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:

Acoustically triggered mechanotherapy using genetically encoded gas vesicles

Abstract

Recent advances in molecular engineering and synthetic biology provide biomolecular and cell-based therapies with a high degree of molecular specificity, but limited spatiotemporal control. Here we show that biomolecules and cells can be engineered to deliver potent mechanical effects at specific locations inside the body through ultrasound-induced inertial cavitation. This capability is enabled by gas vesicles, a unique class of genetically encodable air-filled protein nanostructures. We show that low-frequency ultrasound can convert these biomolecules into micrometre-scale cavitating bubbles, unleashing strong local mechanical effects. This enables engineered gas vesicles to serve as remotely actuated cell-killing and tissue-disrupting agents, and allows genetically engineered cells to lyse, release molecular payloads and produce local mechanical damage on command. We demonstrate the capabilities of biomolecular inertial cavitation in vitro, in cellulo and in vivo, including in a mouse model of tumour-homing probiotic therapy.

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: Purified GVs act as seeds for stable and inertial cavitation.
Fig. 2: Ultrafast optical imaging and acoustic recording of GV-seeded bubble formation and cavitation.
Fig. 3: Molecularly targeted GVs serve as ultrasound-triggered disruptors of mammalian cells.
Fig. 4: GVs as genetically encoded seeds for cellular inertial cavitation and payload release.
Fig. 5: GV-seeded cavitation and tissue disruption in vivo.
Fig. 6: Tumour mechanotherapy seeded by tumour-homing probiotics.

Similar content being viewed by others

Data availability

Plasmids sequences are given in Supplementary Table 3, and will be made available through Addgene. All the raw data related to the plots and graphs are available at https://github.com/shapiro-lab/GV_cavitation.git. All the other materials and data are available from the corresponding author upon reasonable request.

Code availability

MATLAB codes are available from the corresponding author upon reasonable request.

References

  1. Milenic, D. E., Brady, E. D. & Brechbiel, M. W. Antibody-targeted radiation cancer therapy. Nat. Rev. Drug Discov. 3, 488–498 (2004).

    Article  CAS  Google Scholar 

  2. Din, M. O. et al. Synchronized cycles of bacterial lysis for in vivo delivery. Nature 536, 81–85 (2016).

    Article  CAS  Google Scholar 

  3. Danino, T. et al. Programmable probiotics for non-invasive urinary detection of cancer. Sci. Transl. Med. 7, 289ra84 (2015).

    Article  Google Scholar 

  4. Claesen, J. & Fischbach, M. A. Synthetic microbes as drug delivery systems. ACS Synth. Biol. 4, 358–364 (2015).

    Article  CAS  Google Scholar 

  5. Daniel, C., Roussel, Y., Kleerebezem, M. & Pot, B. Recombinant lactic acid bacteria as mucosal biotherapeutic agents. Trends Biotechnol. 29, 499–508 (2011).

    Article  CAS  Google Scholar 

  6. Steidler, L. et al. Treatment of murine colitis by Lactococcus lactus secreting interleukin-10. Adv. Sci. 289, 1352–1355 (2011).

    Google Scholar 

  7. Davili, M. L. et al. Efficacy and toxicity management of 19-28z CAR T cell therapy in B cell acute lymphoblastic leukemia. Sci. Transl. Med. 6, 224ra25 (2014).

    Google Scholar 

  8. Jackson, H. J., Rafiq, S. & Brentjens, R. J. Driving CAR T-cells forward. Nat. Rev. Clin. Oncol. 13, 370–383 (2016).

    Article  CAS  Google Scholar 

  9. Scheller, L. & Fussenegger, M. From synthetic biology to human therapy: engineered mammalian cells. Curr. Opin. Biotechnol. 58, 108–116 (2019).

    Article  CAS  Google Scholar 

  10. Riglar, D. T. & Silver, P. A. Engineering bacteria for diagnostic and therapeutic applications. Nat. Rev. Microbiol. 16, 214–225 (2018).

    Article  CAS  Google Scholar 

  11. Walsby, A. E. Gas vesicles. Annu. Rev. Plant Physiol. 26, 427–439 (1975).

    Article  CAS  Google Scholar 

  12. Pfeifer, F. Distribution, formation and regulation of gas vesicles. Nat. Rev. Microbiol. 10, 705–715 (2012).

    Article  CAS  Google Scholar 

  13. Dutka, P. et al. Measuring gas vesicle dimensions by electron microscopy. Protein Sci. 30, 1081–1086 (2021).

    Article  CAS  Google Scholar 

  14. Shapiro, M. G. et al. Biogenic gas nanostructures as ultrasonic molecular reporters. Nat. Nanotechnol. 9, 311–316 (2014).

    Article  CAS  Google Scholar 

  15. Lu, G. J. et al. Imaging of gas-filled protein nanostructures. Nat. Mater. 17, 456–463 (2018).

    Article  CAS  Google Scholar 

  16. Shapiro, M. G. et al. Genetically encoded reporters for hyperpolarized xenon magnetic resonance imaging. Nat. Chem. 6, 629–634 (2014).

    Article  CAS  Google Scholar 

  17. Lakshmanan, A. et al. Molecular engineering of acoustic protein nanostructures. ACS Nano 10, 7314–7322 (2016).

    Article  CAS  Google Scholar 

  18. Kunth, M., Lu, G. J., Witte, C., Shapiro, M. G. & Schro, L. Protein nanostructures produce self-adjusting hyperpolarized magnetic resonance imaging contrast through physical gas partitioning. ACS Nano 11, 10939–10948 (2018).

    Article  Google Scholar 

  19. Maresca, D., Sawyer, D. P., Renaud, G., Lee-Gosselin, A. & Shapiro, M. G. Nonlinear X-wave ultrasound imaging of acoustic biomolecules. Phys. Rev. X 8, 41002 (2018).

    CAS  Google Scholar 

  20. Maresca, D. et al. Nonlinear ultrasound imaging of nanoscale acoustic biomolecules. Appl. Phys. Lett. 110, 0073704 (2017).

    Article  Google Scholar 

  21. Bourdeau, R. W. et al. Acoustic reporter genes for noninvasive imaging of microorganisms in mammalian hosts. Nature 553, 86–90 (2018).

    Article  CAS  Google Scholar 

  22. Farhadi, A., Ho, G. H., Sawyer, D. P., Bourdeau, R. W. & Shapiro, M. G. Ultrasound imaging of gene expression in mammalian cells. Science 365, 1469–1475 (2019).

    Article  CAS  Google Scholar 

  23. Airan, R. D. et al. Noninvasive targeted transcranial neuromodulation via focused ultrasound gated drug release from nanoemulsions. Nano Lett. 17, 652–659 (2017).

    Article  CAS  Google Scholar 

  24. Kwan, J. J. et al. Ultrasound-propelled nanocups for drug delivery. Small 11, 5305–5314 (2015).

    Article  CAS  Google Scholar 

  25. Kwan, J. J. et al. Ultrasound-induced inertial cavitation from gas-stabilizing nanoparticles. Phys. Rev. E 92, 023019 (2015).

    Article  CAS  Google Scholar 

  26. Ferrara, K., Pollard, R. & Borden, M. Ultrasound microbubble contrast agents: fundamentals and application to gene and drug delivery. Annu. Rev. Biomed. Eng. 9, 415–447 (2007).

    Article  CAS  Google Scholar 

  27. Kooiman, K., Vos, H. J., Versluis, M. & De Jong, N. Acoustic behavior of microbubbles and implications for drug delivery. Adv. Drug Deliv. Rev. 72, 28–48 (2014).

    Article  CAS  Google Scholar 

  28. Kotopoulis, S., Dimcevski, G., Gilja, O. H., Hoem, D. & Postema, M. Treatment of human pancreatic cancer using combined ultrasound, microbubbles, and gemcitabine: a clinical case study. Med. Phys. 40, 072902 (2013).

    Article  Google Scholar 

  29. Sheeran, P. S. & Dayton, P. A. Phase-change contrast agents for imaging and therapy. Curr. Pharm. Des. 18, 2152–2165 (2012).

    Article  CAS  Google Scholar 

  30. Kripfgans, O. D., Fowlkes, J. B., Miller, D. L., Eldevik, O. P. & Carson, P. L. Acoustic droplet vaporization for therapeutic and diagnostic applications. Ultrasound Med. Biol. Biol. 26, 1177–1189 (2000).

    Article  CAS  Google Scholar 

  31. Köse, G., Darguzyte, M. & Kiessling, F. Molecular ultrasound imaging. Nanomaterials 10, 1935 (2020).

    Article  Google Scholar 

  32. Woudstra, L. et al. Development of a new therapeutic technique to direct stem cells to the infarcted heart using targeted microbubbles: StemBells. Stem Cell Res. 17, 6–15 (2016).

    Article  CAS  Google Scholar 

  33. Ashokkumar, M., Lee, J., Kentish, S. & Grieser, F. Bubbles in an acoustic field: an overview. Ultrason. Sonochem. 14, 470–475 (2007).

    Article  CAS  Google Scholar 

  34. Coussios, C. C. & Roy, R. A. Applications of acoustics and cavitation to noninvasive therapy and drug delivery. Annu. Rev Fluid Mech. 40, 395–420 (2008).

    Article  Google Scholar 

  35. Szabo, T. L. Diagnostic Ultrasound Imaging: Inside Out (Academic, 2004).

  36. Tran, B. C., Seo, J., Hall, T. L., Fowlkes, J. B. & Cain, C. A. Microbubble-enhanced cavitation for noninvasive ultrasound surgery. IEEE Trans. Ultrason. Ferroelectr. Freq. Control 50, 1296–1304 (2003).

    Article  Google Scholar 

  37. Church, C. C. Frequency, pulse length, and the mechanical index. Acoust. Res. Lett. 6, 162 (2007).

    Article  Google Scholar 

  38. Holland, C. K. & Apfel, R. E. An improved theory for the prediction of microcavitation thresholds. IEEE Trans. Ultrason. Ferroelectr. Freq. Control 36, 204–208 (1989).

    Article  CAS  Google Scholar 

  39. Cherin, E. et al. Acoustic behavior of Halobacterium salinarum gas vesicles in the high-frequency range: experiments and modeling. Ultrasound Med. Biol. 43, 1016–1030 (2017).

    Article  Google Scholar 

  40. Walsby, A. E. The pressure relationships of gas vacuoles. Proc. R. Soc. Lond. B 178, 301–326 (1971).

    Article  Google Scholar 

  41. Lakshmanan, A. et al. Preparation of biogenic gas vesicle nanostructures for use as contrast agents for ultrasound and MRI. Nat. Protocols 12, 2050–2080 (2017).

    Article  CAS  Google Scholar 

  42. Le Floc’h, J. et al. In vivo biodistribution of radiolabeled acoustic protein nanostructures. Mol. Imaging Biol. 20, 230–239 (2018).

    Article  Google Scholar 

  43. Duong, M. T. Q., Qin, Y., You, S. H. & Min, J. J. Bacteria-cancer interactions: bacteria-based cancer therapy. Exp. Mol. Med. 51, 1–15 (2019).

    Article  CAS  Google Scholar 

  44. Dang, L. H., Bettegowda, C., Huso, D. L., Kinzler, K. W. & Vogelstein, B. Combination bacteriolytic therapy for the treatment of experimental tumors. Proc. Natl Acad. Sci. USA 98, 15155–15160 (2001).

    Article  CAS  Google Scholar 

  45. Leschner, S. et al. Tumor invasion of Salmonella enterica serovar Typhimurium is accompanied by strong hemorrhage promoted by TNF-α. PLoS ONE 4, e6692 (2009).

    Article  Google Scholar 

  46. Kang, S.-R. et al. Imaging of tumor colonization by Escherichia coli using 18F-FDS PET. Theranostics 10, 4958–4966 (2020).

    Article  CAS  Google Scholar 

  47. Gurbatri, C. R. et al. Engineered probiotics for local tumor delivery of checkpoint blockade nanobodies. Sci. Transl. Med. 12, eaax0876 (2020).

    Article  CAS  Google Scholar 

  48. Jiang, S.-N. et al. Inhibition of tumor growth and metastasis by a combination of Escherichia coli-mediated cytolytic therapy and radiotherapy. Mol. Ther. 18, 635–642 (2010).

    Article  CAS  Google Scholar 

  49. Ryan, R. M. et al. Bacterial delivery of a novel cytolysin to hypoxic areas of solid tumors. Gene Ther. 16, 329–339 (2009).

    Article  CAS  Google Scholar 

  50. Groot, A. J. et al. Functional antibodies produced by oncolytic clostridia. Biochem. Biophys. Res. Commun. 364, 985–989 (2007).

    Article  CAS  Google Scholar 

  51. Leventhal, D. S. et al. Immunotherapy with engineered bacteria by targeting the STING pathway for anti-tumor immunity. Nat. Commun. 11, 2739 (2020).

    Article  CAS  Google Scholar 

  52. Hurt, R. C. et al. Genomically mined acoustic reporter genes enable on-demand in vivo monitoring of tumor-homing bacteria. Preprint at bioRxiv https://doi.org/10.1101/2021.04.26.441537 (2021).

  53. Errico, C. et al. Ultrafast ultrasound localization microscopy for deep super-resolution vascular imaging. Nature 527, 499–502 (2015).

    Article  CAS  Google Scholar 

  54. Eranki, A. et al. High-intensity focused ultrasound (hIFU) triggers immune sensitization of refractory murine neuroblastoma to checkpoint inhibitor therapy. Clin. Cancer Res. 26, 1152–1161 (2020).

    Article  CAS  Google Scholar 

  55. Qu, S. et al. Non-thermal histotripsy tumor ablation promotes abscopal immune responses that enhance cancer immunotherapy. J. Immunother. Cancer 8, 1–12 (2020).

    Article  Google Scholar 

  56. Chavez, M. et al. Distinct immune signatures in directly treated and distant tumors result from TLR adjuvants and focal ablation. Theranostics 8, 3611–3628 (2018).

    Article  CAS  Google Scholar 

  57. Häcker, G., Redecke, V. & Häcker, H. Activation of the immune system by bacterial CpG-DNA. Immunology 105, 245–251 (2002).

    Article  Google Scholar 

  58. Forbes, N. S., Munn, L. L., Fukumura, D. & Jain, R. K. Sparse initial entrapment of systemically injected Salmonella typhimurium leads to heterogeneous accumulation within tumors. Cancer Res. 63, 5188–5193 (2003).

    CAS  Google Scholar 

  59. Ling, B. et al. Biomolecular ultrasound imaging of phagolysosomal function. ACS Nano 14, 12210–12221 (2020).

    Article  CAS  Google Scholar 

  60. Jang, M. J. & Nam, Y. NeuroCa: integrated framework for systematic analysis of spatiotemporal neuronal activity patterns from large-scale optical recording data. Neurophotonics 2, 035003 (2015).

    Article  Google Scholar 

  61. 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 

  62. Bar-Zion, A., Yin, M., Adam, D. & Foster, F. S. Functional flow patterns and static blood pooling in tumors revealed by combined contrast-enhanced ultrasound and photoacoustic imaging. Cancer Res. 76, 4320–4331 (2016).

    Article  CAS  Google Scholar 

Download references

Acknowledgements

The authors thank D. Piraner, A. Lakshmanan, A. Farhadi and P. Ramesh for helpful discussions. In addition, we thank A. Farhadi for his help with the GvpC-RGD variant and H. Davis for his inputs on the optical design of the high-speed set-up. We thank M. Harel (www.maayanillustration.com) for the illustrations in this paper. We also thank A. McDowall for help with electron microscopy and C. Rabut for help with the animal experiments. This project was supported by the David and Lucile Packard Fellowship for Science and Engineering (M.G.S.) and the Heritage Medical Research Institute (M.G.S.). In addition, this project received funding from the European Union’s Horizon 2020 research and innovation programme under Marie Skłodowska-Curie grant agreement no. 792866 (A.B.-Z.). A.B.-Z. was also supported by the Lester Deutsch Fellowship. A.N. was supported by the Amgen scholars programme. S.S. is supported by the NSF Graduate Research Fellowship. M.H.A. is supported by the NSF Graduate Research Fellowship and the P.D. Soros Fellowship. M.T.B. is supported by the NSF Graduate Research Fellowship. D. Maresca is supported by the Human Frontiers Science Program Cross-Disciplinary Fellowship.

Author information

Authors and Affiliations

Authors

Contributions

A.B.-Z. and M.G.S. conceived the study. A.B.-Z., A.N., D. Maresca, D.R.M., S.Y. and S.S. designed, planned and conducted the in vitro experiments. A.B.-Z., A.N., M.T.B., R.C.H., A.L.-G. and M.B.S. designed, planned and conducted in vivo experiments. A.B.-Z. edited the gene circuits with the guidance of M.H.A. A.B.-Z., A.N., D.R.M., S.Y. and D. Maresca analysed the data. D. Malounda prepared the purified GVs. All the authors discussed the results. A.B.-Z., A.N. and M.G.S wrote the manuscript with input from all the authors. All the authors have given their approval for the final version of the manuscript. M.G.S. supervised the research.

Corresponding author

Correspondence to Mikhail G. Shapiro.

Ethics declarations

Competing interests

The California Institute of Technology has filed a patent application related to this manuscript. The authors have no other competing interests.

Additional information

Peer review information Nature Nanotechnology thanks Mark Borden 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 GVs attenuate ultrasound at high concentrations.

(a) Illustration of the sample chamber and setup as seen in the images. (b) B-mode images of purified Ana GVs in different concentrations showing acoustic shadowing at high concentrations. Scale bar, 3 mm. (c) Average broadband emissions measured as a function of GV concentration. GVs were insonated with a single 30-cycle pulse with PNP = 1.0 MPa (n = 5).

Extended Data Fig. 2 GV-seeded cavitation at 3 MHz requires higher pressure levels.

(a) Broadband signals recorded from GVs (0.3 nM) and BSA (matched in mg/mL to GVs concentration) insonated at 3 MHz. Broadband signal increased with pressure and was significantly higher for GV samples at PNP≥0.5 MPa (p < 0.05 for PNP < 1.33 MPa, and p < 0.001 at higher pressure levels, n = 8 independent samples). (b) Comparison between broadband signals from GVs insonated with 670 kHz and 3 MHz pulses (n = 16 and n = 8 independent samples, respectively). Error bars, mean ± s.e.m (a-b).

Extended Data Fig. 3 High frame rate optical imaging of GV collapse and bubble cavitation.

High-speed camera frames (left to right then top to bottom) of GV collapse and cavitation (200 ns between each frame, 31×31 μm field of view), focusing on a single bubble. Initial black spots, which correspond to intact GVs, first disappear due to collapse, liberating gas that coalesces into a cavitating bubble in this region. GVs were insonated with a single 30-cycle pulse with PNP of 1.4 MPa and a central frequency of 670 kHz.

Extended Data Fig. 4 Simulation of free bubble dissolution.

The kinetics of gas bubble dissolution were calculated based on the modified-EP (Epstein and Plesset) equation, following the analysis in [25]: \(- \frac{{{{{\mathbf{dr}}}}}}{{{{{\mathbf{dt}}}}}} = \frac{{{{\mathbf{L}}}}}{{{{{\mathbf{r}}}}/{{{\mathbf{D}}}}_{{{\mathbf{w}}}}}}\left( {\frac{{1 + 2\sigma /{{{\mathbf{P}}}}_a{{{\mathbf{r}}}} - {{{\mathbf{f}}}}}}{{1 + 3\sigma /4{{{\mathbf{P}}}}_a{{{\mathbf{r}}}}}}} \right)\), where Pa = 101.3 kPa is the hydrostatic pressure outside the bubble and r is the bubble radius. Here, L = 0.02 is Ostwald’s coefficient, Dw = 2 × 10−5cm2s−1 is the gas diffusivity in water, σ = 72 mNm−1 is the surface tension, and f = 1 is the ratio between the gas concentration in the medium versus that at saturation. This model assumes a perfectly spherical geometry and neglects the potentially stabilizing effects of the nearby collapsed GV shell. However, it provides useful simulations that illustrate the time constants relevant to the process of GV cavitation. (a) Radius-time curves of free air-filled bubbles of different initial sizes. The gas liberated from a collapsed GV occupied the volume of a sphere with a radius of 89 nm under atmospheric conditions and no surface tension, and is expected to have an initial radius slightly larger than 20 nm when surface tension between air and water is assumed across its surface. The actual initial radius is expected to be somewhere between these two values, depending on the degree of stabilization by collapsed GV shells or other solution components. (b) Time before 50% volume reduction for free air-filled bubbles of different sizes. These time constants support the ability of nanobubble to survive the half-cycle between GV collapse (peak pressure) and peak rarefaction. In addition, they can guide the selection of the pulse repetition interval after the initial bubble growth.

Extended Data Fig. 5 High frame rate optical imaging of GVs attached to tumour cells.

(a) GVs attached to U87 cells (0.4 µs) are collapsed by the ultrasound wave (0.8 µs). (b) Differential map comparing pre- and post-collapse images (c) Only after the collapse of the GVs are cavitation events seen (3.4 µs and 9.2 µs). The samples were insonated with a single 30-cycle pulse with PNP = 1.4 MPa a central frequency of 670 kHz. The representative result in panels a-c belongs to one of the 8 repeats presented in Fig. 3g. Scale bars represent 20 µm (a-c).

Extended Data Fig. 6 Targeted and expressed GVs are frequently grouped in close proximity.

The close proximity between expressed or targeted GVs could play an important role in GV cavitation (a) SEM image of Ana GVs attached to U87 cell, forming large patches. (b) TEM image of GVs expressed in S. typhimurium showing a large cluster. Scale bar is 5 μm (a), 1 μm (b). SEM scans of GVs attached to cells and TEM scans of GV expressing S. typhimurium were repeated more than 10 times, all with similar results showing patches or clusters of GVs.

Extended Data Fig. 7 Ultrasound images comparing GV expression in bacteria and mammalian cells.

(a) Ultrasound images of agarose phantoms containing S. typhimurium cells expressing GVs . The initial frame shows the echo from collapsing GVs (left, Peak US,), and the second one presents the residual signal from the cells after bubble dissolution (middle, Collapsed). The GV-specific signal, calculated as the difference between these two images, reveals high GV content in bacteria (right, Difference). (b) Ultrasound images of agarose phantoms containing GV-expressing HEK293 cells. The bacteria and mammalian cell samples were loaded into agarose phantoms at the same concentration as used in cavitation experiments. The volume of each well was 25 µl and it contained either 2 × 105 trypsinized mARG-HEK cells or OD600 = 1 (about 2 × 106) GV-expressing S. typhimurium cells. The combined volume of the mARG-HEK cells greatly exceeds the combined volume of the bacterial cells. However, the partial volume occupied by GVs in mammalian cell is much lower than in bacteria, resulting in lower GV-specific signal.

Extended Data Fig. 8 Color deconvolution of H&E stains reveals effects of GV cavitation on surrounding tissue.

Histologic stains of liver samples were collected after systemic saline injection followed by FUS exposure (negative control, a-d) or GV injection followed by FUS exposure (e-h). Color deconvolution was applied to H&E stains of liver sections (a, e) to obtain separate images of red blood cells (b, f) and tissue (c, g). The residual unmixed images are presented in (d, h). Necrotic regions in the H&E images (e, zoom-in in i) were found around hemorrhagic foci (f, zoom-in in j) in the livers of mice injected with GVs following FUS exposure. Scale bar is 2 mm (d, h), 200 μm (j). The representative results in this figure belong to one of the 5 repeats presented in Fig. 5g, h.

Extended Data Fig. 9 Flow Cytometry Quantification.

Gating strategy for quantifying cell death in mArg-HEK cells, including the SSC-FSC gating of one sample from each population and its Zombie NIR fluorescence histogram. Cell death was quantified by gating the fraction of cells that emitted Zombie NIR fluorescence. The cutoff was the same for all samples.

Supplementary information

Supplementary Information

Supplementary Tables 1–3.

Reporting Summary

41565_2021_971_MOESM3_ESM.avi

Supplementary Video 1 Representative high frame rate movie of GV attached to a Mylar plate. A series of 256 images showing cavitation nucleated by GVs attached to a Mylar plate were collected over 51.2 µs at 5 million frames per second (fps) The movie is displayed at 5 fps, 1 million times slower than the real time.

41565_2021_971_MOESM4_ESM.avi

Supplementary Video 2 Representative high frame rate movie of GVs attached to tumour cells. A series of 256 images showing cavitation nucleated by GVs attached to U87 tumour cells were collected over 51.2 µs at 5 million frames per second (fps) The movie is displayed at 5 fps, 1 million times slower than the real time. The swaying background is a result of the movement of the Mylar membrane at the bottom of the dish during of the ultrasound pulse.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bar-Zion, A., Nourmahnad, A., Mittelstein, D.R. et al. Acoustically triggered mechanotherapy using genetically encoded gas vesicles. Nat. Nanotechnol. 16, 1403–1412 (2021). https://doi.org/10.1038/s41565-021-00971-8

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41565-021-00971-8

This article is cited by

Search

Quick links

Nature Briefing: Translational Research

Sign up for the Nature Briefing: Translational Research newsletter — top stories in biotechnology, drug discovery and pharma.

Get what matters in translational research, free to your inbox weekly. Sign up for Nature Briefing: Translational Research