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.

  • Letter
  • Published:

Leidenfrost wheels

A Publisher Correction to this article was published on 01 October 2018

This article has been updated

Abstract

As reported in 1756 by Johann Gottlob Leidenfrost, volatile liquids on hot solids form “gleaming drops resembling quicksilver”, a consequence of their levitation on a vapour cushion1,2. This makes the drops spectacularly mobile, moving away as soon as they are deposited—an observation commonly attributed to gravity or surrounding airflows. This mobility has been exploited to manipulate drops, because tiny forces such as those generated on asymmetric substrates can move them in well-defined directions3,4,5, a situation that also provides heat evacuation6. Here we report that Leidenfrost droplets initially at rest on horizontal substrates self-rotate and self-propel in the direction they are rolling, in the absence of any source of asymmetry or external force. Their rapid internal flow is found to be accompanied by a tilting of their base, which creates a permanent ratchet-like mechanism, entraining the rolling liquid despite the fact that it is not in contact with its substrate.

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: Behaviour of Leidenfrost drops starting from rest on flat silicon wafers.
Fig. 2: Dynamics of self-propelled Leidenfrost drops.
Fig. 3: Influence of the drop shape on its inner dynamics.
Fig. 4: Focus on the base of self-propelling drops.

Similar content being viewed by others

Data availability

The data that support the plots within this paper and other findings of this study are available from the corresponding author upon request.

Change history

  • 01 October 2018

    In the version of this Letter originally published, the Supplementary Videos were incorrectly labelled; the descriptions of 1–4 should have gone with the videos of 6–9, and the descriptions of 5–9 should have gone with the videos of 1–5. This has now been corrected.

  • 05 May 2020

    In the original online version of this Letter, the links to Supplementary Videos 1–7 pointed to the incorrect files; those for Videos 1–4 pointed to the files for Videos 6–9, respectively, and those for Videos 5–9 pointed to the files for Videos 1–5. The links have now been amended.

References

  1. Boerhaave, H. Elementae Chemiae Vol. 1 (Lugduni Batavorum, Leiden, 1732).

  2. Leidenfrost, J. G. De Aquae Communis Nonnullis Qualitatibus Tractatus (Ovenius, Duisburg, 1756); transl. Wares, C. On the fixation of water in diverse fire. Int. J. Heat Mass Trans. 9, 1153–1166 (1966).

    Article  Google Scholar 

  3. Linke, H. et al. Self-propelled Leidenfrost droplets. Phys. Rev. Lett. 96, 154502 (2006).

    Article  ADS  Google Scholar 

  4. Marín, A. G. et al. Capillary droplets on Leidenfrost micro-ratchets. Phys. Fluids 24, 1–9 (2012).

    Article  Google Scholar 

  5. Dupeux, G. et al. Self-propelling uneven Leidenfrost solids. Phys. Fluids 25, 051704 (2013).

    Article  ADS  Google Scholar 

  6. Singh Dhillon, N., Buongiorno, J. & Varanasi, K. K. Critical heat flux maxima during boiling crisis on textured surfaces. Nat. Commun. 6, 8247 (2015).

    Article  ADS  Google Scholar 

  7. Bernardin, J. D. & Mudawar, I. The Leidenfrost point: experimental study and assessment of existing models. J. Heat Transf. 121, 894–903 (1999).

    Article  Google Scholar 

  8. Bernardin, J. D. & Mudawar, I. Film boiling heat transfer of droplet streams and sprays. Int. J. Heat Mass Transf. 40, 2579–2593 (1997).

    Article  Google Scholar 

  9. Weickgenannt, C. M. et al. Inverse-Leidenfrost phenomenon on nanofiber mats on hot surfaces. Phys. Rev. E 84, 036310 (2011).

    Article  ADS  Google Scholar 

  10. Arnaldo del Cerro, D. et al. Leidenfrost point reduction on micropatterned metallic surfaces. Langmuir 28, 15106–15110 (2012).

    Article  Google Scholar 

  11. Vakarelski, I. U., Patankar, N. A., Marston, J. O., Chan, D. Y. C. & Thoroddsen, S. T. Stabilization of Leidenfrost vapour layer by textured superhydrophobic surfaces. Nature 489, 274–277 (2012).

    Article  ADS  Google Scholar 

  12. Biance, A. L., Clanet, C. & Quéré, D. Leidenfrost drops. Phys. Fluids 15, 1632–1637 (2003).

    Article  ADS  Google Scholar 

  13. Gottfried, B. S., Lee, C. J. & Bell, K. J. The Leidenfrost phenomenon: film boiling of liquid droplets on a flat plate. Int. J. Heat Mass Trans. 9, 1167–1188 (1966).

    Article  Google Scholar 

  14. Myers, T. G. & Charpin, J. P. F. A mathematical model of the Leidenfrost effect on an axisymmetric droplet. Phys. Fluids 21, 63101–1632 (2009).

    Article  Google Scholar 

  15. Cousins, T. R., Goldstein, R. E., Jaworski, J. W. & Pesci, A. I. A ratchet trap for Leidenfrost drops. J. Fluid Mech. 696, 215–227 (2012).

    Article  ADS  Google Scholar 

  16. Burton, J. C., Sharpe, A. L., Van der Veen, R. C. A., Franco, A. & Nagel, S. R. Geometry of the vapor layer under a Leidenfrost drop. Phys. Rev. Lett. 109, 074301 (2012).

    Article  ADS  Google Scholar 

  17. Pomeau, Y., Le Berre, M., Celestini, F. & Frisch, T. The Leidenfrost effect: from quasi-spherical droplets to puddles. C. R. Mec. 340, 867–881 (2012).

    Article  ADS  Google Scholar 

  18. Duchemin, L., Lister, J. R. & Lange, U. Static shapes of levitated viscous drops. J. Fluid Mech. 533, 161–170 (2005).

    Article  ADS  MathSciNet  Google Scholar 

  19. Snoeijer, J. H., Brunet, P. & Eggers, J. Maximum size of drops levitated by an air cushion. Phys. Rev. E 79, 036307 (2009).

    Article  ADS  Google Scholar 

  20. Quéré, D. Leidenfrost dynamics. Annu. Rev. Fluid Mech. 45, 197–215 (2013).

    Article  ADS  MathSciNet  Google Scholar 

  21. Mahadevan, L. & Pomeau, Y. Rolling droplets. Phys. Fluids 11, 2449–2453 (1999).

    Article  ADS  MathSciNet  Google Scholar 

  22. Wciślik, S. Thermal infrared mapping of the Leidenfrost drop evaporation. J. Phys. Conf. Ser. 745, 32064 (2016).

    Article  Google Scholar 

  23. Mrinal, M., Wang, X. & Luo, C. Self-rotation induced propulsion of a Leidenfrost drop on a ratchet. Langmuir 33, 6307–6313 (2017).

    Article  Google Scholar 

  24. Touihri, R., Ben Hadid, H. & Henry, D. On the onset of convective instabilities in cylindrical cavities heated from below. I. Pure thermal case. Phys. Fluids 11, 2078–2088 (1999).

    Article  ADS  MathSciNet  Google Scholar 

  25. Tam, D., von Arnim, V., McKinley, G. H. & Hosoi, A. E. Marangoni convection in droplets on superhydrophobic surfaces. J. Fluid Mech. 624, 101–123 (2009).

    Article  ADS  Google Scholar 

  26. Dash, S., Chandramohan, A., Weibel, J. A. & Garimella, S. V. Buoyancy-induced on-the-spot mixing in droplets evaporating on nonwetting surfaces. Phys. Rev. E 90, 062407 (2014).

    Article  ADS  Google Scholar 

  27. Duchesne, A., Savaro, C., Lebon, L., Pirat, C. & Limat, L. Multiple rotations of a drop rolling inside a horizontal circular hydraulic jump. Eur. Phys. Lett. 102, 64001 (2013).

    Article  ADS  Google Scholar 

  28. Sobac, B., Rednikov, A., Dorbolo, S. & Colinet, P. Self-propelled Leidenfrost drops on a thermal gradient: a theoretical study. Phys. Fluids 29, 082101 (2017).

    Article  ADS  Google Scholar 

  29. Couder, Y., Fort, E., Gautier, C. H. & Boudaoud, A. From bouncing to floating: noncoalescence of drops on a fluid bath. Phys. Rev. Lett. 94, 177801 (2005).

    Article  ADS  Google Scholar 

  30. Dorbolo, S., Terwagne, D., Vandewalle, N. & Gilet, T. Resonant and rolling droplet. New J. Phys. 10, 113021 (2008).

    Article  ADS  Google Scholar 

  31. Kang, K. H., Lee, S. J., Lee, C. M. & Kang, I. S. Quantitative visualization of flow inside an evaporating droplet using the ray tracing method. Meas. Sci. Technol. 15, 1104–1112 (2004).

    Article  ADS  Google Scholar 

Download references

Acknowledgements

We thank C. Frot for her help in designing the set-up, J. Quintela Casal for preliminary experiments and C. Josserand and É. Pirot for fruitful discussions.

Author information

Authors and Affiliations

Authors

Contributions

T.M., P.B., C.C. and D.Q. conceived the project. A.B., T.M., P.B. and D.Q. designed the project. A.B. performed most experiments and analyses to which A.L. also contributed. A.B., T.M., C.C. and D.Q. discussed the models. A.B. and D.Q. wrote the manuscript with inputs from all other authors.

Corresponding author

Correspondence to David Quéré.

Ethics declarations

Competing interests

The authors declare no competing interests.

Additional information

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

Supplementary information

Supplementary Information

Supplementary Information, Supplementary Figures 1–10

Supplementary Video 1

Side view of the experiment shown in Fig. 1 slowed down by a factor of 40. A water drop with radius R = 1.10 mm is dispensed from a needle at the centre of a hot wafer (T = 350 °C). Tracers (predispersed in the liquid) reveal an internal rolling motion before and after the drop detaches from the needle and self-propels in the rolling direction, like a liquid wheel

Supplementary Video 2

Animation associated with Fig. 1c,d showing top-view trajectories of drops with radii R = 1.0 mm (left), R = 1.5 mm (centre) and R = 2.0 mm (right) sped up by a factor 20. The surface is a 10-cm-diameter reflective wafer heated at T = 300 °C. The grey zone on the bottom right is hidden by the needle and experimentally inaccessible. Drops with radius R = 1 mm all self-propel after detachment with straight, isotropic trajectories. In stark contrast, drops with R = 2 mm follow roughly straight trajectories in a biased direction. In between (R = 1.5 mm)—that is, on the brink of the onset of self-propulsion—drops move away along the biased direction but can turnaround and adopt a random propelling direction.

Supplementary Video 3

Internal motion in a Leidenfrost drop (R = 2.5 mm) slowed down by a factor of 10 (corresponding to Fig. 3a). The drop is immobilized in a groove heated at T = 350 °C and illuminated with a 400-µm-thick laser sheet. Tracers rise up along the interface and pursue by a downward motion along the drop centre.

Supplementary Video 4

Internal motion in a Leidenfrost droplet (R = 1.1 mm) slowed down by a factor of 25 (corresponding to Fig. 3b). The drop is immobilized by a needle on a flat wafer heated at T = 360 °C and illuminated with a laser sheet. The droplet rolls with an angular velocity of 85 rad s–1 and the flow is found to be stable in time.

Supplementary Video 5

Surface flows of a Leidenfrost drop viewed from the top throughout its life played in real time. A puddle (R ≈ 3.5 mm) is initially deposited in a groove heated at T = 350 °C. Hydrophobic tracers standing at the interface draw convective chaotic patterns that tend to organize into a four-cell symmetric structure. For R ≈ 1.8 mm, drop vibrations intensify and the symmetry breaks. Then, the drop starts to roll. Rolling persists until tracers saturate and form a static solid shell at the surface. The white bar indicates 2 mm.

Supplementary Video 6

Simultaneous PIV and interferometric visualization of the bottom interface of a Leidenfrost droplet (R = 0.73 mm) slowed down by a factor of ten (corresponding to Fig. 4a). The drop is immobilized by a needle on a transparent sapphire heated at T = 300 °C. Tracers near the surface move along the symmetry axis Ox of the interference pattern, showing a correlation between the inner flow in the drop and the deformation of the vapour cushion. The drop detaches at the end of the video, and the contact zone (that is, the drop) is found to accelerate along the Ox-direction—that is, in the direction of the tilt of the bottom interface. We also observe a few oscillations caused by the detachment. The white bar indicates 200 µm.

Supplementary Video 7

Side view of a submillimetric droplet levitating on a plate heated at 340 °C and disturbed by vertical oscillations. The black bar indicates 1 mm, and the video is slowed down by a factor of 100. A perturbation of the drop’s vertical position changes heat exchange and generates surface waves propagating from the bottom of the drop, likely to disturb the drop dynamics after detachment.

Supplementary Video 8

Side view of a water drop levitating on a plate heated at 350 °C observed with an infrared camera using a calibration range from −40 °C to 150 °C, only suitable for water, and not brass. The bar indicates 5 mm, and the video is sped up by a factor of 3.2, while the right-handed lateral colour bar gives access to the surface temperature of the drop.

Supplementary Video 9

Top view of a large puddle (R ≈ 4 mm) levitating on a Glaco-coated unpolished silicon wafer heated at 200 °C. The video is sped up by a factor of 20. The restricted mobility at the beginning can be accounted for adhesion of water on such surfaces, resulting from intermittent contact with roughness. Once the drop radius becomes millimetric, water suddenly chooses one direction and promptly leaves the substrate.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Bouillant, A., Mouterde, T., Bourrianne, P. et al. Leidenfrost wheels. Nature Phys 14, 1188–1192 (2018). https://doi.org/10.1038/s41567-018-0275-9

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/s41567-018-0275-9

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