events. Nat. Immunol. 12, 655–662 (2011).

    Article  CAS  Google Scholar 

  28. Owen, D.M., Williamson, D.J., Magenau, A. & Gaus, K. Sub-resolution lipid domains exist in the plasma membrane and regulate protein diffusion and distribution. Nat. Commun. 3, 1256 (2012).

    Article  Google Scholar 

  29. Durisic, N., Laparra-Cuervo, L., Sandoval-Alvarez, A., Borbely, J.S. & Lakadamyali, M. Single-molecule evaluation of fluorescent protein photoactivation efficiency using an in vivo nanotemplate. Nat. Methods 11, 156–162 (2014).

    Article  CAS  Google Scholar 

  30. Annibale, P., Vanni, S., Scarselli, M., Rothlisberger, U. & Radenovic, A. Identification of clustering artifacts in photoactivated localization microscopy. Nat. Methods 8, 527–528 (2011).

    Article  CAS  Google Scholar 

  31. Macgillavry, H.D., Song, Y., Raghavachari, S. & Blanpied, T.A. Nanoscale scaffolding domains within the postsynaptic density concentrate synaptic AMPA receptors. Neuron 78, 615–622 (2013).

    Article  CAS  Google Scholar 

  32. Fukata, Y. et al. Local palmitoylation cycles define activity-regulated postsynaptic subdomains. J. Cell Biol. 202, 145–161 (2013).

    Article  CAS  Google Scholar 

  33. Hosy, E., Butler, C. & Sibarita, J.B. Organization and dynamics of AMPA receptors inside synapses—nano-organization of AMPA receptors and main synaptic scaffolding proteins revealed by super-resolution imaging. Curr. Opin. Chem. Biol. 20, 120–126 (2014).

    Article  CAS  Google Scholar 

  34. Giannone, G. et al. Dynamic superresolution imaging of endogenous proteins on living cells at ultra-high density. Biophys. J. 99, 1303–1310 (2010).

    Article  CAS  Google Scholar 

  35. Kiskowski, M.A., Hancock, J.F. & Kenworthy, A.K. On the use of Ripley's K-function and its derivatives to analyze domain size. Biophys. J. 97, 9 (2009).

    Article  Google Scholar 

  36. Samet, H. The Quadtree and related hierarchical data structures. ACM Comput. Surv. 16, 187–260 (1984).

    Article  Google Scholar 

  37. Lee, S.-H., Shin, J.Y., Lee, A. & Bustamante, C. Counting single photoactivatable fluorescent molecules by photoactivated localization microscopy (PALM). Proc. Natl. Acad. Sci. USA 109, 17436–17441 (2012).

    Article  CAS  Google Scholar 

  38. Puchner, E.M., Walter, J.M., Kasper, R., Huang, B. & Lim, W.A. Counting molecules in single organelles with superresolution microscopy allows tracking of the endosome maturation trajectory. Proc. Natl. Acad. Sci. USA 110, 16015–16020 (2013).

    Article  CAS  Google Scholar 

  39. Thompson, R.E., Larson, D.R. & Webb, W.W. Precise nanometer localization analysis for individual fluorescent probes. Biophys. J. 82, 2775–2783 (2002).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

We thank D. Nair and K. Haas for the GluA1-mEOS2 and GPI-mEOS2 acquisitions, G. Giannone and O. Rossier for the integrin-β3−mEOS2 PALM data and M. Lakadamyali for providing GlyR data and feedback on the manuscript. This work was supported by the Ministère de l'Enseignement Supérieur et de la Recherche (ANR NanoDom, Labex BRAIN and ANR-10-INBS-04 France-BioImaging), the European Research Council (ERC; grant nano-dyn-syn to D.C.), the Centre National de la Recherche Scientifique, the Conseil Régional d'Aquitaine and the Institut National de la Santé et de la Recherche Médicale.

Author information

Authors and Affiliations

Authors

Contributions

F.L. developed the software and carried out the simulations. F.L. and J.-B. S. designed the analysis method. E.H. and D.C. designed the biological experiments. A.K. developed the single-molecule localization software. C.B. and A.B. worked on the quantitative analysis and the molecular counting. All the authors contributed to the manuscript. J.-B.S. came up with the original idea and supervised the work.

Corresponding author

Correspondence to Jean-Baptiste Sibarita.

Ethics declarations

Competing interests

The authors declare no competing financial interests.

Integrated supplementary information

Supplementary Figure 1 Segmentation of various proteins and cell types using SR-Tesseler and several intensity-based thresholding methods.

(a–d) Automatic segmentation obtained with SR-Tesseler using the local 1st rank order density feature > 2δ (left column) and four intensity-based thresholding techniques (Isodata=cyan, Intermode=green, Huang=magenta and Percentile=red, see http://fiji.sc/Auto_Threshold#Default) implemented in ImageJ (right column). In each case, SR-Tesseler identified the different levels of molecular organization. Using the same automatic density parameters, we could successfully segmented small clusters, non-isotropic structures such as microtubules or adhesion sites, and the cell contour. This illustrates the versatility and robustness to molecular organization of our method. On the contrary, results obtained with the four intensity-based thresholding techniques provided much more arbitrary and variable results. For example, while the Isodata method shows good segmentation performance on microtubules, it failed segmenting correctly any organization level in neurons and fibroblast cells. Percentile and Huang gave very similar segmentations on the microtubule data but very different results on neuronal and fibroblast data. The Intermode method identified efficiently some nanoclusters in neuronal data but failed to capture any other structure on fibroblast and microtubule data.

Supplementary Figure 2 Comparison between density and cutting distance and DBSCAN.

(a) Segmentation comparison between density ( > 2δ, top-red outlines) and cutting distance ( < 30 nm bottom-magenta outlines) based analysis performed on simulated data. Simulated clusters (cyan outlines) have a density of 0.01 mol/nm², a radius of 50 nm and an enrichment factor R of ∞ (left), 20 (center) and 10 (right). In absence of background, the density criterion tends to underestimate the cluster size, while the use of the cutting distance defined from the localization accuracy is more accurate. When R decreases, the cutting distance criterion tends to fuse the clusters with the background localizations, quickly overestimating their sizes, while the density criterion performs robustly. (b) Segmentation comparison between the density-based spatial clustering analysis with noise (DBSCAN) (left) and SR-Tesseler (right) methods performed on 50 nm radius clusters of 0.01 mol/nm² density. Data with R = 2.5 (top) could be segmented using a set of (ε, Nε) parameters ((40, 80), (30, 40) and (20, 20) from left to right), where ε and Nε denote the neighborhood’s radius (in nm) and the minimum number of localizations in the neighborhood ε, respectively, two parameters of the DBSCAN method12. Using the same set of parameters on data with lower background (R = 20), all the segmentation failed (bottom). In comparison, SR-Tesseler analysis, performed with a unique density parameter > 2δ, could successfully segment all the clusters (right).

Supplementary Figure 3 Comparison with the Ripley function and benchmarking.

(a) Degradation of the simulations by randomizing the localization positions with various accuracies (0, 5, 10 and 20 nm) for different R ratios of 25, 10 and 5 (left to right). (b) H-Ripley functions for various simulation and experimental datasets. While a peak of maximal aggregation is observed on simulations data containing only sparse clusters, no clustering is visible in any experimental data of GPI::mEOS2 and GluA1::mEOS2 expressed in neurons. (c) On experimental data, when performed on ROIs centered on spines, Ripley function could measure clusters of 91.2 nm (IQR 83.2-112 nm). (d) Computation time of the H-Ripley function for different searching radii and different implementations. Since the Voronoï diagram is a space partitioning technique, it can be used to speed-up the H-Ripley function brute-force implementation. The Quadtree space-partitioning technique performs even faster.

Source data

Supplementary Figure 4 Multiple-localization correction.

(a–e) Multiple detections occurring in a vicinity of space (ω) and time (τ) are identified as the same molecule and merged. (a) Fluorophore photophysical parameters, such as the number of blinks undergone by the molecules (Nblinks), the fluorescent on-time (Ton) before a blinking or photobleaching event and the fluorescent off-time (Toff), are determined for each localization. (b) Four detections (3 squares and 1 triangle) are located within their respective space vicinity ω but only 3 of them (squares) are within the blinking tolerance time interval τ(for clarity, the neighborhood radius of the triangle was not displayed). Their corresponding Voronoï diagram is represented in red. Square localizations are then identified as the same molecule, merged together and replaced by a new detection (circle) at a location corresponding to their barycenter. The new Voronoï diagram is displayed in black. (c–e) To determine the correct time interval τ, the photophysics of the fluorophore is analyzed. From the initial dataset, the off-time (c), number of blinks per molecule (d) and on-time (e) distributions are computed. For a dataset composed of 473,983 localizations, the average number of blinks per molecule was 1.4, and the number of molecules after cleaning was 220,775. As a control, the number of emission bursts (319,473), counted with τ= 0, divided by the average number of blinks per molecule (1.4) was only 3.36% different.

Supplementary information

Supplementary Text and Figures

Supplementary Figures 1–4 (PDF 1280 kb)

Supplementary Software

SR-Tesseler software (ZIP 20052 kb)

Source data

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Levet, F., Hosy, E., Kechkar, A. et al. SR-Tesseler: a method to segment and quantify localization-based super-resolution microscopy data. Nat Methods 12, 1065–1071 (2015). https://doi.org/10.1038/nmeth.3579

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nmeth.3579

This article is cited by

Search

Quick links

Nature Briefing AI and Robotics

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

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