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.

  • Technical Report
  • Published:

A reversible shearing DNA probe for visualizing mechanically strong receptors in living cells

Abstract

In the last decade, DNA-based tension sensors have made significant contributions to the study of the importance of mechanical forces in many biological systems. Albeit successful, one shortcoming of these techniques is their inability to reversibly measure receptor forces in a higher regime (that is, >20 pN), which limits our understanding of the molecular details of mechanochemical transduction in living cells. Here, we developed a reversible shearing DNA-based tension probe (RSDTP) for probing molecular piconewton-scale forces between 4 and 60 pN transmitted by cells. Using these probes, we can easily distinguish the differences in force-bearing integrins without perturbing adhesion biology and reveal that a strong force-bearing integrin cluster can serve as a ‘mechanical pivot’ to maintain focal adhesion architecture and facilitate its maturation. The benefits of the RSDTP include a high dynamic range, reversibility and single-molecule sensitivity, all of which will facilitate a better understanding of the molecular mechanisms of mechanobiology.

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: Design and characterization of the RSDTPs.
Fig. 2: Revealing the magnitude and spatial dynamics of integrin forces in living cells with RSDTPs.
Fig. 3: Imaging of different levels of integrin force in real time with multiplexed RSDTPs.
Fig. 4: Mechanically strong integrins maintain the FA architecture.
Fig. 5: Drug treatment shows that the activity of myosin II dominates strong integrin forces.
Fig. 6: Characterization of actin-polymerization-driven integrin forces.

Similar content being viewed by others

Data availability

The raw image data generated and analysed that support the findings of this study are available from the corresponding author upon reasonable request. Source data are provided with this paper.

Code availability

The behaviour of FAs in cells was analysed using the Focal Adhesion Analysis Server (FAAS): a web tool for analysing FA dynamics (https://faas.bme.unc.edu).

References

  1. Iskratsch, T., Wolfenson, H. & Sheetz, M. P. Appreciating force and shape—the rise of mechanotransduction in cell biology. Nat. Rev. Mol. Cell Biol. 15, 825–833 (2014).

    Article  CAS  PubMed  Google Scholar 

  2. Schwartz, M. A. The force is with us. Science 323, 588–589 (2009).

    Article  CAS  PubMed  Google Scholar 

  3. Parsons, J. T., Horwitz, A. R. & Schwartz, M. A. Cell adhesion: integrating cytoskeletal dynamics and cellular tension. Nat. Rev. Mol. Cell Bio. 11, 633–643 (2010).

    Article  CAS  Google Scholar 

  4. Roca-Cusachs, P., Conte, V. & Trepat, X. Quantifying forces in cell biology. Nat. Cell Biol. 19, 742–751 (2017).

    Article  CAS  PubMed  Google Scholar 

  5. Liu, Y., Galior, K., Ma, V. P. Y. & Salaita, K. Molecular tension probes for imaging forces at the cell surface. Acc. Chem. Res. 50, 2915–2924 (2017).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Polacheck, W. J. & Chen, C. S. Measuring cell-generated forces: a guide to the available tools. Nat. Methods 13, 415–423 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Kechagia, J. Z., Ivaska, J. & Roca-Cusachs, P. Integrins as biomechanical sensors of the microenvironment. Nat. Rev. Mol. Cell Bio. 20, 457–473 (2019).

    Article  CAS  Google Scholar 

  8. Kanchanawong, P. et al. Nanoscale architecture of integrin-based cell adhesions. Nature 468, 580–584 (2010).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  9. Hoffman, B. D., Grashoff, C. & Schwartz, M. A. Dynamic molecular processes mediate cellular mechanotransduction. Nature 475, 316–323 (2011).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  10. Plotnikov, S. V., Pasapera, A. M., Sabass, B. & Waterman, C. M. Force fluctuations within focal adhesions mediate ECM-rigidity sensing to guide directed cell migration. Cell 151, 1513–1527 (2012).

    Article  CAS  PubMed  Google Scholar 

  11. Morimatsu, M., Mekhdjian, A. H., Adhikari, A. S. & Dunn, A. R. Molecular tension sensors report forces generated by single integrin molecules in living cells. Nano Lett. 13, 3985–3989 (2013).

    Article  CAS  PubMed  Google Scholar 

  12. Morimatsu, M., Mekhdjian, A. H., Chang, A. C., Tan, S. J. & Dunn, A. R. Visualizing the interior architecture of focal adhesions with high-resolution traction maps. Nano Lett. 15, 2220–2228 (2015).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Ma, V. P.-Y. & Salaita, K. DNA nanotechnology as an emerging tool to study mechanotransduction in living systems. Small 15, 1900961 (2019).

    Article  CAS  Google Scholar 

  14. Stabley, D. R., Jurchenko, C., Marshall, S. S. & Salaita, K. S. Visualizing mechanical tension across membrane receptors with a fluorescent sensor. Nat. Methods 9, 64–67 (2011).

    Article  PubMed  CAS  Google Scholar 

  15. Liu, Y., Yehl, K., Narui, Y. & Salaita, K. Tension sensing nanoparticles for mechano-imaging at the living/nonliving interface. J. Am. Chem. Soc. 135, 5320–5323 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Chang, Y. et al. A general approach for generating fluorescent probes to visualize piconewton forces at the cell surface. J. Am. Chem. Soc. 138, 2901–2904 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Blakely, B. L. et al. A DNA-based molecular probe for optically reporting cellular traction forces. Nat. Methods 11, 1229–1232 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Zhang, Y., Ge, C., Zhu, C. & Salaita, K. DNA-based digital tension probes reveal integrin forces during early cell adhesion. Nat. Commun. 5, 5167 (2014).

    Article  CAS  PubMed  Google Scholar 

  19. Brockman, J. M. et al. Live-cell super-resolved PAINT imaging of piconewton cellular traction forces. Nat. Methods 17, 1018–1024 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Tan, S. J. et al. Regulation and dynamics of force transmission at individual cell-matrix adhesion bonds. Sci. Adv. 6, eaax0317 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  21. Jurchenko, C., Chang, Y., Narui, Y., Zhang, Y. & Salaita, K. S. Integrin-generated forces lead to streptavidin-biotin unbinding in cellular adhesions. Biophys. J. 106, 1436–1446 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Galior, K., Liu, Y., Yehl, K., Vivek, S. & Salaita, K. Titin-based nanoparticle tension sensors map high-magnitude integrin forces within focal adhesions. Nano Lett. 16, 341–348 (2016).

    Article  CAS  PubMed  Google Scholar 

  23. Wang, X. & Ha, T. Defining single molecular forces required to activate integrin and notch signaling. Science 340, 991–994 (2013).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  24. Zhao, Y. C., Wang, Y. L., Sarkar, A. & Wang, X. F. Keratocytes generate high integrin tension at the trailing edge to mediate rear de-adhesion during rapid cell migration. iScience 9, 502–512 (2018).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Zhao, Y. C., Pal, K., Tu, Y. & Wang, X. F. Cellular force nanoscopy with 50 nm resolution based on integrin molecular tension imaging and localization. J. Am. Chem. Soc. 142, 6930–6934 (2020).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  26. Jo, M. H., Cottle, W. T. & Ha, T. Real-time measurement of molecular tension during cell adhesion and migration using multiplexed differential analysis of tension gauge tethers. ACS Biomater. Sci. Eng. 5, 3856–3863 (2019).

    Article  CAS  PubMed  Google Scholar 

  27. Woodside, M. T. et al. Nanomechanical measurements of the sequence-dependent folding landscapes of single nucleic acid hairpins. Proc. Natl Acad. Sci. USA 103, 6190–6195 (2006).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Hatch, K., Danilowicz, C., Coljee, V. & Prentiss, M. Demonstration that the shear force required to separate short double-stranded DNA does not increase significantly with sequence length for sequences longer than 25 base pairs. Phys. Rev. E 78, 011920 (2008).

    Article  CAS  Google Scholar 

  29. Mosayebi, M., Louis, A. A., Doye, J. P. & Ouldridge, T. E. Force-induced rupture of a DNA duplex: from fundamentals to force sensors. ACS Nano 9, 11993–12003 (2015).

    Article  CAS  PubMed  Google Scholar 

  30. Zhang, C. et al. The mechanical properties of RNA–DNA hybrid duplex stretched by magnetic tweezers. Biophys. J. 116, 196–204 (2019).

    Article  CAS  PubMed  Google Scholar 

  31. Liu, Y. et al. DNA-based nanoparticle tension sensors reveal that T-cell receptors transmit defined pN forces to their antigens for enhanced fidelity. Proc. Natl Acad. Sci. USA 113, 5610–5615 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  32. Ma, R. et al. DNA probes that store mechanical information reveal transient piconewton forces applied by T cells. Proc. Natl Acad. Sci. USA 116, 16949–16954 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  33. Berginski, M. E. & Gomez, S. M. The Focal Adhesion Analysis Server: a web tool for analyzing focal adhesion dynamics. F1000Res 2, 68 (2013).

    Article  PubMed  PubMed Central  Google Scholar 

  34. Glazier, R. et al. DNA mechanotechnology reveals that integrin receptors apply pN forces in podosomes on fluid substrates. Nat. Commun. 10, 4507 (2019).

    Article  PubMed  PubMed Central  CAS  Google Scholar 

  35. Liu, Y. et al. Nanoparticle tension probes patterned at the nanoscale: impact of integrin clustering on force transmission. Nano Lett. 14, 5539–5546 (2014).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  36. Swaminathan, V., Fischer, R. S. & Waterman, C. M. The FAK-Arp2/3 interaction promotes leading edge advance and haptosensing by coupling nascent adhesions to lamellipodia actin. Mol. Biol. Cell 27, 1085–1100 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Harris, A. R., Jreij, P. & Fletcher, D. A. Mechanotransduction by the actin cytoskeleton: converting mechanical stimuli into biochemical signals. Annu. Rev. Biophys. 47, 617–631 (2018).

    Article  CAS  Google Scholar 

  38. Gordon, M. P., Ha, T. & Selvin, P. R. Single-molecule high-resolution imaging with photobleaching. Proc. Natl Acad. Sci. USA 101, 6462–6465 (2004).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  39. Basu, R. et al. Cytotoxic T cells use mechanical force to potentiate target cell killing. Cell 165, 100–110 (2016).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  40. Wang, J. et al. Profiling the origin, dynamics, and function of traction force in B cell activation. Sci. Signal. 11, eaai9192 (2018).

    Article  PubMed  CAS  Google Scholar 

  41. Chen, Y. et al. An integrin αIIbβ3 intermediate affinity state mediates biomechanical platelet aggregation. Nat. Mater. 18, 760–769 (2019).

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  42. Zhang, Y. et al. Platelet integrins exhibit anisotropic mechanosensing and harness piconewton forces to mediate platelet aggregation. Proc. Natl Acad. Sci. USA 115, 325–330 (2018).

    Article  CAS  PubMed  Google Scholar 

  43. Myers, D. R. et al. Single-platelet nanomechanics measured by high-throughput cytometry. Nat. Mater. 16, 230–235 (2017).

    Article  CAS  PubMed  Google Scholar 

  44. Li, H., Hu, Y., Sun, F., Chen, W. & Liu, Z. A reversible shearing DNA-based tension probe for cellular force measurement. Protocol Exchange https://doi.org/10.21203/rs.3.pex-1451/v1 (2021).

  45. Piella, J., Bastús, N. G. & Puntes, V. Size-controlled synthesis of sub-10-nanometer citrate-stabilized gold nanoparticles and related optical properties. Chem. Mater. 28, 1066–1075 (2016).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

This work was supported by the National Natural Science Foundation of China (21775115, 32071305, 31670760, 11704286 and 11674403), the start-up funding from Wuhan University for financial support, and the Fundamental Research Funds for the Central Universities (2042018kf02). We thank V. Ma for comments on the manuscript and for helpful discussion.

Author information

Authors and Affiliations

Authors

Contributions

Z.L. and H.L. conceived and initiated the project. H.L. designed and performed all the experiments with the help of Y.H., F.S., P.L., W.C., J.M., W.W., L.W. and P.W. C.Z and X.Z. performed the SMMT experiments. Z.L. supervised the project and wrote the manuscript.

Corresponding author

Correspondence to Zheng Liu.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

Peer review information Nature Cell Biology thanks Khalid Salaita and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

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

Extended data

Extended Data Fig. 1 Synthesis scheme and characterizations of RSDTPs.

a, The structures and oligonucleotide sequences of 17-pN, 45-pN and 56-pN RSDTPs. b, Detailed chemical schemes for the synthesis of RSDTPs. c, Representative HPLC traces of DNA strand I’ modified with a Cy3B dye (denoted by dashed rectangle frame.) (d) Analysis of assay products by 10% denaturing polyacrylamide gel electrophoresis, lane 1: strand I’-Cy3B; lane 2: strand I’-Cy3B-c(RGDfK); lane 3: extra-long single-stranded DNA (elssDNA, highlighted by a black arrow). Note that, the weaker band above is a complex formed by a small amount of elssDNA and the template strand. The image is the representative of at least three independent biological replicates. e, Representative electrospray ionization mass spectrums of elssDNA.

Source data

Extended Data Fig. 2 Schematic diagrams of the composition of the hairpin tension probes for calibration and experimental method of single molecule magnetic tweezer experiments.

a, The geometries and compositions of the force probes used for calibration. All probes, combined with a complementary strand (blue) at their 3′end, were ligated to the biotin modified dsDNA handle. For the 56-pN probe, the sequence highlighted by a rectangular is only used to extend the entire length of probe without significantly affecting the mechanical properties. b, Schematic depiction of the single-molecule magnetic tweezer experiments. The probes are immobilized to the streptavidin-coated glass surface through their DNA handles. To unfold the probe, pulling forces were applied to a streptavidin-coated magnetic bead by permanent magnets. c, Mechanical stability measurements for 56-pN reversible shearing DNA probe. The results show that the 56-pN probe does not unfold even if probe was exposed to a lower level of forces (30–44 pN) for more than 10 minutes, and the unfolding event is only observed when applied force increases above 50 pN and lasted for more than ten seconds.

Extended Data Fig. 3 Characterization of AuNPs and RSDTPs functionalized surfaces preparation.

a-b, Representative TEM images of AuNPs (a) and corresponding size distribution profiles (b). c, UV-VIS spectra of 5 nm AuNPs. d, Stepwise procedures for preparing RSDTPs functionalized surfaces (see online methods). e, The representative AFM image showing the spatial distribution of the AuNPs immobilized on a glass coverslip.

Source data

Extended Data Fig. 4 Determination of DNA probe density on Au nanoparticles surface.

To calibrate the surface density of the DNA tension sensor, we used the method described by Wang and Ha et al.23 to plot the relationship between the fluorescence grayscale units of the sample (mean intensity) and the DNA probe density. First, we incubated aqueous solutions containing 25 nM, 12.5 nM, and 6.25 nM Cy3B-unfolded DNA probe (a scrambled DNA hairpin probe annealing with a complementary strand, the structure was shown in a) on the AuNPs-modified coverslips for 30 minutes, then washed off the unbinding probes and measured the average fluorescence intensity of glass as a function of the incubation concertation of the Cy3B-unfolded DNA probe (b). There is a linear relationship between the mean fluorescence intensity of the surface and the DNA incubation concentration within this concentration range (c). By extrapolation, incubation of 0.05 nM Cy3B-unfolded DNA probe on the surface should correspond to mean fluorescence intensity of 14.4. we also found a molecular density of 0.149 molecules/μm2 on the surface (0.05 nM) using a single-molecule TIRF image (b). Taken together, 96.6 units in mean fluorescence intensity on the Cy3B-unfolded DNA probe surface is equivalent to 1 molecule/ μm2. Given that the average QE of the Cy3B-folded DNA probe is 98.3% (Fig. 1 in main text), the 1.6 units mean fluorescence intensity on the RSDTP-Cy3B coated surface should be equivalent to 1 molecule/ μm2. We could also plot a calibration curve between intensity and molecular density for RSDTP (Cy3B) coated surface, as shown in d. This calibration curve allows us to determine RSDTP (Cy3B) surface density by measuring the average intensity of a coverslip. For example, the 45-pN RSDTP coated surface’s molecular density was estimated to be 671 molecules/μm2 (e).

Source data

Extended Data Fig. 5 Determination of fluorescence quenching efficiency (QE) of RSDTPs on a coverslip.

a, Estimate the distance between the AuNP surface and fluorophores on mechanically unfolding RSDTPs with different rupture forces by assuming a contour length of 0.44 nm per nt. b, Plots of the quenching efficiency as a function of distance from fluorophore to AuNPs with diameters of 5 nm and 10 nm based on the NSET model (Supplementary Note 2). The quenching efficiency (QE)-distance simulation results suggested that the effect of 5 nm AuNP on the Cy3B fluorophore beyond a distance of 20 nm is negligible. c, Schematic diagram of measuring the fluorescence quenching efficiency of RSDTPs on a solid surface. To mimic the folded and unfolded states of different RSDTPs, we hybridize complementary DNA strands (preheated to 95 °C) with different length to a scrambled hairpin probe (sequences shown in e). The QE values are calculated by measuring the fluorescence intensities before and after the addition of complementary strands. d, An example of QE measurement. The QE value of 17-pN RSDTP was calculated as 97.3% by measuring the readout noise of EMCCD, surface fluorescence intensities before and after addition of complementary strands. e, Table listing the sequences of the DNA used in this assay.

Extended Data Fig. 6 Control experiments confirmed the reliability and reversibility of the tension probes.

a, Comparison of DIC images for cells plated on sensors with and without RGD peptide. No cell attachment was found on surface coated with sensors lacking RGD peptide, suggesting that tension signal is specially generated through RGD-integrin interactions. b, Representative DIC images of NIH 3T3 cells seeded on tension probes and fibronectin-coated coverslips at different time points. c, A 45-pN RSDTP that lacks the quencher oligonucleotide was used to further validate the stability and reversibility of tension probes. The images show that the fluorescent background of the surface that lacks the quencher oligonucleotide is three times higher than a surface with the quencher strand DNA. Despite the high fluorescence background of these surfaces, we can still detect a tension signal during cell spreading on these surfaces, because Au NP plays a second quencher here. After treating cells with Cyto D, we found that the tension signals disappeared immediately and didn’t leave any dark features, which are often used to determine the stability of tension probes on the surface18. The fluorescence profiles along the yellow lines in the images are also shown in the right panel.

Extended Data Fig. 7 Further confirmation of the force gradient within FA and the relationship between FA longevity and mechanically hotspots.

a, Representative 17-pN (Atto647N), 56-pN (Cy3B) and the overlay tension images for a single NIH/3T3 cells chosen from 2 independent replicates spreading on a coverslip encoded with multiplexed RSDTPs. b, Representative 56-pN tension image along with GFP-paxillin of a 3T3 cell chosen from 4 independent replicates. Contour of GFP-paxillin was overlaid with tension signal. c, Zoomed-in time-lapse images from yellow-boxed region in b showing adhesion structures bearing weak forces undergo a rapid disassembly process. d, Longevity of weak and strong adhesions. The box represents 25th and 75th percentiles, the median is denoted by the middle horizontal line with mean value is labeled above each box and whiskers represent 1.5-fold the interquartile range. n = 49 for weak adhesions (structures without 56-pN signal) and n = 34 for strong adhesions (structures contain 56-pN signal) in the cell from b. Two-tailed, Student’s t-test was used to measure statistical significance, and P-values are shown on the graph.

Source data

Extended Data Fig. 8 Testing the reliability of photocleavable RSDTPs.

a, Representative RICM and tension images before and after UV illumination (time=30 s, UV-light is operating with a periodic illumination, power density = 1.3 × 10-4 μW/μm2) for cells seeding on different photocleavable RSDTPs coated surfaces. After cleavage of the loop of probes, the cell exhibited a rapid contraction on 17-pN photocleavable RSDTP and slow contraction on 56-pN photocleavable RSDTP, suggested that the RSDTP have been switched to a TGT-like probe. The images were representative of at least 2 independent replicates. b, FA dynamics of GFP-Paxillin expressing 3T3 cells seeded on 56-pN RSDTP, 56-pN TGT and 56-pN photocleavable RSDTP coated surface in response to UV illumination. Same illumination condition as in a were used here. No significant effect on cells seeded on 56-pN RSDTPs and 56-pN TGT probe coated surfaces were observed. The images were representative of at least 3 independent replicates.

Extended Data Fig. 9 Images of myosin-independent integrin force.

a, Tension images of blebbistatin pre-treated cells adhering to RSDTPs with different rupture forces. NIH 3T3 cells were pretreated with 10 μM blebbistatin for 30 minutes and then seeded on the coverslips. The images were representative of 3 independent replicates. b, Representative 45-pN tension images of a Y-27632 pretreated NIH 3T3 cell chosen from 2 independent replicates before and after adding Cytochalasin D. c, Co-localization of the integrin tension signal with GFP-Paxillin and GFP-Actin in 3T3 cells treated with Y27632 on 45-pN RSDTP coated surface. Images of RICM (gray), GFP (green), tension (red) and overlay channels were shown. Line scan analysis plot of the region highlighted with dashed white arrows in the overlay images shows the co-localization of the tension with GFP-paxillin and GFP-actin. The images were representative of 3 independent replicates. d, Representative 33-pN tension images of a Y-27632 pretreated NIH 3T3 cell chosen from 3 independent biological replicates before and after adding 100 mM sucrose. The right panel shows the change of the thin edge tension signal before and after hypertonic treatment, n=9 cells.

Source data

Extended Data Fig. 10 Single-molecule subunit counting of the force-bearing integrins at thin edges of Y27632 pretreated NIH 3T3 cells.

a, Left, a representative tension image (45-pN RSDTP) of a single living cell treated with Y27632 showing punctate fluorescent points. Right, representative intensity traces, under continuous excitation sufficient to cause bleaching of the fluorophores, from four clusters denoted by yellow arrows in the tension image. Below, histogram and gaussian fit of fluorescence intensities for the bleaching steps in the traces (n = 39 steps from cell 1). b, Additional data were performed on two fixed cells, and similar results were obtained. (n = 38 steps for cell 2; n = 36 steps for cell 3).

Source data

Supplementary information

Supplementary Information

Supplementary Fig. 1 and Supplementary Notes 1 and 2.

Reporting Summary

Supplementary Video 1

Time-lapse video showing RICM and tension imaging of NIH-3T3 cells spreading on the coverslips functionalized by RSDTPs with different rupture forces (top, 17 pN; middle, 45 pN; bottom, 56 pN). The left panel represents the RICM channel, the middle panel represents the integrin tension reported as fluorescent signals by the RSDTPs, and the right channel represents the overlay of RICM and tension channels. Images were acquired for a duration of 80 min at 5-min intervals.

Supplementary Video 2

Time-lapse video showing the dynamics of FAs (marked by GFP–paxillin) in NIH-3T3 cells spreading on 56-pN TGTs, 56-pN RSDTPs, mixed-probe surface (2% 56-pN RSDTPs and 98% 56-pN TGTs) or fibronectin-functionalized surfaces. Images were acquired for a duration of 60 min at 1-min intervals.

Supplementary Video 3

Time-lapse video showing the changes in FA dynamics and 56-pN tension signals before and after UV illumination when a NIH-3T3 cell expressing GFP–paxillin spread on a mixed-probe surface (2% 56-pN photocleavable RSDTPs and 98% 56-pN non-fluorescent TGTs). Images were acquired every 2 min, UV illumination condition: 365 nm, 0.25 Hz frequency, 1.3 × 10–4 μW μm–2.

Supplementary Video 4

Time-lapse video showing the dynamics of the 17-pN (Cy3B) and 56-pN (Atto647N) tension signals along with the cytoskeleton during the recovery of myosin activity. NIH-3T3 cells were pretreated with 20 μM Y-27632 before seeding on a multiplexed tension probe surface. Y-27632 was removed by washing with culture medium without drugs twice, and video was recorded at t = 0 when the fresh medium was added. Images were acquired for a duration of 80 min at 5-min intervals.

Supplementary Video 5

Time-lapse video showing GFP–actin (inverted contrast) and tension map (red) of a Y-27632-pretreated NIH-3T3 cell spreading on a 45-pN RSDTP functionalized surface.

Supplementary Tables

Supplementary Tables 1 and 2.

Source data

Source Data Fig. 1

Statistical source data.

Source Data Fig. 2

Statistical source data.

Source Data Fig. 3

Statistical source data.

Source Data Fig. 4

Statistical source data.

Source Data Fig. 4

Unprocessed gel.

Source Data Fig. 6

Statistical source data.

Source Data Extended Data Fig. 1

Unprocessed gel.

Source Data Extended Data Fig. 3

Statistical source data.

Source Data Extended Data Fig. 4

Statistical source data.

Source Data Extended Data Fig. 7

Statistical source data.

Source Data Extended Data Fig. 9

Statistical source data.

Source Data Extended Data Fig. 10

Statistical source data.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Li, H., Zhang, C., Hu, Y. et al. A reversible shearing DNA probe for visualizing mechanically strong receptors in living cells. Nat Cell Biol 23, 642–651 (2021). https://doi.org/10.1038/s41556-021-00691-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41556-021-00691-0

This article is cited by

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

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