Hostname: page-component-7c8c6479df-995ml Total loading time: 0 Render date: 2024-03-28T02:44:33.689Z Has data issue: false hasContentIssue false

Free-Lagrange simulations of the expansion and jetting collapse of air bubbles in water

Published online by Cambridge University Press:  25 February 2008

C. K. TURANGAN
Affiliation:
School of Engineering Sciences, University of Southampton, Highfield, SO17 1BJ, UK
A. R. JAMALUDDIN
Affiliation:
School of Engineering Sciences, University of Southampton, Highfield, SO17 1BJ, UK
G. J. BALL
Affiliation:
Atomic Weapons Establishment, Aldermaston, Reading, RG7 4PR, UK
T. G. LEIGHTON*
Affiliation:
Institute of Sound & Vibration Research, University of Southampton, Highfield, SO17 1BJ, UK
*
Author to whom correspondence should be addressed: tgl@soton.ac.uk

Abstract

A free-Lagrange numerical method is implemented to simulate the axisymmetric jetting collapse of air bubbles in water. This is performed for both lithotripter shock-induced collapses of initially stable bubbles, and for free-running cases where the bubble initially contains an overpressure. The code is validated using two test problems (shock-induced bubble collapse using a step shock, and shock–water column interaction) and the results are compared to numerical and experimental results. For the free-running cases, simulations are conducted for a bubble of initial radius 0.3 mm located near a rigid boundary and near an aluminium layer (planar and notched surfaces). The simulations suggest that the boundary and its distance from the bubble influence the flow dynamics, inducing bubble elongation and jetting. They also indicate stress concentration in the aluminium and the likelihood of aluminium deformation associated with bubble collapse events. For the shock-induced collapse, a lithotripter shock, consisting of 56 MPa compressive and −10 MPa tensile waves, interacts with a bubble of initial radius 0.04 mm located in a free field (case 1) and near a rigid boundary (case 2). The interaction of the shock with the bubble causes it to involute and a liquid jet is formed that achieves a velocity exceeding 1.2 km s−1 for case 1 and 2.6 km s−1 for case 2. The impact of the jet on the downstream wall of the bubble generates a blast wave with peak overpressure exceeding 1 GPa and 1.75 GPa for cases 1 and 2, respectively. The results show that the simulation technique retains sharply resolved gas/liquid interfaces regardless of the degree of geometric deformation, and reveal details of the dynamics of bubble collapse. The effects of compressibility are included for both liquid and gas phases, whereas stress distributions can be predicted within elastic–plastic solid surfaces (both planar and notched) in proximity to cavitation events. There is a movie with the online version of the paper.

Type
Papers
Copyright
Copyright © Cambridge University Press 2008

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

Footnotes

Present address: Institute of High Performance Computing, Singapore Science Park II, Singapore 117528.

Present address: Romax Technology Limited, Nottingham Science & Technology Park, Nottingham, NG7 2PZ, UK.

References

REFERENCES

Bailey, M.R., Cleveland, R.O., Colonius, T., Crum, L. A., Evan, A. P., Lingeman, J. E., McAteer, J. A., Sapozhnikov, O. A. & Williams, J. C. Jr 2003 Cavitation in shock wave lithotripsy: the critical role of bubble activity in stone breakage and kidney trauma. Proc. 2003 Ultrasonics Symp. pp. 724–727.Google Scholar
Ball, G. J. 1996 A free-Lagrange method for unsteady compressible flow: simulation of a confined cylindrical blast wave. Shock Waves 5, 311325.CrossRefGoogle Scholar
Ball, G. J., Howell, B. P., Leighton, T. G. & Schofield, M. J. 2000 Shock-induced collapse of a cylindrical air cavity in water: a free-Lagrange simulation. Shock Waves 10, 265276.CrossRefGoogle Scholar
Blake, J. R., Hooton, M. C., Robinson, P. B. & Tong, R. P. 1997 Collapsing cavities, toroidal bubbles and jet impact. Phil. Trans. R. Soc. Lond. A 355, 537550.CrossRefGoogle Scholar
Bourne, N. K. & Field, J. E. 1991 Bubble collapse and the initiation of explosion. Proc. R. Soc. Lond. A Math. Phys. Engng Sci 435, 423435.Google Scholar
Bourne, N. K. & Milne, A. M. 2003 The temperature of shock-collapsed cavity. Proc. R. Soc. Lond. A Math. Phys. Engng Sci 459, 18511861.CrossRefGoogle Scholar
Brunton, J. H. 1967 In Proc. Intl Conf. Rain Eros., 2nd (ed. Fyall, A. A. & King, R. B.), p. 291. Royal Aircraft Establishment, Farnborough, UK.Google Scholar
Church, C. C. 1989 A theoretial study of cavitation generated by an extracorporeal shock wave lithotripter. J. Acoust. Soc. Am 86, 215227.CrossRefGoogle Scholar
Coleman, A. J., Choi, M. J., Saunders, J. E. & Leighton, T. G. 1992 Acoustic emission and sonoluminescence due to cavitation at the beam focus of an electrohydraulic shock wave lithotripter. Ultrasound Med. Biol 18, 267281.CrossRefGoogle ScholarPubMed
Cunningham, K. B., Coleman, A. J., Leighton, T. G. & White, P. R. 2001 Characterising in vivo acoustic cavitation during lithotripsy with time-frequency methods. Acoust. Bull. 26 (5), 1016.Google Scholar
Dear, J. P. & Field, J. E. 1988a High-speed photography of surface geometry effects in liquid/solid impact. J. Appl. Phys 63, 1015.CrossRefGoogle Scholar
Dear, J. P. & Field, J. E. 1988b A study of the collapse of arrays of cavities. J. Fluid. Mech 190, 409425.CrossRefGoogle Scholar
Dear, J. P., Field, J. E. & Walton, A. J. 1988 Gas compression and jet formation in cavities collapsed by a shock wave. Nature 332, 505.CrossRefGoogle Scholar
Ding, Z. & Gracewski, S. M. 1996 The behaviour of a gas cavity impacted by a weak or strong shock wave. J. Fluid Mech 309, 183209.CrossRefGoogle Scholar
Field, J. E. 1991 The physics of liquid impact, shock wave interactions with cavities, and the implications to shock-wave lithotripsy. Phys. Med. Biol. 36 (11), 14751484.CrossRefGoogle ScholarPubMed
Fong, S. W., Klaseboer, E., Turangan, C. K., Khoo, B. C. & Hung, K. C. 2006 Numerical anslysis of a gas bubble near bio-materials in an ultrasound field. Ultrasound Med. Biol 32, 925942.CrossRefGoogle Scholar
Hansson, I., Kendrinskii, V. & Morch, K. A. 1982 On the dynamics of cavity clusters. J. Phys. D: Appl. Phys 15, 17261734.CrossRefGoogle Scholar
Howell, B. P. & Ball, G. J. 2000 Damping of mesh-induced errors in free-Lagrange simulations of Richtmyer–Meshkov instability. Shock Waves 10, 253264.CrossRefGoogle Scholar
Howell, B. P. & Ball, G. J. 2002 A free-Lagrange augmented Godunov method for the simulation of elastic–plastic solids. J. Comput. Phys 175, 128167.CrossRefGoogle Scholar
Igra, D. & Takayama, K. 1999 Investigations of aerodynamic breakup of a cylindrical water droplet. Tech. Rep. 11. Institute of Fluid Science, Tohoku University.CrossRefGoogle Scholar
Jackson, M. J. & Field, J. E. 2000 Modelling liquid impact fracture thresholds in brittle materials. Brit. Ceramic Trans. 99 (1), 113.CrossRefGoogle Scholar
Jamaluddin, A. R., Ball, G. J. & Leighton, T. G. 2002 Free-Lagrange simulations of shock/bubble interaction in shock wave lithotripsy. Proc. Second Intl Conf. Computational Fluid Mech., pp. 541–546.Google Scholar
Klaseboer, E., Turangan, C. K., Fong, S. W., Liu, T. G., Hung, K. C. & Khoo, B. C. 2006 Simulations of pressure-pulse bubble interaction using boundary element method. Comput. Meth. Appl. Engng 195, 42874302.CrossRefGoogle Scholar
Kuwahara, M., Ioritani, N., Kambe, K., Orikasa, S. & Takayama, K. 1991 Anti-miss-shot control device for selective stone disintegration in extracorporeal shock wave lithotripsy. Shock Waves 1, 145148.CrossRefGoogle Scholar
Lauterborn, W. 1972 High-speed photography of laser-induced breakdown in liquids. Appl. Phys. Lett 21, 27.CrossRefGoogle Scholar
Leighton, T. G. 2004 From seas to surgeries, from babbling brooks to baby scans: the acoustics of gas bubbles in liquids. Intl J. Mod. Phys. B 18, 32673314.CrossRefGoogle Scholar
Leighton, T. G., Cox, B. T. & Phelps, A. D. 2000 The Rayleigh-like collapse of a conical bubble. J. Acoust. Soc. Am 107, 130142.CrossRefGoogle ScholarPubMed
Leighton, T. G., Fedele, F., Coleman, A. J., McCarthy, C., Ryves, S., Hurrell, A. M., DeStefano, A. Stefano, A. & White, P. R. 2008 A device for monitoring the efficacy of ESWL using passive acoustic emissions. Ultrasound Med. Biol. (submitted).Google Scholar
Lindau, O. & Lauterborn, W. 2003 Cinematographic observation of the collapse and rebound of a laser-produced cavitation bubble near a wall. J. Fluid Mech 479, 327348.CrossRefGoogle Scholar
Lush, P. A., Takayama, K., Tomita, T. & Obara, T. 1992 Cavitation and induced shock wave–bubble interaction as a cause of human tissue damage in extracorporeal shock wave lithotripsy. In Cavitation, Proc. IMech E, pp. 55–64.Google Scholar
Ohl, C. D., Kurz, T., Geisler, R., Lindau, O. & Lauterborn, W. 1999 Bubble dynamics, shock waves and sonoluminescence. Phil. Trans. R. Soc. Lond. A 357, 269294.CrossRefGoogle Scholar
Philipp, A. & Lauterborn, W. 1998 Cavitation erosion by single laser-produced bubbles. J. Fluid Mech 361, 75116.CrossRefGoogle Scholar
Philipp, A., Delius, M., Scheffczyk, C., Vogel, V. & Lauterborn, W. 1993 Interaction of lithotripter-generated shock waves with air bubbles. J. Acoust. Soc. Am 93, 24962509.CrossRefGoogle Scholar
Plesset, M. S. & Prosperetti, A. 1977 Bubble dynamics and cavitation. Annu. Rev. Fluid Mech 9, 145185.CrossRefGoogle Scholar
Rayleigh, Lord 1917 On the pressure developed in a liquid during the collapse of a spherical cavity. Phil. Mag 34, 9498.CrossRefGoogle Scholar
Sapozhnikov, O. A., Bailey, M. R., Crum, L. A., Miller, N. A., Cleveland, R. O., Pishchalnikov, Y. A., Pishchalnikova, I. V., McAteer, J. A., Connors, B. A., Blomgren, P. M. & Evan, A. P. 2001 Ultrasound-guided localized detection of cavitation during lithotripsy in pig kidney in vivo. Proc. 2001 Ultrasonics Symp. vol. 2, 13471350.Google Scholar
Shima, A. 1997 Studies of bubble dynamics. Shock Waves 7, 3342.CrossRefGoogle Scholar
Tomita, Y. & Shima, A. 1986 Mechanisms of impulsive pressure generation and damage pit formation by bubble collapse. J. Fluid Mech 169, 535564.CrossRefGoogle Scholar
Tomita, Y., Obara, T., Takayama, K. & Kuwahara, M. 1994 Cavitation phenomena in extracorporeal microexplosion lithotripsy. Shock Waves 3, 149157.CrossRefGoogle Scholar
Turangan, C. K. 2004 Free-Lagrange simulations of cavitation bubble collapse. PhD thesis, University of Southampton, UK.Google Scholar
Turangan, C. K., Ong, G. P., Klaseboer, E. & Khoo, B. C. 2006 Experimental and numerical study of transient bubble–elastic membrane interaction. J. Appl. Phys 100, 054910.CrossRefGoogle Scholar
Vogel, A., Lauterborn, W. & Timm, R. 1989 Optical and acoustic investigations of the dynamics of laser-produced cavitation bubbles near a solid boundary. J. Fluid Mech 206, 299338.CrossRefGoogle Scholar
Vyas, B. & Preece, C. M. 1976 Stress produced in a solid by cavitation. J. Appl. Phys 47, 51335138.CrossRefGoogle Scholar
Wilkins, M. L. 1999 In Computer Simulation of Dynamic Phenomena, pp. 128167. Springer.CrossRefGoogle Scholar
Zeman, R. K., Davros, W. J., Goldberg, J. A., Garra, B. S., Hayes, W. S., Cattau, E. L., Horii, S. C., Cooper, C. J. & Silverman, P. M. 1990 Cavitation effects during lithotripsy. Part II. Clinical observations. Radiology 177, 163166.CrossRefGoogle ScholarPubMed

Turangan et al. supplementary movie

Movie 1. This animation of simulated Schlieren images shows the collapse of a bubble in an infinite body of liquid, induced by the passage over the bubble of the pressure pulse from an extracorporeal shock wave lithotripter (travelling from left to right, and labelled IS in the movie and in figure 1(a), below). Except for the wall of the air bubble (which is initially spherical with radius 0.06 mm) and the horizontal line just below the middle of the image (which indicates the axis of rotational symmetry), the darker shading indicates perturbations of refractive index in the usual manner for Schlieren images. An expansion wave (travelling upwards and to the left; and labelled EX in the movie and in figure 1(a), below) is generated when the lithotripter pulse first impacts the bubble. The upstream bubble wall (on the left) then involutes and forms a liquid jet (labelled LJ), which accelerates to over 1200 m s-1 (figure 1(b), below) before impacting the downstream bubble wall. This impact produces a blast wave (labelled BW in the movie and figure 1(a)) that generates a spatial peak pressure of over 1 GPa. Further details on this animation can be found below the figures, at the base of this page.

Download Turangan et al. supplementary movie(Video)
Video 827.4 KB

Turangan et al. supplementary movie

Movie 1. This animation of simulated Schlieren images shows the collapse of a bubble in an infinite body of liquid, induced by the passage over the bubble of the pressure pulse from an extracorporeal shock wave lithotripter (travelling from left to right, and labelled IS in the movie and in figure 1(a), below). Except for the wall of the air bubble (which is initially spherical with radius 0.06 mm) and the horizontal line just below the middle of the image (which indicates the axis of rotational symmetry), the darker shading indicates perturbations of refractive index in the usual manner for Schlieren images. An expansion wave (travelling upwards and to the left; and labelled EX in the movie and in figure 1(a), below) is generated when the lithotripter pulse first impacts the bubble. The upstream bubble wall (on the left) then involutes and forms a liquid jet (labelled LJ), which accelerates to over 1200 m s-1 (figure 1(b), below) before impacting the downstream bubble wall. This impact produces a blast wave (labelled BW in the movie and figure 1(a)) that generates a spatial peak pressure of over 1 GPa. Further details on this animation can be found below the figures, at the base of this page.

Download Turangan et al. supplementary movie(Video)
Video 420.3 KB
Supplementary material: Image

Turangan et al. supplementary material

Figure 1. (a) The time histories of bubble volume (solid line) and the pressure as measured at a point on the axis of symmetry, 0.18 mm upstream of the initial position of the centre of the bubble (broken line). The labels show the moments when the pressure at that point is dominated by the lithotripter shock (IS), the expansion wave (EX), and the blast wave (BW). The datum of time is the moment when the lithotripter shock first impacts the upstream bubble wall (corresponding to t=0.25 μs in the above animation). (b) The time histories of the speeds of the points on the bubble wall where the upstream (solid line) and downstream (broken line) walls cross the axis of rotational symmetry. Once the jet has formed, the speed of its tip corresponds to the speed of the upstream wall (solid line). The datum of time is as for part (a)

Download Turangan et al. supplementary material(Image)
Image 476.5 KB