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.

  • Analysis
  • Published:

A comparison of microfluidic methods for high-throughput cell deformability measurements

Abstract

The mechanical phenotype of a cell is an inherent biophysical marker of its state and function, with many applications in basic and applied biological research. Microfluidics-based methods have enabled single-cell mechanophenotyping at throughputs comparable to those of flow cytometry. Here, we present a standardized cross-laboratory study comparing three microfluidics-based approaches for measuring cell mechanical phenotype: constriction-based deformability cytometry (cDC), shear flow deformability cytometry (sDC) and extensional flow deformability cytometry (xDC). All three methods detect cell deformability changes induced by exposure to altered osmolarity. However, a dose-dependent deformability increase upon latrunculin B-induced actin disassembly was detected only with cDC and sDC, which suggests that when exposing cells to the higher strain rate imposed by xDC, cellular components other than the actin cytoskeleton dominate the response. The direct comparison presented here furthers our understanding of the applicability of the different deformability cytometry methods and provides context for the interpretation of deformability measurements performed using different platforms.

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: Comparison of the microfluidics-based approaches for the determination of cell deformability used in this study.
Fig. 2: Effects of osmolarity changes on cell deformability.
Fig. 3: Effects of LatB-induced actin disassembly on cell deformability.

Similar content being viewed by others

Data availability

Source data for Figs. 13 are provided alongside the manuscript. The dataset containing numerical values from all deformability measurements supporting findings of this study is available on figshare60.

Code availability

MATLAB and R codes used to perform statistical analysis and generate data representations shown in this manuscript are available on GitHub at https://github.com/dicarlo-lab/metadeformability.

References

  1. Di Carlo, D. A mechanical biomarker of cell state in medicine. J. Lab. Autom. 17, 32–42 (2012).

    PubMed  Google Scholar 

  2. Nematbakhsh, Y. & Lim, C. T. Cell biomechanics and its applications in human disease diagnosis. Acta Mech. Sin. 31, 268–273 (2015).

    CAS  Google Scholar 

  3. Darling, E. M. & Di Carlo, D. High-throughput assessment of cellular mechanical properties. Annu. Rev. Biomed. Eng. 17, 35–62 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  4. Otto, O. et al. Real-time deformability cytometry: on-the-fly cell mechanical phenotyping. Nat. Methods 12, 199–202 (2015).

    CAS  PubMed  Google Scholar 

  5. Guck, J. et al. Optical deformability as an inherent cell marker for testing malignant transformation and metastatic competence. Biophys. J. 88, 3689–3698 (2005).

    CAS  PubMed  PubMed Central  Google Scholar 

  6. Swaminathan, V. et al. Mechanical stiffness grades metastatic potential in patient tumor cells and in cancer cell lines. Cancer Res. 71, 5075–5080 (2011).

    CAS  PubMed  PubMed Central  Google Scholar 

  7. Byun, S. et al. Characterizing deformability and surface friction of cancer cells. Proc. Natl Acad. Sci. USA 110, 7580–7585 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  8. Tse, H. T. K. et al. Quantitative diagnosis of malignant pleural effusions by single-cell mechanophenotyping. Sci. Transl. Med. 5, 212ra163 (2013).

    PubMed  Google Scholar 

  9. Bufi, N. et al. Human primary immune cells exhibit distinct mechanical properties that are modified by inflammation. Biophys. J. 108, 2181–2190 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  10. Bashant, K. R. et al. Real-time deformability cytometry reveals sequential contraction and expansion during neutrophil priming. J. Leukoc. Biol. 105, 1143–1153 (2019).

    CAS  PubMed  Google Scholar 

  11. Gossett, D. R. et al. Hydrodynamic stretching of single cells for large population mechanical phenotyping. Proc. Natl Acad. Sci. USA 109, 7630–7635 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  12. Toepfner, N. et al. Detection of human disease conditions by single-cell morpho-rheological phenotyping of blood. eLife 7, e29213 (2018).

    PubMed  PubMed Central  Google Scholar 

  13. Rosenbluth, M. J., Lam, W. A. & Fletcher, D. A. Analyzing cell mechanics in hematologic diseases with microfluidic biophysical flow cytometry. Lab Chip 8, 1062–1070 (2008).

    CAS  PubMed  PubMed Central  Google Scholar 

  14. Ekpenyong, A. E. et al. Viscoelastic properties of differentiating blood cells are fate- and function-dependent. PLoS ONE 7, e45237 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  15. Urbanska, M. et al. Single-cell mechanical phenotype is an intrinsic marker of reprogramming and differentiation along the mouse neural lineage. Development 144, 4313–4321 (2017).

    CAS  PubMed  Google Scholar 

  16. Lin, J. et al. High-throughput physical phenotyping of cell differentiation. Microsyst. Nanoeng. 3, 17013 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  17. Mammoto, T. & Ingber, D. E. Mechanical control of tissue and organ development. Development 137, 1407–1420 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  18. Chan, C. J., Heisenberg, C. P. & Hiiragi, T. Coordination of morphogenesis and cell-fate specification in development. Curr. Biol. 27, R1024–R1035 (2017).

    CAS  PubMed  Google Scholar 

  19. Radmacher, M. Studying the mechanics of cellular processes by atomic force microscopy. Methods Cell Biol. 83, 347–372 (2007).

    CAS  PubMed  Google Scholar 

  20. Hochmuth, R. M. Micropipette aspiration of living cells. J. Biomech. 33, 15–22 (2000).

    CAS  PubMed  Google Scholar 

  21. Guck, J. et al. The optical stretcher: a novel laser tool to micromanipulate cells. Biophys. J. 81, 767–784 (2001).

    CAS  PubMed  PubMed Central  Google Scholar 

  22. Thoumine, O., Ott, A., Cardoso, O. & Meister, J.-J. Microplates: a new tool for manipulation and mechanical perturbation of individual cells. J. Biochem. Biophys. Methods 39, 47–62 (1999).

    CAS  PubMed  Google Scholar 

  23. Wu, P. H. et al. A comparison of methods to assess cell mechanical properties. Nat. Methods 15, 491–498 (2018).

    CAS  PubMed  PubMed Central  Google Scholar 

  24. Adamo, A. et al. Microfluidics-based assessment of cell deformability. Anal. Chem. 84, 6438–6443 (2012).

    CAS  PubMed  PubMed Central  Google Scholar 

  25. Lange, J. R. et al. Microconstriction arrays for high-throughput quantitative measurements of cell mechanical properties. Biophys. J. 109, 26–34 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  26. Nyberg, K. D. et al. Quantitative deformability cytometry: rapid, calibrated measurements of cell mechanical properties. Biophys. J. 113, 1574–1584 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  27. Guillou, L. et al. Measuring cell viscoelastic properties using a microfluidic extensional flow device. Biophys. J. 111, 2039–2050 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  28. Armistead, F. J., De Pablo, J. G., Gadêlha, H., Peyman, S. A. & Evans, S. D. Cells under stress: an inertial-shear microfluidic determination of cell behaviour. Biophys. J. 4, 1127–1135 (2019).

    Google Scholar 

  29. Golfier, S. et al. High-throughput cell mechanical phenotyping for label-free titration assays of cytoskeletal modifications. Cytoskeleton 74, 283–296 (2017).

    CAS  PubMed  Google Scholar 

  30. Di Carlo, D. Inertial microfluidics. Lab Chip 9, 3038 (2009).

    PubMed  Google Scholar 

  31. Guilak, F., Erickson, G. K. & Ting-Beall, H. P. The effects of osmotic stress on the viscoelastic and physical properties of articular chondrocytes. Biophys. J. 82, 720–727 (2002).

    CAS  PubMed  PubMed Central  Google Scholar 

  32. Zhou, E. H. et al. Universal behavior of the osmotically compressed cell and its analogy to the colloidal glass transition. Proc. Natl Acad. Sci. USA 106, 10632–10637 (2009).

    CAS  PubMed  PubMed Central  Google Scholar 

  33. Moeendarbary, E. et al. The cytoplasm of living cells behaves as a poroelastic material. Nat. Mater. 12, 253–261 (2013).

    CAS  PubMed  PubMed Central  Google Scholar 

  34. Guo, M. et al. Cell volume change through water efflux impacts cell stiffness and stem cell fate. Proc. Natl Acad. Sci. USA 114, E8618–E8627 (2017).

    CAS  PubMed  PubMed Central  Google Scholar 

  35. Silverthorn, D. U., Johnson, B. R., Ober, W. C., Garrison, C. W. & Silverthorn, A. C. Human Physiology: An Integrated Approach Ch. 5 (Pearson Education, Boston, 2013).

  36. Wakatsuki, T., Schwab, B., Thompson, N. C. & Elson, E. L. Effects of cytochalasin D and latrunculin B on mechanical properties of cells. J. Cell Sci. 114, 1025–1036 (2001).

    CAS  PubMed  Google Scholar 

  37. Salbreux, G., Charras, G. & Paluch, E. Actin cortex mechanics and cellular morphogenesis. Trends Cell Biol. 22, 536–545 (2012).

    CAS  PubMed  Google Scholar 

  38. Spector, I., Shorlet, N. R., Blasberger, D. & Kashman, Y. Latrunculins—novel marine macrolides that disrupt microfilament organization and affect cell growth: I. Comparison with cytochalasin D. Cell Motil. Cytoskeleton 13, 127–144 (1989).

    CAS  PubMed  Google Scholar 

  39. Morton, W. M., Ayscough, K. R. & Mclaughlin, P. J. Latrunculin alters the actin-monomer subunit interface to prevent polymerization. Nat. Cell Biol. 2, 376–378 (2000).

    CAS  PubMed  Google Scholar 

  40. Guck, J. & Chilvers, E. R. Mechanics meets medicine. Sci. Transl. Med. 5, 3–6 (2013).

    Google Scholar 

  41. Finan, J. D. & Guilak, F. The effects of osmotic stress on the structure and function of the cell nucleus. J. Cell. Biochem. 109, 460–467 (2010).

    CAS  PubMed  PubMed Central  Google Scholar 

  42. Hallows, K. R., Packman, C. H. & Knauf, P. A. Acute cell volume changes in anisotonic media affect F-actin content of HL-60 cells. Am. J. Physiol. 261, C1154–C1161 (1991).

    CAS  PubMed  Google Scholar 

  43. Maruyama, K., Kaibara, M. & Fukada, E. Rheology of F-actin I. Network of F-actin in solution. Biochim. Biophys. Acta Protein Struct. 371, 20–29 (1974).

    CAS  Google Scholar 

  44. Janmey, P. A., Euteneuer, U., Traub, P. & Schliwa, M. Viscoelastic properties of vimentin compared with other filamentous biopolymer networks. J. Cell Biol. 113, 155–160 (1991).

    CAS  PubMed  Google Scholar 

  45. Burg, T. P. et al. Weighing of biomolecules, single cells and single nanoparticles in fluid. Nature 446, 1066–1069 (2007).

    CAS  PubMed  Google Scholar 

  46. Nawaz, A. A. et al. Using real-time fluorescence and deformability cytometry and deep learning to transfer molecular specificity to label-free sorting. Preprint at bioRxiv https://doi.org/10.1101/862227 (2019).

  47. Rosendahl, P. et al. Real-time fluorescence and deformability cytometry. Nat. Methods 15, 355–358 (2018).

    CAS  PubMed  Google Scholar 

  48. Mietke, A. et al. Extracting cell stiffness from real-time deformability cytometry: theory and experiment. Biophys. J. 109, 2023–2036 (2015).

    CAS  PubMed  PubMed Central  Google Scholar 

  49. Mokbel, M. et al. Numerical simulation of real-time deformability cytometry to extract cell mechanical properties. ACS Biomater. Sci. Eng. 3, 2962–2973 (2017).

    CAS  PubMed  Google Scholar 

  50. Fregin, B. et al. High-throughput single-cell rheology in complex samples by dynamic real-time deformability cytometry. Nat. Commun. 10, 415 (2019).

    CAS  PubMed  PubMed Central  Google Scholar 

  51. Guck, J. Some thoughts on the future of cell mechanics. Biophys. Rev. 11, 667–670 (2019).

    PubMed  PubMed Central  Google Scholar 

  52. Lee, J. et al. Suspended microchannel resonators with piezoresistive sensors. Lab Chip 11, 645–651 (2011).

    CAS  PubMed  Google Scholar 

  53. Cermak, N. et al. High-throughput measurement of single-cell growth rates using serial microfluidic mass sensor arrays. Nat. Biotechnol. 34, 1052–1059 (2016).

    CAS  PubMed  PubMed Central  Google Scholar 

  54. Herbig, M. et al. in Flow Cytometry Protocols. Methods in Molecular Biology Vol. 1678 (eds Hawley, T. & Hawley, R.) 347–369 (Humana Press, 2018).

  55. Herold, C. Mapping of deformation to apparent Young’s modulus in real-time deformability cytometry. Preprint at http://arxiv.org/abs/1704.00572 (2017).

  56. Hoffmann, E., Lambert, I. H. & Pedersen, S. F. Physiology of cell volume regulation in vertebrates. Physiol. Rev. 89, 193–277 (2009).

    CAS  PubMed  Google Scholar 

  57. Motulsky, H. & Christopoulos, A. Fitting Models to Biological Data Using Linear and Nonlinear Regression: A Practical Guide to Curve Fitting Ch. F.22, 138–142 (Oxford Univ. Press, 2004).

  58. Abdi, H. in Encyclopedia of Measurement and Statistics (ed. Salkind, N.) 103–107 (Sage, 2007).

  59. Ritz, C., Baty, F., Streibig, J. C. & Gerhard, D. Dose-response analysis using R. PLoS ONE 10, e0146021 (2015).

    PubMed  PubMed Central  Google Scholar 

  60. Urbanska, M. et al. MetaDeformability Dataset (Figshare, 2020); https://doi.org/10.6084/m9.figshare.11704119

Download references

Acknowledgements

We thank P. Janmey for helpful discussions, J. H. Kang and D. Soteriou for useful comments on the manuscript, A. Mietke (Massachusetts Institute of Technology) and C. Herold (Zellmechanik Dresden) for sharing parts of analysis codes, A.L. and D.E. Olins (University of New England) for providing the HL60/S4 cell line and the Microstructure Facility at the Center for Molecular and Cellular Bioengineering at Technische Universität Dresden (in part funded by the State of Saxony and the European Regional Development Fund) for help with the production of sDC chips. We acknowledge funding from the Alexander von Humboldt-Stiftung (Alexander von Humboldt Professorship to J.G.), the Ludwig Center for Molecular Oncology (S.R.M.), the Cancer Systems Biology Consortium U54 CA217377 and the Koch Institute Support Grant P30 CA14051 from the NCI (S.R.M.), the German Federal Ministry of Research and Education (ZIK grant to O.O. under grant agreement 03Z22CN11), and a Presidential Early Career Award for Scientists and Engineers (N00014-16-1-2997, to D.D.).

Author information

Authors and Affiliations

Authors

Contributions

J.G., D.D. and S.R.M. conceptualized the project. M.U., H.E.M. and J.S.B. performed the experiments and analyzed the data. O.O. provided methodological support with sDC data acquisition and analysis. M.U. and H.E.M. visualized the data and prepared the original manuscript draft. All authors revised and edited the manuscript. J.G., D.D., S.R.M. and O.O. acquired funding.

Corresponding authors

Correspondence to Scott R. Manalis, Dino Di Carlo or Jochen Guck.

Ethics declarations

Competing interests

M.U., H.E.M., J.S.B. and J.G. declare no competing interests. O.O. is a shareholder of Zellmechanik Dresden GmbH distributing real-time deformability cytometry; Zellmechanik Dresden GmbH owns a patent for Real-Time Deformability Cytometry (RT-DC): EU patent under the number EP 30 036 520 B1. S.R.M. is a founder of Travera and Affinity Biosensors. D.D. has financial interests in Cytovale Inc., which is commercializing deformability cytometry technology.

Additional information

Peer review information Nina Vogt 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.

Integrated supplementary information

Supplementary Fig. 1 Distribution of passage times measured with cDC for untreated HL60 cells.

Bars represent binned data of control HL60 cells pulled from 10 experiments (total cell number, n = 9,734). Dashed curve represents Kernel density estimation (KDE) of probability distribution with vertical line at the most represented value equal to 23 ms.

Supplementary Fig. 2 Cell diameter of HL60 cells exposed to different osmotic shock conditions.

(ac) Violin plots of cell diameter in a single experiment as measured by cDC (a), sDC (b) and xDC (c). Black boxes extend from 25th to 75th percentiles, with a dot at the median, whiskers indicate 1.5× IQR (interquartile range). (df) Summary of median cell diameter values obtained in all experiment series with cDC (d), sDC (e) and xDC (f). Data points correspond to medians of individual experiments (n = 3, 4 and 4, for cDC, sDC and xDC, respectively). Conditions measured in the same experimental series are color-coded. Boxes span 2× standard deviation with a line at the mean of all medians. In (df) statistical significance of overall differences among mean cell sizes at different osmolarities was tested using analysis of variance (ANOVA) and its result is shown on top of the horizontal line overarching all conditions. The P values reported above each box come from comparison of the given treatment to the control condition (300 mOsm) obtained through post-hoc analysis using two-sided pairwise t-tests for multiple comparison with Benjamin-Hochberg P-value adjustment.

Supplementary Fig. 3 Visualization of bin selection and data processing for osmolarity experiments.

(ac) 50%-density contour plots of deformability vs cell diameter for an exemplary experiment on HL60 cells subjected to different osmolarity conditions. The contour plots are accompanied by deformability and cell diameter histograms for cDC (a), sDC (b), and xDC (c). The most represented 1-μm wide diameter bins used for relative deformability, RD, calculations and and the corresponding deformability histograms are outlined in grey. (df) Jitter plots showing distribution of RD from cDC (d), sDC (e), and xDC (f) measurements for a single experiment. Boxes extend from 25th to 75th percentiles, with a dot at the median, whiskers indicate 1.5× IQR (interquartile range) and each data point corresponds to an individual cell. (gi) Summary of RD values obtained in all experimental series with cDC (g), sDC (h) and xDC (i). Data points correspond to medians of every experiment and conditions measured in same experimental series are color-coded. Boxes span 2× standard deviation with a line at the mean of all medians. (j) Number of events in the selected 1-μm wide diameter bin for each condition and method. (k) Events selected within the 1-μm wide diameter bin as a percentage of all events measured. In j and k, the boxes span 2× standard deviation with a line at the mean. In g-k, the statistics have been calculated for n = 3, 4 and 4 independent measurement replicates, for cDC, sDC and xDC, respectively.

Supplementary Fig. 4 Time-resolved effect of osmotic shock on HL60 deformability and size as measured by sDC.

(ac) The changes in HL60 relative deformability, RD (a), and cell diameter (b) over time after exposure to medium with altered osmolarity as measured by sDC. The experiments were performed in 30 × 30-μm channels at a flowrate of 0.16 μl s−1. Data points represent medians of consecutive measurements taken at different times after the exposure to altered osmolarity medium. On average 3,000 events (and not less than 1,800) were analyzed for each data point.

Supplementary Fig. 5 Fitting of the relation between osmolarity and relative deformability for hyperosmotic shock data.

(a-c) Exponential (red), power law (green) and linear (blue) fits to relative deformability, RD, vs osmolarity data obtained with cDC (a), sDC (b) and xDC (c). Data points in a-c represent means of medians of multiple experimental replicates (n = 3, 4, and 4, for cDC, sDC, and xDC, respectively), error bars represent standard deviation. (d-e) Bar graphs of mean absolute residuals (d) and Bayesian information criterion (BIC) (e), that assess the quality of different fits. Values estimated for each method, as well as mean of values for all methods (n = 3) for given fit function, are presented. The error bars on the mean plots represent standard deviation. Mean absolute residuals give an information on how much the values predicted by the fitted function deviate from the experimental data. Lower values of residuals indicate better agreement of experimental data with proposed function. In case of BIC, lower values indicate a better fit.

Supplementary Fig. 6 Visualization of bin selection and data processing for LatB treatment experiments.

(ac) 50%-density contour plots of deformability vs cell diameter for an exemplary experiment on HL60 cells treated with increasing concentration of LatB. The contour plots are accompanied by deformability and cell diameter histograms for cDC (a), sDC (b), and xDC (c). The most represented 1-μm wide diameter bins used for relative deformability, RD, calculations and the corresponding deformability histograms are outlined in grey. (df), Jitter plots showing distribution of RD from cDC (d), sDC (e), and xDC (f) measurements for a single experiment. Boxes extend from 25th to 75th percentiles, with a dot at the median, whiskers indicate 1.5 × IQR (interquartile range) and each data point corresponds to an individual cell. (gi), Summary of RD values obtained in all experimental series with cDC (g), sDC (h) and xDC (i). Data points correspond to medians of every experiment and conditions measured in the same experimental series are color-coded. Boxes span 2× standard deviation with a line at the mean of all medians. (j) Number of events in the selected 1-μm wide diameter bin for each condition and method. (k) Events selected within the 1-μm wide diameter bin as a percentage of all events measured. In (j) and (k) the boxes span 2× standard deviation with a line at the mean. In g, k, the statistics have been calculated for n = 3, 5 and 4 independent measurement replicates, for cDC, sDC and xDC, respectively.

Supplementary Fig. 7 Dose-response to LatB treatment measured with sDC at three different flow rates.

Deformability, D (a), and relative deformability, RD (b), as a function of LatB concentration at three different flowrates (fr1 = 2.4 µl min-1, fr2 = 4.8 µl min-1, and fr3 = 7.2 µl min-1). The different flowrates are color-coded as indicated in the figure legend (fr1 – gray, fr2 – blue, fr3 – green). Open circles indicate medians of individual measurements, lines connect means of measurement replicates for each flowrate (n = 5), error bars correspond to standard deviation of the mean distributions. Bin-selected data was used.

Supplementary Fig. 8 Response to high LatB concentrations measured with cDC and sDC.

The graph shows relative deformability, RD, as a function of LatB concentration. Dots represent medians of individual measurements. Error bars represent median absolute deviation. Bin-selected data was used. One measurement series was performed using cDC (purple, from left to right n = 296 and 271 analyzed cells in the selected size bin per data point) and two measurement series were performed using sDC (bright and dark green, from left to right n = 656, 537, 420, 550, and n = 734, 615, 336, 541 analyzed cells in the selected size bin, for bright and dark green data points, respectively). The concentration range used for main analysis is shaded in gray.

Supplementary Fig. 9 Cell diameter of HL60 cells treated with different concentrations of LatB.

(ac) Violin plots of cell diameter in a single experiment as measured by cDC (a), sDC (b) and xDC (c). Black boxes extend from 25th to 75th percentiles, with a dot at the median, whiskers indicate 1.5× IQR (interquartile range). (df) Summary of median cell diameter values obtained in all experiment series with cDC (d), sDC (e) and xDC (f). Data points correspond to medians of individual experiments (n = 3, 5 and 4, for cDC, sDC and xDC, respectively). Conditions measured in same experimental series are color-coded. Boxes span 2 × standard deviation with a line at the mean of all medians. In (df) statistical significance of overall differences among mean cell sizes at different concentrations was tested using analysis of variance (ANOVA) and its result is shown on top of the horizontal line overarching all conditions. The P values reported above each box come from comparison of the given treatment to the control condition obtained through post-hoc analysis using pairwise two-sided t-tests for multiple comparisons with Benjamin-Hochberg P-value adjustment.

Supplementary Fig. 10 The influence of size bin selection on relative deformability response to osmotic shock.

For all three methods RD was calculated for either all data, 3-μm wide cell diameter bin or 1-μm wide cell diameter bin. For the ease of comparison, the data is grouped based on binning strategy and all three methods are plotted together (a), or the data is grouped by method and all binning strategies are compared (b). The lines connect the data points representing means of medians from measurement replicates (n = 3, 4, and 4, for cDC, sDC, and xDC, respectively). Error bars present the standard deviation of the medians.

Supplementary Fig. 11 The influence of size bin selection on relative deformability response to LatB treatment.

For all three methods RD was calculated for either all data, 3-μm wide cell diameter bin or 1-μm wide cell diameter bin. For the ease of comparison, the data is grouped based on binning strategy and all three methods are plotted together (a), or the data is grouped by method and all binning strategies are compared (b). The lines connect the data points representing means of medians from measurement replicates (n = 3, 5, and 4, for cDC, sDC, and xDC, respectively). Error bars present the standard deviation of the medians.

Supplementary information

Supplementary Information

Supplementary Figs. 1–11, Note and Tables 1–3.

Reporting Summary

Source data

Source Data Fig. 1

NMETH-AS38212B_SourceData_Figure1.

Source Data Fig. 2

NMETH-AS38212B_SourceData_Figure2.

Source Data Fig. 3

NMETH-AS38212B_SourceData_Figure3.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Urbanska, M., Muñoz, H.E., Shaw Bagnall, J. et al. A comparison of microfluidic methods for high-throughput cell deformability measurements. Nat Methods 17, 587–593 (2020). https://doi.org/10.1038/s41592-020-0818-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41592-020-0818-8

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