ArticlePDF Available

Experiments on bubble dynamics between a free surface and a rigid wall

Authors:

Abstract and Figures

Experiments were conducted where the underwater bubble oscillates between two boundaries, a free surface and a horizontal rigid wall. The motion features of both the bubble and the free surface were investigated, via the consideration of two key factors, i.e., the non-dimensional distances from the bubble to the two boundaries. To support the investigation, experiments were conducted in the first place where the bubble oscillates near only one of the two boundaries. Then the other boundary was inserted at different positions to observe the changes in the motion features, including the types, maximum speed and height of the water spike and skirt, the form and speed of the jets, and bubble shapes. Correspondence is found between the motion features of the free surface and different stages of bubble oscillation. Intriguing details such as gas torus around the jet, double jets, bubble entrapment, and microjet of the water spike, etc., are observed.
Content may be subject to copyright.
RESEARCH ARTICLE
Experiments on bubble dynamics between a free surface
and a rigid wall
A. M. Zhang P. Cui Y. Wang
Received: 8 April 2013 / Revised: 17 August 2013 / Accepted: 14 September 2013 / Published online: 28 September 2013
ÓSpringer-Verlag Berlin Heidelberg 2013
Abstract Experiments were conducted where the under-
water bubble oscillates between two boundaries, a free
surface and a horizontal rigid wall. The motion features of
both the bubble and the free surface were investigated, via
the consideration of two key factors, i.e., the non-dimen-
sional distances from the bubble to the two boundaries. To
support the investigation, experiments were conducted in
the first place where the bubble oscillates near only one of
the two boundaries. Then the other boundary was inserted
at different positions to observe the changes in the motion
features, including the types, maximum speed and height of
the water spike and skirt, the form and speed of the jets,
and bubble shapes. Correspondence is found between the
motion features of the free surface and different stages of
bubble oscillation. Intriguing details such as gas torus
around the jet, double jets, bubble entrapment, and microjet
of the water spike, etc., are observed.
1 Introduction
Cavitation bubble has been studied for more than a century
and still remains at the center of attention in various fields.
The interest in bubble interaction with boundaries mainly
arose from the damage effect to nearly all known surfaces,
such as ship propellers, turbines, and other fluid machinery.
Researches are also motivated by the important role of
bubble in a wide variety of applications from ocean engi-
neering to medical science.
Spherical bubble collapse in infinite free fluid field can
be roughly modeled by the Rayleigh–Plesset equations
(Rayleigh 1917) or by other models involving compress-
ibility, viscosity, heat transfer, etc. (Gilmore 1952; Prosp-
eretti and Lezzi 1986). On the other hand, the collapse of
bubbles near boundaries will be asymmetrical, with the
formation of liquid jet in suitable conditions. The physical
property of the boundary is crucial to bubble collapse,
especially the direction of bubble jet. Various experiments
of bubble collapse near free surface were carried out (Blake
and Gibson 1981; Chahine 1977; Dadvand et al. 2009;
Robinson et al. 2001, etc.) in which the bubble migrates
and produces a jet away from the boundary and becomes
toroidal. Numerical simulations were also carried out on
bubble-free surface interaction using the boundary integral
method (Blake and Gibson 1981; Blake et al. 1987; Pear-
son et al. 2004b; Wang et al. 1996, etc.). Most of the
simulations were focused on the collapse phase before jet
penetration, while Pearson et al. (2004b), Wang et al.
(1996), etc., obtained the subsequent toroidal bubble, and
the simulations stopped before bubble rebound. In the
current work, more detailed observations are made to the
late collapse phase of the bubble, including a second
toroidal bubble, bubble rebounds, as well as the scenario
where two jets were found within a single collapse. The
free surface deformation motivated by the bubble was also
discussed in the above works and by Longuet-Higgins
(1983), etc. Variations in the width, height, and speed of
the free surface liquid jet (referred to as water ‘‘spike’’)
with the standoff distance of the bubble were discussed. In
suitable conditions, the ‘‘lateral jets’’ (lateral deformation
on the spike, referred to as water ‘‘skirt’’) were observed.
The numerical simulations were mostly focused on the
A. M. Zhang (&)P. Cui
College of Shipbuilding Engineering, Harbin Engineering
University, Harbin 150001, China
e-mail: zhangaman@hrbeu.edu.cn
Y. Wang
Naval Academy of Armament, Beijing 100161, China
123
Exp Fluids (2013) 54:1602
DOI 10.1007/s00348-013-1602-7
bubble rather than on the water spike that keeps evolving
for a long time after bubble rebound; fewer simulations
were focused on the water skirt. In this paper, a number of
experiments focused on free surface motions are carried
out over a range of standoff distance, in order to obtain a
collection of typical morphologies of the water spike and
skirt. Besides, a coincidence between the different stages of
bubble oscillation and the features of free surface motion is
found, which leads to a discussion on the formation of the
water skirt.
More extensive and abundant researches were carried
out on bubble collapse in the neighborhood of solid
boundaries (see, e.g., Naude
´and Ellis (1961), Lauterborn
and Bolle (1975), Blake and Gibson (1987), Philipp and
Lauterborn (1998), Shima et al. (1981), etc.). Apart from
the bubble jet directed toward the boundary, efforts were
also focused on the formation and collapse of toroidal
bubble (Brujan et al. 2002; Lindau and Lauterborn 2003;
Vogel et al. 1989), shock wave emission by bubble
rebound (Blake et al. 1998; Lindau and Lauterborn 2003;
Shima et al. 1983), and damage done to the boundary
(Bourne and Field 1995; Philipp and Lauterborn 1998;
Tomita and Shima 1986). In the above studies, a second
toroidal cavity was observed after jet penetration with
larger non-dimensional standoff; with smaller standoff,
the splashing of the jet against the boundary was studied.
Numerical simulations related to solid boundaries are
also prosperous. Many have successfully simulated the
toroidal bubble after jet penetration and the splashing
(Best 1993; Blake et al. 1997,1999; Brujan et al. 2002;
Pearson et al. 2004a; Tong et al. 1999; Wang et al.
2005; Zhang et al. 1993), though it is still hard to
numerically treat small-standoff cases and the bubble
rebound accurately. Interest of this paper is therefore on
the effect of the bubble rebound on the water skirt,
especially from those against the solid wall in the case
of small standoff distances.
In addition to the free surface and the solid boundary,
bubble collapse near elastic boundary was also examined
both experimentally and numerically (Brujan et al. 2001;
Shaw et al. 1999; Turangan et al. 2006, etc.), where the
bubble may produce a jet away from the boundary or split.
Bubble dynamics were well studied in the neighborhood of
a single boundary in the above-mentioned papers. Never-
theless, works are rare involving multiple boundaries,
especially those of different natures. In this paper, experi-
ments will be conducted where a bubble is initiated
between a free surface and a horizontal rigid wall. The
forms of the water spike and skirt, their height and speed,
the speed of the reentrant jets, and other bubble behaviors
are investigated by keeping one of the two standoffs con-
stant and the other varied.
2 Experimental setup
2.1 Bubble generation and recording
The low-voltage electric-spark bubble-generating method is
adopted in the current work. It was initially applied by Tur-
angan et al. (2006) and then modified over time (Dadvand
et al. 2009;Fongetal.2009). This method is suitable for the
current experiment because the period of the bubble generated
is longer (*5 ms in the following experiments) and the radius
is larger (*10 mm) compared to a laser-induced bubble [e.g.,
Lindau and Lauterborn (2003)], which will lower the
requirements in photographing and enables us to capture more
motion details with higher spatial and temporal resolution,
especially tocapture the bubble and the water spike withinthe
same video. The circuit adopted in the current work is based
on that of Zhang et al. (2011), as demonstrated in Fig. 1.The
simple circuit includes a 6,600 lF capacitor and a 200 V DC
power supply.Two copper-wire electrodes from the two poles
of the capacitor are in contact with each other at the ‘‘con-
nection point’’. With the discharge of the capacitor,the copper
wires are ignited into bright plasma and a bubble is generated.
According to repetitive experiments, the maximum average
radius of a bubble is closely related to the duration of the
plasma and falls between 11.50 ±0.60 mm when the dura-
tion is controlled within 2.5–3.0 ms. In very few cases is the
average radius or the plasma duration out of the above-men-
tioned ranges, and such results are discarded to ensure the
repeatability of the maximum radius. Despite the unstable
plasma, the bubble center (in free field) always coincides with
the connection point of the copper wire, which is, therefore,
referred to as the ‘‘initial bubble center’’.
Fig. 1 The experimental setup
Page 2 of 18 Exp Fluids (2013) 54:1602
123
To capture the motion of the bubble and the free surface,
a high-speed camera (Phantom V12.1) is adopted. The
camera works at 24,000 frames per second, with an
exposure time of 10 ls for each frame. The whole exper-
iment section is placed in a 0.5 90.5 90.5 m glass tank
and illuminated by a continuous light source from oppose
to the camera. A piece of polished glass is used as the rigid
bottom wall, which is adjusted to parallelism with the static
water surface. The distance from the tank boundary to the
bubble is more than 20 times of the bubble maximum
radius; hence, its effect could be neglected. Also ignored is
the effect of the electrodes, which have a diameter of
0.25 mm, only about 1.3 % of the bubble diameter, and
unlikely to cause substantial influence to the bubble
behaviors. The experiments are conducted at room tem-
perature (22 °C) and atmospheric pressure; the pressure
discrepancy between bubble top and bottom is negligible
given the small radius, and hence, the buoyancy or gravity
effect will be overcome by inertia. In fact, neither jet nor
upward migration of the bubbles is observed in free fluid
field.
The content of the bubble may include water vapor and
electrolysis products; hence, the non-equilibrium vapor
condensation could exist and cause additional cushioning
of the bubble collapse as argued by Fujikawa and Ak-
amatsu (1980).
2.2 Measurements and accuracy
The capturing time of the last image before the plasma is
taken as time zero. The maximum error in time measuring
is the image interval, 41.66 ls, which is small enough
compared to the first bubble oscillation period (*5 ms).
Before bubble generation, a ruler is placed at the initial
center perpendicular to the lens axle and recorded as length
calibration with which the spatial measurements are
directly carried out over discrete pixels of the captured
images. Hence, the special precision is up to the pixel
length, calculated to be 0.12 and 0.30 mm for the cases
with and without water spike, respectively. In addition,
there are blurs in some images on object edges with a
typical thickness of 1–2 pixels; therefore, the spatial error
could be 0.12–0.24 mm (for bubble) and 0.30–0.60 mm
(for water spike).
An equivalent maximum radius R
eq
is adopted since in
many cases the bubbles develop into non-spherical shapes.
Thirty points are picked around the bubble’s circumference
on the image captured at maximum expansion to calculate
the enclosed area A. The equivalent radius is then defined,
inversing the formula of circle area, as Req ¼ffiffiffiffiffiffiffiffi
A=p
p. For
spherical bubble, R
eq
is 0.37 % smaller than actual radius
due to the use of discrete points to represent the circle.
Considering the error in length (0.12–0.24 mm), the final
error in R
eq
is calculated to be 0.17–0.34 mm, i.e.,
1.5–3.0 % for an 11.50-mm bubble. As displayed in the
camera view in Fig. 1, there are two length parameters in
the current work, namely
1. the non-dimensional standoff distance from the free
surface, defined as c
f
=d
f
/R
eq
2. the non-dimensional standoff distance from the rigid
wall, defined as c
b
=d
b
/R
eq.
where d
f
and d
b
are the vertical distances from the initial
center of the bubble to the free surface and the rigid bot-
tom, respectively. The two parameters are found to be the
major factors affecting the motion features of the bubble
and the free surface.
The speed of a jet is measured at its tip. The speed of a
water spike is measured at its highest point or the upper edge
of the first droplet (if present) is pinched off from the spike.
The spike height is measured vertically from the static free
surface to the top. As for the water skirt, its speed and height
are measured at the highest point of the integrated liquid bulk,
ignoring the liquid droplets sprayed around.
The Weber number (We), the Reynolds number (Re), the
Froude number (Fr), and the Ohnesorge number (Oh) are
introduced as follows: We =qDU
2
/r,Re =UD/v,
Oh ¼vffiffiffiffiffiffiffiffiffiffi
q=rL
p, and Fr ¼U=ffiffiffiffiffiffi
gD
pto delineate the differ-
ent regimes of the free surface motions observed, where q,
v,r,D, and Uare density, kinetic viscosity and surface
tension coefficient of water (at 22 °C), characteristic
length, and characteristic velocity, respectively. In the
current experiments, Re is between 10
4
and 10
5
for bubble
jets of 10–20 m/s in velocity with Dbeing jet width
(1–5 mm), up to 1.2 910
5
for bubble-wall motion with
Dbeing instantaneous bubble diameter, and of order 10
4
for jet torus migration with Dbeing torus diameter; hence,
the viscous effect of the bubble is less significant against
inertia. For the motion of water spikes and skirts with
maximum speed being 1–10 m/s, Fr is relatively low
(approximately 2.0–21.0) with Dset as the maximum
bubble diameter.
3 Experiment results
The aim of the experiments is to observe the dynamics of
the bubble and the free surface when the bubble oscillates
between two different boundaries; however, the experi-
ments where only one of the boundaries exists are carried
out in the first place, for the following reasons.
1. Some major features are not unique to the two-
boundary cases and will be better illustrated with
single-boundary experiments.
Exp Fluids (2013) 54:1602 Page 3 of 18
123
2. The single-boundary cases are necessary references to
investigate the changes of these features when a
second boundary is inserted.
3. Some of the features in single-boundary cases that are
not emphasized or observed in previous studies are
presented in the current work to facilitate the mor-
phology analysis in two-boundary cases.
4. The temporal and spatial scale and the generation
method can be kept identical between the single-
boundary and the two-boundary cases.
3.1 Single boundary: the free surface
3.1.1 Bubble motion
First of all, the behavior of the spark-generated oscillatory
bubble near the free surface is investigated. The time
evaluation of a typical case, with c
f
=0.72, is demon-
strated in Fig. 2. Sphericity of the bubble is retained (frame
1) until the end of the growth. The upper wall of the bubble
involutes away from the surface as the contraction begins
and then produces a reentrant jet (frame 2) with a maxi-
mum speed of 17.2 m/s.
The liquid jet hits the lower side of the collapsing
bubble which then becomes toroidal. A protrusion is pro-
duced (frame 4 and onward) which is also a toroidal gas
bubble centered around the ongoing jet. To avoid confu-
sion, we refer to the main bubble as the ‘‘toroidal bubble’
and the protrusion as the ‘‘jet torus’’ since it is drawn by the
jet. The morphology of two parts (frames 1–5) linked was
numerically simulated by Pearson et al. (2004b) (see
Figs. 5,6therein).
The motion of the toroidal bubble is characterized by a
rapid diminishing in volume (frames 5–7) until rebounding
in frame 7. The content of the main bubble is being
squeezed into the jet torus possibly under high pressure of
non-condensable gas and uncondensed vapor inside the
bubble. At the rebound, the jet torus breaks away from the
toroidal bubble (frame 7), migrates downward, and invo-
lutes. The involution can be told by the motion of the
fringes of the torus, as marked in frames 8 and 10.
A collapse generating two tori was reported experimen-
tally in the vicinity of solid boundaries (Brujan et al. 2002;
Zhang et al. 1993). However, to the best knowledge of the
authors, there is a lack of report on such collapse near the
free surface, except for that of Robinson et al. (2001), but
for detailed observation, higher resolution would be pre-
ferred. This can be achieved by, e.g., generating a larger
bubble and lowering the frame rate requirement, as carried
out in this paper. In addition, a schematic diagram is drawn
in Fig. 3a to demonstrate the sectional profiles of the bubble
corresponding to the captured frames. In the 3-D toroidal
bubble simulations by Li et al. (2012), the jet torus is not
present probably due to mesh-related issues.
3.1.2 Instability and shock wave emission
It should be noted that instability of the bubble surface
(Brennen 2002; Menon and Lal 1998) occurs during the
collapse phase. The collapse pauses, and the bubble
undergoes a slight re-expansion with its surface becoming
less smooth between frames 4 and 5, probably a result of
non-equilibrium condensation (i.e., vapor condensation
rate being lower than reducing rate of bubble volume) and
Fig. 2 Bubble collapse near free surface, c
f
=0.72. The frame number is placed at the corner of each frame, with its capturing time (in ms) in
the square bracket. The motion features are labeled in italics; bubble rebounds are captured at minimum volume. Same for the figures hereinafter
Page 4 of 18 Exp Fluids (2013) 54:1602
123
existence of non-condensable gas contents. Brennen (2002)
suggested that the Rayleigh–Taylor instability is likely to
be excited. Very soon the collapse resumes.
It was theoretically predicted that a local high pressure is
induced in the final stage of bubble collapse (Rayleigh 1917).
The inertia of the contracting liquid is overcome by the gas
pressure inside the bubble, which increases dramatically with
the bubble collapse. The bubble then rebounds, and a shock
wave is emitted to the liquid. The shock wave has been
visualized by Shima et al. (1981), Shaw et al. (1996), Tomita
and Shima (1986), Ward and Emmony (1991), Ohl et al.
(1995), and others in later works. For the toroidal bubble here
near the free surface, a ‘‘rebound’’ is also observed (see
frames 7–8). Shock wave is not directly seen due to exposure
limitations, but can be indicated. Debris of copper was cata-
pulted out during the plasma stage, piercing through the
bubble surface (see the spots around the bubblein frames 1–2
in Figs. 2and 4, especially on the bottom-left side). The
debris is merely *0.5 mm in diameter and unlikely to have
substantial influence on bubble motion, but brings minor
portions of gas away from the bubble to form microbubbles
appearing as the dark spots suspending around the main
bubble. Right after the rebound, the microbubbles are sud-
denly excited into pulsations. For instance, the microbubble
boxed out in frame 7 in Fig. 2can be seen to expand, contract,
and expand again in a chronological series of magnified
subgraphs on the bottom-left side of the figure. According to
Ohl and Ikink (2003) and Tomita and Shima (1986), micro-
bubbles were found to produce forced pulsation when
impacted by shock waves. Therefore, it is reasonable to
assume that in the current experiments, a shock wave is
radiated and acts upon the microbubbles, especially consid-
ering that shock wave emission was readily observed by
Brujan et al. (2002), Lindau and Lauterborn (2003), etc., from
similar collapsing toroidal bubble near a rigid boundary.
3.1.3 Thin jet formation and the double-jet scenario
With different c
f
, the bubble motion possesses practically
the same features as described above. Nevertheless, there is
(a) (b)
Fig. 3 Schematic diagram for bubble collapse near free surface, ac
f
=0.72; profiles AEcorrespond to frames 1, 2, 4, 6, and 7, respectively, in
Fig. 2.bc
f
=0.51; profiles AEcorrespond to frames 3–5 and 8–9, respectively, in Fig. 4
Fig. 4 Bubble collapse with two jets (the double-jet scenario), c
f
=0.51. The pulsation of the microbubbles (e.g., marked out in frame 9) is
shown in a series of magnified subgraphs with capturing time labeled at the corner
Exp Fluids (2013) 54:1602 Page 5 of 18
123
one more special behavior that needs to be pointed out. In
previous experiments and simulations (Blake and Gibson
1981,1987; Chahine 1977; Pearson et al. 2004b), thinner
reentrant jet was observed with smaller standoff distance.
Nevertheless, in the present experiments, there are not only
changes in jet width but also a coexistence of two jets with
proper standoff distances, as demonstrated in Fig. 4in an
overall view with c
f
.
The first jet is formed near the end of growth, originated,
and migrates away from the liquid veneer above the bubble
(frame 1); this narrow jet is referred to as ‘‘the thin jet’’ to
avoid confusion. It penetrates the bubble in early collapse
phase, producing a protrusion, i.e., a jet torus similar to that
in Fig. 2, frame 5. The upper side of the bubble continues
to involute during bubble collapse and turns into a ‘‘con-
ventional’’ reentrant jet in frame 5. This jet follows and
maintains coaxial with the thin jet through the bubble.
Their coexistence is referred to as the ‘‘double-jet’’ sce-
nario. The toroidal bubble formed as the reentrant (second)
jet reaches the bubble bottom, and another jet torus is
formed (frame 7 and onward).
The average width of the thin jet is 0.86 mm, about 7 %
of the maximum bubble radius and 26 % of the reentrant jet
width. The maximum jet velocities are about 9.6 and
15.4 m/s. It is found in further experiments that as c
f
increases, the formation of the thin jet is delayed while the
reentrant jet becomes advanced in time. A schematic dia-
gram is given in Fig. 3b to demonstrate the thin jet for-
mation and the double-jet scenario in Fig. 4.
More details of the thin jet formation are provided in
Fig. 5. In the first frame, a flimsy liquid veneer is left between
the bubble and the free surface. Next, the veneer develops a
conelike shape with liquid accumulating at its top which is
subsequently elongated (frame 2).In the meantime, the ‘‘thin
jet’’ is initiated from the top (see frames 3–6 (magnified)).
The downward migration of the thin jet is coupled with the
rising of the top (frame 4 and onward).
3.1.4 Free surface motion
A liquid jet migrating upward was found when the free
surface is disturbed by rapid bubble expansion. There
might be obvious distortions on the lateral part of such
liquid jets as well when c
f
changes, contributing to various
water surface morphologies. The free surface dynamics
affected by small-scale (neglecting buoyancy) oscillating
bubble have been reported experimentally by Gibson
(1968), Chahine (1977), Blake and Gibson (1981), etc.,
featuring a main water jet from the surface, and also by
Dadvand et al. (2009) where ‘‘lateral jets’’ (deformation on
lateral parts of the main jet) were also noticed. Neverthe-
less, a more comprehensive study on the configuration of
the main or lateral jets is still preferable since in most
works the number of cases carried out was limited in
addition to the difficulty in enclosing both the bubble and
the entire spike in the images.
In this section, we first try to explain the key free surface
motion features with bubble behaviors, present images with
fair details, and then give a collection of the distinctive
morphologies of the free surface with different c
f
. To begin
with, a typical example with c
f
=0.80 is demonstrated in
Fig. 6. For convenience, the main part of the upward sur-
face jet is referred to as the ‘‘water spike’’ and its deformed
lateral parts as the ‘‘water skirt’’, as marked out in Fig. 6
(continued), frame 11. According to repeated observations,
we suggest that the spike is motivated by the expansion of
the bubble, while the skirt is induced mainly by the
‘rebound’’ of the bubble, which will be detailed below.
The form of the free surface motion is directly affected by
the standoff distance c
f
.
In Fig. 6, the free surface rises as a dome with bubble
expansion and becomes conical with its lateral part falling;
as the bubble collapses, the top rises pertinaciously prob-
ably due to the high water pressure at the rear of the
reentrant jet, as simulated by Blake and Gibson (1987), Li
et al. (2012), and Robinson et al. (2001); the water dome
then elongates into a spike. Right following the instability
stage (introduced in Sect. 3.1.2) between frames 4 and 5, a
minor ‘‘fold’’ appears around the spike (frame 6).
As the toroidal bubble rebounds, the liquid motion is
inverted from contracting toward the bubble center to
expanding away from it. This expansion of liquid becomes
visible at the free surface, resulting in the up-lifting of a
relatively large liquid bulk underneath the fold (frame 8
and onward). This portion of liquid collides with the liquid
above it and produces a water ‘‘skirt’’ at the position where
Fig. 5 Details of thin jet formation, c
f
=0.50. Frames 3–6 are magnified
Page 6 of 18 Exp Fluids (2013) 54:1602
123
Fig. 6 a Evolution of the water spike and the initial water skirt along with bubble oscillation, c
f
=0.80; b(continued) evolution of the water
spike and the integrated water skirt after bubble collapse, c
f
=0.80
Exp Fluids (2013) 54:1602 Page 7 of 18
123
the fold was (frame 9 and onward). In addition, the ‘‘fold’
was possibly caused similarly by the slight re-expansion of
the bubble during the instability stage.
Frames 9–17 capture the formation and development of
the skirt; with the bubble diameter (2R
eq
) and maximum
skirt speed being the characteristic length and speed,
respectively, We &4.01 910
2
and Re &2.6 910
4
. The
‘skirt’’ remains integrated as it evolves. In this process, the
spike out-speeds the skirt (We &4.50 910
3
,
Re &8.4 910
4
around 6–8 ms with the spike speed being
the characteristic speed). Droplets pinch off from the top
due to surface tension at the end of the inertial upward
migration (frame 13 and onward. With the droplet diameter
being characteristic length, Oh &2.13 910
-3
and
We &2.42 910
2
). With an upward motion, the skirt
(especially the top part) radially converges toward the
spike (frames 14–16).
Correspondence between bubble oscillation and free
surface motion is reflected in Fig. 7. A sharp increase
(peaking *6.9 m/s) in spike speed is found in alignment
with high-speed bubble expansion; the first decay in spike
speed (0.7–2.0 ms) can be accounted for by the decelera-
tion in bubble-wall speed in late expansion phase. The
spike speed passes an inflection point at *2 ms and rises
to a second peak (*3.5 m/s), which is simultaneous with
the bubble jet and hence probably accelerated by the high
pressure between the spike and the rear of the jet as
mentioned before.
Acceleration of the ‘‘fold’’ (to *1.1 m/s) is ensued from
the toroidal bubble rebound; the ‘‘fold’’ then bursts into a
water skirt, while the decay in spike speed is not disturbed
since its top is adequately far from the bubble.
More types of free surface motion are shown in Figs. 8,
9,10,11,12,withc
f
ranging from 0.31 to 1.60. In Fig. 8,
the effect of the rebound on the liquid surface is magnified
due to a smaller standoff distance c
f
=0.54. Consequently,
the water skirt acquires a higher speed of *2.6 m/s and
sprays around into droplets (frames 7–10) rather than stays
integrated, with We =2.24 910
3
, one order of magnitude
larger than the last case, and Re =6.24 910
4
.Oh for
these droplets (generally 0.5–1 mm in diameter) reaches
3.0 910
-3
–5.0 910
-3
. Nevertheless, the lower part of
the skirt top still converges toward the spike. In addition,
the double-jet scenario occurs in this case.
c
f
is reduced to 0.31 in Fig. 9. The bubble bursts into
open air; a screen of liquid is catapulted above the surface
and soon converges radially to form a very unstable water
spike, which is in rapid rise up to 16 m/s (We reaches
8.49 910
4
). No bubble rebound is observed; a possible
reason is that the pressure inside the bubble is presumably
normalized to atmospheric pressure when channeled to
open air. Consequently, there is no water skirt formed later
on.
With farther distance, the water spike is less pro-
nounced (see Fig. 10 where c
f
=1.12), acquiring lower
maximum speed (*1.0 and *0.31 m/s at two peaks) and
height. On the contrary, the effect of the rebounding is
less attenuated, and the skirt obtains a faster speed and
‘swallows’’ the spike (frames 4–6, first row); the top of
the spike then turns into a pit (see frames 1–2, second
Fig. 7 Speed variations in the
spike and skirt with time. The
oscillation stages of the bubble
are also marked along the time
axis, and its correspondence
with the speed curves is
analyzed in the text. The spike
speed is measured at the top
point of the spike or at the upper
edge of the first droplet (if
present) pinched off from the
spike. The skirt speed is
measured at the highest point of
the integrated skirt
Page 8 of 18 Exp Fluids (2013) 54:1602
123
row). The skirt merges over the pit, entrapping a bubble in
the liquid. In the meantime, a microliquid jet is catapulted
upward (frames 3–5, second row); then droplets pinch off
from the top of the jet (with Oh &6.10 910
-3
, frames
6–8, second row) with maximum speed of 4.5 m/s. Later,
the tip of the microjet inflates into a small dome (frames
8–9, second row). The high-speed microjet may be
explained as the consequence of the high-pressure stag-
nation resulting from the radial focusing of the bulk flow
during the merge of the skirt over the pit (Deng et al.
2007). The microjet formation is depicted schematically
in Fig. 11.
In this case, lower We is retained as 1.65 910
2
/
2.55 910
2
for the spike (at second speed peak)/skirt (at
maximum speed), while Re is still of order 10
4
. It is indi-
cated that the spike and the skirt migrate independently and
their relative heights are quite different from that in other
cases; this adds to the conjecture that they are generated by
different sources (i.e., by the bubble expansion and the
‘rebound’’, respectively).
When c
f
grows even larger to 1.60, neither the spike nor
the skirt is pronounced (Fig. 12). Though still generated
separately if observed very carefully (and the skirt is
dominating in the early stage), they rise as a whole at
*0.54 m/s maximum, in a cone shape where surface ten-
sion plays a more important role (We &96.7).
A summary is given in Fig. 13, where the maximum spike
heights rise exponentially with decreasing c
f
and reach more
than 200 times of R
eq
above the static surface. Five regions are
identified for the five typical free surface morphologies.
3.2 Single boundary: rigid bottom
The bubble behaviors (neglecting buoyancy effect) near
rigid boundaries with different c
b
have been comprehen-
sively studied (see, e.g., Benjamin and Ellis (1966), Blake
Fig. 8 Evolution of the water spike and the spraying skirt, c
f
=0.54
Fig. 9 Evolution of the bursting bubble and the unstable water spike, c
f
=0.31
Exp Fluids (2013) 54:1602 Page 9 of 18
123
and Gibson (1987), Shima et al.(1981), Tomita and Shima
(1986), Lindau and Lauterborn (2003), etc., and the refer-
ences therein). The late collapse phase of the bubble with
suitable c
b
(e.g., *1.0) is usually characterized by the two
toroidal fractions, i.e., the toroidal bubble and the jet torus,
similar to the free surface cases in Sect. 3.1, as captured in
Fig. 4a in Brujan et al. (2002) and Fig. 4a in Tomita and
Shima (1986), etc. Shock wave emissions are recorded
from the rebound of both fractions in Fig. 4a in Brujan
et al. (2002). However, in general, a jet induced by a rigid
wall forms at a later stage during bubble collapse than a jet
induced by a free surface and acquires a higher speed. In
addition, the formation of a jet torus is not necessary. In
some experiments, e.g., those by Lauterborn and Bolle
(1975) and Philipp and Lauterborn (1998), the bubble
contracts to minimum as a sole torus probably due to a
higher vapor condensation rate.
With a smaller c
b
(e.g., \0.6), the liquid between the
bubble and the rigid boundary was completely evacuated;
hence, the jet impacts directly onto the boundary rather
than to form a jet torus. Afterward, a toroidal bubble is
formed in close contact with the rigid boundary, collapses
and rebounds against it, and also radially expands along the
rigid surface and breaks into small segments. Typical cases
were presented in Fig. 2g–i by Philipp and Lauterborn
(1998), Fig. 14 by Lindau and Lauterborn (2003), and
Fig. 4d by Tomita and Shima (1986). In the subsequent
sections, the effect of the above bubble behaviors on the
free surface will be discussed.
3.3 Two boundaries: the insertion of a rigid bottom
The following two-boundary experiments are carried out
on the joint dynamics of the bubble, the rigid bottom, and
the free surface. To investigate its effect, the rigid bottom
wall is inserted at different c
b
(ranging from 0.18 to 4.80),
while c
f
remains constant. The image series of three typical
cases are listed in Figs. 14,15 and 16.
Figure 14 shows the comparative case without a rigid
bottom. The maximum speed and height of the spike are
6.8 m/s and 183 mm (16.6 times of R
eq
), respectively; the
skirt stays integrated with maximum height and speed being
54 mm and 1.2 m/s, respectively (We &4.77 910
2
,
Re &2.88 910
4
); the thin jet as in Sect. 3.1.3 is not
produced.
The case with c
b
being 0.97 is presented in Fig. 15. The
jet torus is blocked by the wall from further migration.
Fig. 10 Evolution of the lower water spike ‘‘swallowed’’ by the skirt, c
f
=1.10; the swallowing, bubble entrapment, and the microjet are shown
in the magnified images in the second row
D
C
B
F
E
A
Fig. 11 Schematic diagram for the microjet; profiles AFcorrespond
to frames 1–5 and 9 in Fig. 10, respectively
Page 10 of 18 Exp Fluids (2013) 54:1602
123
Fig. 12 Evolution of the lower water spike and skirt, c
f
=1.60
Bursting
bubble
&
unstable
high spike
Higher
spike
&
spraying
skirt
Higher
spike
&
integrated
skirt
Lower
spike
swallowed
by
skirt
Lower
spike
&
skirt
no buttom wall
Fig. 13 Variations in the non-
dimensional maximum spike
height with c
f
, no rigid bottom.
The maximum spike height is
measured as the vertical
distance from the highest
position of the spike top or the
first droplet pinched off from
the spike to the static water
surface. The non-dimensional
height H is the ratio between the
maximum spike height and the
maximum bubble radius. The
horizontal axis can be divided
into five regions according to
five different combinations of
water spikes and skirts. A
typical snapshot of each is given
in the corresponding region
Fig. 14 Evolution of the water spike and skirt along with bubble oscillation, without bottom wall, c
f
=0.89
Exp Fluids (2013) 54:1602 Page 11 of 18
123
Different from the previous case, a thin jet is produced
(frame 4) because the upper bubble wall is pressed closer to
the free surface by the rigid bottom during expansion,
although the initial bubble center is not changed. This also
caused the water spike to acquire a 31 % higher maximum
speed, 8.9 m/s, and a 164 % higher height, 484 mm, i.e.,
44 times of R
eq
, when compared to the non-bottom case,
with droplets pinching off with Oh between 2.0 910
-3
and 3.7 910
-3
. On the other hand, with We =5.60 910
2
and Re =3.12 910
4
, the form and height (48 mm) of the
water skirt are not much altered, probably because the
rebound of the toroidal bubble (frame 7) is not strength-
ened by the presence of the rigid wall and the effect of the
jet torus rebounding against the bottom is weak and late in
time.
The bottom is lifted to c
f
=0.23 in Fig. 16. As a result,
the bubble expands in direct contact with the bottom wall
into a hemisphere (frames 2–3); its upper wall is pressed
even closer to the free surface. Predictably, a thin jet is
generated (frame 3); following the thin jet, the upper
bubble wall involutes from above (frames 4–5) to produce
a reentrant jet which is induced by the joint effect of both
the free surface and the rigid bottom since the same
occurred with the presence of either boundary.
In frame 6 where the toroidal bubble rebounds against
the wall, dark spots are observed in surrounding fluid
which, as explained in Sect. 3.1.2, are microbubbles split
from the main bubble. The microbubbles are exited into
abrupt pulsation (subgraph B), and the copper remnants are
catapulted outward with their tracks captured as the radial
dark stripes (subgraph A); hence, there is possibly, fol-
lowing the same inference in Sect. 3.1.2, a shock wave
emitted by the rebound.
It is seen from Fig. 17 that with c
b
=0.23 the spike/
skirt acquires a maximum speed of 9.8/2.9 m/s and a height
of 608/91 mm (55/8.2 times of R
eq
). The insertion of the
rigid wall elevated the maximum spike speed by *41 %,
accompanied with a thin jet. Again, the second rise in spike
speed coincides with the jet, possibly due to the high liquid
pressure generated at jet rear. The skirt, initiated as bubble
rebound against the wall, has a 140 % increase in maxi-
mum speed compared to the non-bottom case. This could
be attributed to that more gas is retained in the toroidal
bubble rather than injected into a jet torus, contributing to a
stronger gas compression in the collapse. Thus, a stronger
radial liquid expansion is triggered, producing a more
vigorous water skirt at the free surface which sprays its
fringes into water drops (frame 9 and onward) with Oh
ranging from 2.5 910
-3
to 5.5 910
-3
. For the skirt, We
is elevated to 2.78 910
3
. The increase in spike height, on
the other hand, is more likely to be caused by that the
bubble is pushed closer to the free surface (equivalent to a
decrease in c
f
), since the toroidal bubble is less effective to
the spike top as concluded in Sect. 3.1.4.
Measurements from more cases are presented below. It
is displayed in Fig. 18 that with a fixed c
f
, the maximum
Fig. 15 Evolution of the water spike and skirt along with bubble oscillation with both free surface and rigid bottom; c
f
=0.89, c
b
=0.97
Page 12 of 18 Exp Fluids (2013) 54:1602
123
Fig. 16 Evolution of the water spike and skirt along with bubble oscillation, with both the free surface and the rigid bottom; c
f
=0.89,
c
b
=0.23
Fig. 17 Speed variations in the
spike and skirt with time. The
oscillation stages of the bubble
are marked along the time axis
Exp Fluids (2013) 54:1602 Page 13 of 18
123
height of the water spike rises sharply as c
b
falls below 0.5.
The trend is very similar to that in Fig. 13, and hence, we
suggest that both are results of the increased proximity
between the upper bubble wall and the free surface.
Figure 19 demonstrates the change in bubble jet with
different c
b
. Pushed nearer to the free surface by small c
b
,
the bubble is enabled to produce a thin jet. Generally, for c
b
above 1.2, only reentrant jet is produced; for c
b
under 0.4,
only thin jet is observed since there is no enough space for
the reentrant jet to develop. Between 0.4 and 1.2 there exist
the double-jet scenarios, in which the reentrant jet is
always faster. Both kinds of jets gain in speed as c
b
decreases.
In conclusion, the insertion of a rigid wall in the prox-
imity of an oscillating bubble near free surface will result
in (1) a smaller distance between the upper bubble wall and
the free surface which may cause a thin jet and a higher
water spike and (2) rebound of the toroidal bubble against
the surface of the rigid wall which possibly causes a more
vigorous water skirt.
Displacement curves of the top and the bottom points
of the bubble from its initial center are plotted in Fig. 20;
the top bubble boundary will be partially blocked by a
dark strip on the image caused by meniscus effect when
surpassing the static water surface. To overcome this
problem, a 2-degree spline is fitted to the visible part of
the upper bubble boundary (see Fig. 14, frames 2–3) to
represent the blocked part. Fortunately, this meniscus
problem occurs merely in a few images near bubble
maximum, and the displacement of the top point mea-
sured on the spline fits well to the whole displacement
curve. Back to Fig. 20, with the rigid wall being the only
boundary, the bubble jet is developed at a relatively late
stage and appears in the figure as the sharp downfall of
the bubble’s top point between 3.5 and 4.0 ms. On the
other hand, for the free surface-only case, the reentrant jet
is produced earlier and possesses a lower speed. Further,
when both boundaries are present, the thin jet appears in
the late expansion phase, earlier but slower than the
reentrant jet, as reflected by the lower steepness of the
curve. In addition, the maximum displacement of bubble
top is slightly larger, as the top part is elongated at the
free surface; the bubble bottom is pressed when the rigid
wall is inserted.
3.4 Two boundaries: the insertion of the free surface
The effect of the free surface is investigated with different
heights (c
f
) above the bubble initial center, while the
bubble-wall distance (c
b
) is kept constant. Two groups of
experiments are conducted with c
b
being 0.97 and 0.10,
respectively. For each group, the snapshots of the selected
cases at the same moment or stage are displayed together
for comparison.
In the first group, c
f
ranges from 1.75 to 0.20 and six
cases are chosen for display. Images in the first row of
Fig. 21 are captured at maximum bubble radius. The free
surface becomes more disturbed with smaller c
f
. Thin jets
are observed with c
f
=0.37 and 0.58, which are initiated at
approximately 1.9 and 1.3 ms, respectively, earlier with
lower free surface (see case (e) and (f)).
The second row is captured at 5.00 ±0.08 ms. Bubble
oscillation is more advanced with smaller c
f
. In the first two
cases with large c
f
, the upper bubble walls are flattened but
have not yet turned into reentrant jets, while in case (c) and
(d) the reentrant jets are penetrating the bubble. The bub-
bles in the last two cases are at later collapse stages, where
the jet tori are expanding along the rigid wall. The spike
acquires a greater height as c
f
decreases.
Fig. 18 Variations in non-dimensional maximum spike height with
c
b
when c
f
=0.89. Non-dimensional height H is the ratio between the
maximum spike height and R
eq
Fig. 19 Variations in jet speed with c
b
when c
f
=0.89. For double-
jet scenarios, the coexisting thin jet and reentrant jet of the same
bubble are linked by the dot line
Page 14 of 18 Exp Fluids (2013) 54:1602
123
In the third row, the free surface motion is fully devel-
oped. For (a) and (b), the motions are not pronounced. In
case (c), the fringe of the skirt is smooth and integrated,
while it starts to revolute outward in case (d) and sprays
radially into droplets in case (e); finally, the skirt is dis-
torted in case (f).
Fig. 20 The displacement
curves of the top and the bottom
point of the bubble in different
boundary conditions. The free
surface is inserted at c
f
=0.89;
the rigid bottom is inserted at
c
b
=0.97
Fig. 21 Bubble and free surface motion with different c
f
while c
b
=0.97. First row captured at maximum bubble radius. Second row captured at
5.00 ±0.08 ms. Third row captured at 13.33 ±0.08 ms. c
f
in the six cases are a1.69, b1.35, c0.97, d0.81, e0.58, and f0.37
Exp Fluids (2013) 54:1602 Page 15 of 18
123
For a more comprehensive investigation, a smaller
c
b
(0.10) is assigned to the second group. The c
f
ranges from
2.03 to 1.06 in the cases captured in Fig. 22.
The maximum bubbles are shown in the first row of
Fig. 22, while the second row is captured right before the
water skirts become visible. With falling c
f
, the spike is
initiated with higher speed. Next, the skirts emerge in the
third row. In case (b) and (c) where the spikes are less
pronounced due to larger c
f
(1.65 and 1.50), the skirts
overpass the spikes, similar to the case in Fig. 10. The
shape to which a skirt evolves into is less integrated as c
f
reduces.
There are two factors contributing to the intensified free
surface motion, i.e., the proximity of the upper bubble wall
to the free surface during expansion, which is also attrib-
uted to the pressing from the rigid wall, and the outward
fluid motion at bubble rebound. As has been discussed in
Sect. 3.1.4, the spike height is not likely to be affected by
the jet torus; hence, the elevation of spike heights (shown
later in Fig. 23) is mainly a result of smaller c
f
. As for the
water skirts, vigorous ones are observed in the non-bottom
cases with small c
f
; besides, it is indicated by both the
images and the speed curves that the water skirt formation
is closely linked to the rebound of the toroidal bubble
Fig. 22 Bubble and free
surface motion with different c
f
when c
b
=0.10, captured at
maximum bubble radius (first
row) and right before water skirt
formation (second row)
Page 16 of 18 Exp Fluids (2013) 54:1602
123
(either the one with a jet torus or the one clinging to the
wall); therefore, the intensified skirts should be attributed
to both factors mentioned above.
In the third row of Fig. 22, the concealed spikes in
(b) and (c) re-emerge above the skirts. Again, a microjet is
catapulted upward in case (c), with a speed up to 7.5 m/s;
only this time the gas entrapment is missing.
Figure 23 demonstrates the maximum spike height
variation with c
f
, with the presence of the rigid wall. Three
groups of experiments are conducted where c
b
is kept at
0.10, 0.63, and [10, respectively, and c
f
falls from 2.0 to
0.3. The maximum height of the spike always decreases
exponentially with growing c
f
. Compared to the group
where c
b
[10, the maximum heights at different c
f
are
elevated with c
b
=0.63 and are elevated again when c
b
is
reduced to 0.10.
4 Conclusions
The dynamics of the bubble and the free surface have been
experimentally studied using high-speed photography. The
bubbles are generated by underwater electric discharge and
pulsate in the vicinity of the free surface and/or a hori-
zontal rigid boundary, with varying bubble-boundary dis-
tances (c
b
and c
f
). Intriguing motion features have been
found with both single and double boundaries. The current
results benefit from higher temporal and spatial resolution
allowed by the relatively large scale (*10 mm for radius,
*5 ms for first period) of the spark-generated bubbles.
The bubble splits into two toroidal bubbles due to the
reentrant jet in late collapse phase near the free surface as
well as at certain distances from the rigid boundary.
Besides, with proper c
f
, the bubble was found to produce
two successive jets in the first collapse phase. According to
configurations of the water spike and skirt, the free surface
motions are categorized into five types. The maximum
height of the spike rises exponentially with decreasing c
f
.
There are two peaks in the upward speed of the spike; the
first is induced by the initial growth of the bubble, while
the second is probably induced by the high fluid pressure at
the rear of the bubble jet. Based on the observations of the
transient pulsation of microbubbles suspending around the
main bubble, shock wave emissions were inferred at the
rebound of the toroidal bubble or the jet torus.
For a constant bubble-free surface distance (c
f
), the
decrease in bubble-wall distance (c
b
) leads to elevations of
the speed and maximum height of the water spike and skirt,
increase in reentrant jet speed, and formation of the thin jet.
For a constant bubble-rigid bottom distance (c
b
) and a
descending free surface, the bubble is advanced in oscil-
lation stages, and an exponential rise in maximum spike
height was found.
Most of the motion features observed in the double-
boundary cases are inherited from the single-boundary
cases but change in speed, height, etc. Therefore, additional
considerations are required in bubble applications with
multiple boundaries, especially those of different natures.
Acknowledgments This work was supported by the project of the
Outstanding Youth Fund of China (Grant No. 51222904) and the
National Security Major Fundamental Research Program of China
(Grant No. 613157). The authors are also grateful for the precious
advice and support from Dr. Q. X. Wang from University of
Birmingham.
References
Best JP (1993) The formation of toroidal bubbles upon the collapse of
transient cavities. J Fluid Mech 251:79–107
Benjamin TB, Ellis AT (1966) The collapse of cavitation bubbles and
the pressures thereby produced against solid boundaries. Phil
Trans R Soc Lond A 260:221–240
Blake JR, Gibson DC (1981) Growth and collapse of a vapour cavity
near a free surface. J Fluid Mech 111:123–140
Blake JR, Gibson DC (1987) Cavitation bubbles near boundaries.
Annu Rev Fluid Mech 19:99–123
Blake JR, Taib BB, Doherty G (1987) Transient cavities near
boundaries. Part 2 Free surface. J Fluid Mech 181:197–212
Blake JR, Hooton MC, Robinson PB, Tong RP (1997) Collapsing
cavities, toroidal bubbles and jet impact. Phil Trans R Soc Lond
A 355:537–550
Blake JR, Tomita Y, Tong RP (1998) The art, craft and science of
modelling jet impact in a collapsing cavitation bubble. In:
Biesheuvel A, Heijst G (eds) In fascination of fluid dynamics.
Springer, Netherlands, pp 77–90
Blake JR, Keen GS, Tong RP, Wilson M (1999) Acoustic cavitation:
the fluid dynamics of non–spherical bubbles. Phil Trans R Soc
Lond A 357:251–267
Bourne NK, Field JE (1995) A high speed photographic study of
cavitation damage. J Appl Phys 78:4423–4427
Fig. 23 Variations in maximum spike heights with c
f
when c
b
is kept
constant at 0.10, 0.63, and [10, respectively. The non-dimensional
maximum spike height H is the ratio between the maximum spike
height and R
eq
Exp Fluids (2013) 54:1602 Page 17 of 18
123
Brennen CE (2002) Fission of collapsing cavitation bubbles. J Fluid
Mech 472:153–166
Brujan EA, Nahen K, Schmidt P, Vogel A (2001) Dynamics of laser-
induced cavitation bubbles near an elastic boundary. J Fluid
Mech 433:251–281
Brujan EA, Keen GS, Vogel A, Blake JR (2002) The final stage of the
collapse of a cavitation bubble close to a rigid boundary. Phys
Fluids 14:85–92
Chahine GL (1977) Interaction between an oscillating bubble and a
free surface. J Fluids Eng 99:709–716
Dadvand A, Khoo BC, Shervani-Tabar MT (2009) A collapsing
bubble-induced microinjector: an experimental study. Exp Fluids
46:419–434
Deng Q, Anilkumar AV, Wang TG (2007) The role of viscosity and
surface tension in bubble entrapment during drop impact onto a
deep liquid pool. J Fluid Mech 578:119–138
Fong SW, Adhikari D, Klaseboer E, Khoo BC (2009) Interactions of
multiple spark-generated bubbles with phase differences. Exp
Fluids 46:705–724
Fujikawa S, Akamatsu T (1980) Effects of the non-equilibrium
condensation of vapour on the pressure wave produced by the
collapse of a bubble in a liquid. J Fluid Mech 97:481–512
Gibson DC (1968) Cavitation adjacent to plane boundaries 3rd
Australasian Conference on Hydraulic fluid mechanics
Gilmore FR (1952) The growth or collapse of a spherical bubble in a
viscous compressible liquid. Cal Tech Inst Rep 26–4. Hydrody-
namics Laboratory, California Institute of Technology, Pasa-
dena, CA. http://authors.library.caltech.edu/561/
Lauterborn W, Bolle H (1975) Experimental investigations of
cavitation-bubble collapse in the neighbourhood of a solid
boundary. J Fluid Mech 72:391–399
Li Z, Sun L, Zong Z, Dong J (2012) Some dynamical characteristics
of a non-spherical bubble in proximity to a free surface. Acta
Mech 223:2331–2355
Lindau O, Lauterborn W (2003) Cinematographic observation of the
collapse and rebound of a laser-produced cavitation bubble near
a wall. J Fluid Mech 479:327–348
Longuet-Higgins MS (1983) Bubbles, breaking waves and hyperbolic
jets at a free surface. J Fluid Mech 127:103–121
Menon S, Lal M (1998) On the dynamics and instability of bubbles
formed during underwater explosions. Exp Therm Fluid Sci
16:305–321
Naude
´CF, Ellis AT (1961) On the mechanism of cavitation damage
by nonhemispherical cavities collapsing in contact with a solid
boundary. J Fluids Eng 83:648–656
Ohl CD, Ikink R (2003) Shock-wave-induced jetting of micron-size
bubbles. Phys Rev Lett 90:214502
Ohl CD, Philipp A, Lauterborn W (1995) Cavitation bubble collapse
studied at 20 million frames per second. Ann Physik 4:26–34
Pearson A, Blake JR, Otto SR (2004a) Jets in bubbles. J Eng Math
48:391–412
Pearson A, Cox E, Blake JR, Otto SR (2004b) Bubble interactions
near a free surface. Eng Anal Bound Elem 28:295–313
Philipp A, Lauterborn W (1998) Cavitation erosion by single laser-
produced bubbles. J Fluid Mech 361:75–116
Prosperetti A, Lezzi A (1986) Bubble dynamics in a compressible
liquid. Part 1. First-order theory. J Fluid Mech 168:457–478
Rayleigh L (1917) On the pressure developed in a liquid during the
collapse of a spherical cavity. Philos Mag 34:94–98
Robinson PB, Blake JR, Kodama T, Shima A, Tomita Y (2001)
Interaction of cavitation bubbles with a free surface. J Appl Phys
89:8225–8237
Shaw SJ, Jin YH, Schiffers WP, Emmony DC (1996) The interaction
of a laser-generated cavity in water with a solid surface. J Acoust
Soc Am 99:2811–2824
Shaw SJ, Jin YH, Gentry TP, Emmony DC (1999) Experimental
observations of the interaction of a laser generated cavitation
bubble with a flexible membrane. Phys Fluids 11:2437–2439
Shima A, Takayama K, Tomita Y, Miura N (1981) An experimental
study on effects of a solid wall on the motion of bubbles and
shock waves in bubble collapse. Acta Acust Acust 48:293–301
Shima A, Takayama K, Tomita Y (1983) Mechanism of impact
pressure generation from spark-generated bubble collapse near a
wall. AAIA J 21:55–59
Tomita Y, Shima A (1986) Mechanisms of impulsive pressure
generation and damage pit formation by bubble collapse. J Fluid
Mech 169:535–564
Tong RP, Schiffers WP, Shaw SJ, Blake JR, Emmony DC (1999) The
role of ‘splashing’ in the collapse of a laser-generated cavity near
a rigid boundary. J Fluid Mech 380:339–361
Turangan CK, Ong GP, Klaseboer E, Khoo BC (2006) Experimental
and numerical study of transient bubble-elastic membrane
interaction. J Appl Phys 100:054910-1–054910-7
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 206:299–338
Wang QX, Yeo KS, Khoo BC, Lam KY (1996) Nonlinear interaction
between gas bubble and free surface. Comput Fluids 25:607–628
Wang QX, Yeo KS, Khoo BC, Lam KY (2005) Vortex ring modelling
of toroidal bubbles. Theor Comp Fluid Dyn 19:303–317
Ward B, Emmony DC (1991) Interferometric studies of the pressure
developed in a liquid during infrared-laser-induced cavitation
bubble oscillation. Infrared Phys 32:489–515
Zhang S, Duncan JH, Chahine GL (1993) The final stage of the collapse
of a cavitation bubble near a rigid wall. J Fluid Mech 257:147–181
Zhang AM, Wang SP, Bai ZH (2011) Experimental study on bubble
pulse features under different circumstances. Chin J Theor Appl
Mech 43:71–83
Page 18 of 18 Exp Fluids (2013) 54:1602
123
... The physical properties of the boundary are crucial for the collapse of the bubble, especially the asymmetric deformation. [6] For the study of bubble near a single boundary, there are mainly near free surface, elastic boundary and rigid boundary. When bubble grows at a certain distance from the free surface, the free surface is pushed upward as the bubble expands, and a substantial water dome is formed on the free surface, which then evolves into a spike 2 after the bubble shrinks. ...
... (www.preprints.org) | NOT PEER-REVIEWED | Posted: 5 February 2024 doi:10.20944/preprints202402.0239.v16 ...
Preprint
Full-text available
Cavitation bubbles are commonly existing in shipbuilding engineering, ocean engineering, mechanical engineering, chemical industry, and aerospace. Asymmetric deformation of bubble occurs near the boundary, and the non-spherical cavitation bubbles have strong destructive effects, inducing high amplitude loading at the nearby boundary. The purpose of this paper is to investigate the laser-induced of cavitation bubble, non-spherical collapse behavior near the boundary. In this study, experimental data such as bubble pulsation process and bubble surface velocity distribution were obtained by high-speed camera technique and full-field velocity calculation. The results show that the bubble appeared non-spherical collapse shapes, near different boundaries, with near-hemispherical, near-ellipsoidal, near-cone and near-pea shapes, and the bubble surface velocity distribution is not uniform. When bubble near the free surface or rigid boundary, the smaller the stand-off r is, the more obvious the repulsive effect of the free surface and the attractive effect of the rigid boundary are. As the stand-off r decreases, the larger the Bjerknes force and the bubble surface velocity difference, the more pronounced the non-spherical shape.
... They found that the collapse time of the bubbles was shortened and prolonged by the water surface and rigid wall boundary, respectively. Zhang et al. [34] argued that most of the motion features observed in the double-boundary cases originate from the single-boundary cases, in addition to changes in the speed and height. Huang et al. [35] experimentally observed three distinct patterns identified from the morphologies of the bubbles and water surfaces. ...
Article
Full-text available
Wall vortex occurs when a cavitation bubble oscillates far from a single rigid wall (at a dimensionless standoff distance of γr>1.3). This study reveals that introducing a water surface expands the wall vortex regime. A wall vortex in an expanded new regime forms instead of a free vortex at a smaller γr value. Because of the influence of the water surface, a broader jet pierces the bottom of a bubble. This causes the bubbles to expand easily along the wall and form a flat shape during the second cycle. Here an outwards flow forms instead of an upward flow after the bubble recollapses. This study investigates the formation and development of a wall vortex in the new expanded regime via a combination of experiments, numerical simulations, and theoretical modeling. To this end, a theoretical model describing the radial motion R and centroid position h of the bubble between the boundaries is developed using Lagrangian formulation. Two infinite sets of image bubbles are used to satisfy the conditions of the water surface and rigid wall based on image theory. The criteria for the vortex flow patterns are proposed based on the direction of the centroid migration ḣ(tc) of the bubble at the beginning of the second cycle tc. A free vortex occurs when the upward flow dominates [ḣ(tc)>0], whereas a downwards flow dominates the wall vortex [ḣ(tc)<0]. A phase diagram of the vortex flows is obtained from the theoretical model and is verified using the experimental results. Numerical analysis reveals that the wall vortex flow with the influence of the water surface contributes to a greater wall shear stress and larger area, thereby increasing the potential for surface cleaning. These findings provide new insights for engineering applications such as ultrasonic cleaning.
... The asymmetric collapse of cavitation bubbles often occurs near the boundaries, and micro-jets are formed. The physical properties of the boundaries are crucial for the bubble, especially asymmetric deformation [6]. The study of bubbles near a single boundary comprises studies mainly of near the free surface, elastic boundary, and rigid boundary. ...
Article
Full-text available
Cavitation bubbles commonly exist in shipbuilding engineering, ocean engineering, mechanical engineering, chemical industry, and aerospace. Asymmetric deformation of the bubble occurs near the boundary and then has strong destructiveness, such as high amplitude loading. Therefore, the research on non-spherical deformation is of great significance, and the objective of this paper is to investigate the non-spherical collapse dynamics of laser-induced cavitation bubbles when near different boundaries. In this study, experimental data, such as the bubble pulsation process and bubble surface velocity distribution, were obtained by high-speed camera techniques and full-field velocity calculations. Near the different boundaries, the results show that the bubbles appeared to have different collapse shapes, such as near-hemispherical, near-ellipsoidal, near-cone, and near-pea shapes, and the surface velocity distribution is extremely non-uniform. When the bubble near the free surface or rigid boundary collapses, the smaller the stand-off r is, the more obvious the repulsive effect of the free surface or the attractive effect of the rigid boundary is. As the stand-off r decreases, the larger the Bjerknes force and the bubble surface velocity difference and the more pronounced the non-spherical shape becomes.
... When a cavitation bubble collapses between a rigid and a free surface, most of the motion features of the bubble and boundaries share some similarities with the single-boundary case. However, the main dynamics are changed, such as the jet speed and jet height [189] . The bubble also splits into two toroidal bubbles due to the same mechanism as that behind bubble collapses near a single free surface. ...
Article
Full-text available
Cavitation occurs widely in nature and engineering and is a complex problem with multiscale features in both time and space due to its associating violent oscillations. To understand the important but complicated phenomena and fluid mechanics behind cavitation, a great deal of effort has been invested in investigating the collapse of a single bubble near different boundaries. This review aims to cover recent developments in the collapse of single bubbles in the vicinity of complex boundaries, including single boundaries and two parallel boundaries, and open questions for future research are discussed. Microjets are the most prominent features of the non-spherical collapse of cavitation bubbles near boundaries and are directed toward rigid walls and away from free surfaces. Such a bubble generally splits, resulting in the formation of two axial jets directed opposite to each other under the constraints of an elastic boundary or two parallel boundaries. The liquid jet penetrates the bubble, impacts the boundary, and exerts a great deal of stress on any nearby boundary. This phenomenon can cause damage, such as the erosion of blades in hydraulic machinery, the rupture of human blood vessels, and underwater explosions, but can also be exploited for applications, such as needle-free injection, drug and gene delivery, surface cleaning, and printing. Many fascinating developments related to these topics are presented and summarized in this review. Finally, three directions are proposed that seem particularly fruitful for future research on the interaction of cavitation bubbles and boundaries.
... Since the pioneering establishment of the ideal spherical cavitation bubble model by Rayleigh [12], numerous scientists have conducted extended research on bubble dynamics, encompassing phenomena like microjets during bubble collapse. Some researchers have investigated the bubbles collapse near various boundaries through experimental methods and have made numerous discoveries, such as the evolution of jets near wall [13][14][15], the development of microjets under the influence of a free surface [16], mechanism of microjets in cavitation erosion [17,18], the formation mechanisms of microjets and counterjets [19], microjets under different gravity [20], the development of microjets affected by a wall and a free surface [21]. Another group of scientists, by combining numerical simulations with their research, have also achieved significant outcomes, such as the development of microjets near wall surfaces [22][23][24], the impact pressure characteristics of microjets on the wall [25], scale laws between different types of cavitation bubbles [26], unified studies on microjets [27], the characteristics of microjets near a wall in compressible mediums [28], unified studies on different classical bubble equations [29]. ...
Article
Full-text available
The bubble dynamics under the influence of particles is an unavoidable issue in many cavitation applications, with a fundamental aspect being the shockwave affected by particles during bubble collapse. In our experiments, the method of spark-induced bubbles was used, while a high-speed camera and a piezoresistive pressure sensor were utilized to investigate how particle shape affects the evolution of shockwaves. Through the high-speed photography, we found that the presence of the particle altered the consistency of the liquid medium around the bubble, which result in the emitting of water hammer shockwave and implosion shockwave respectively during the collapse of the bubble. This stratification effect was closely related to the bubble-particle relative distance φ and particle shape δ. Specifically, when the bubble-particle relative distance φ < 1.34 e-0.10δ, particles disrupted the medium consistency around the bubbles and led to a nonspherical collapse and the consequent stratification of the shockwave. By measuring the stratified shockwave intensity affected by different particle shapes, we found that the stratified shockwave intensity experienced varying degrees of attenuation. Furthermore, as the particle shape δ increased, the attenuation of the particle on shockwave intensity gradually reduced. These new findings hold significant theoretical implications for elucidating cavitation erosion mechanisms in liquid–solid two-phase flows and applications and prevention strategies in liquid–solid two-phase cavitation fields.
... The development of theoretical models is also an important part of studying the dynamic behavior of bubbles, and the current research on cavitation bubbles mainly focuses on the formation of bubbles, the influence of shock waves and boundary conditions on the collapse characteristics of bubbles [10][11][12]. Fujikawa and Akamastu [13] considered the compressibility, heat transfer, and non-equilibrium evaporation and condensation of vapors, and proposed an analytical model that showed that evaporation and condensation strongly affect the bubble dynamics. Han [14] experimentally, numerically and theoretically investigate the nonlinear interaction between a spark cavitation bubble and the interface of two immiscible fluids (oil and water) on multiple time scales, but she also only focused on the first cycle of bubble motion. ...
Article
Full-text available
The main composition within a spark-generated bubble primarily consists of vapor, accompanied by a minor presence of noncondensable gases. The phase transition exerts a substantial influence on bubble dynamics throughout various stages, a facet that has been frequently overlooked in prior research. In this study, we introduce a modified theoretical model aimed at accurately predicting the multiple oscillations of spark-generated bubbles. Leveraging the Plesset equation, which integrates second-order corrections for compressibility and non-equilibrium evaporation, we further incorporate the thermal boundary layer approximation for bubbles, as proposed by Zhong et al. We employ an adjusted phase transition duration tailored to the unique characteristics of spark-generated bubbles. Furthermore, we meticulously ascertain initial conditions through repeated gas content measurements within the bubble. Our proposed theoretical model undergoes rigorous validation through quantitative comparisons with experimental data, yielding commendable agreement in modeling the dynamic behavior of bubbles across multiple cycles. Remarkably, we uncover that the condensation rate significantly governs the behavior of spark bubbles during their initial two cycles. Finally, we investigate the dependence of spark-generated bubble dynamics on the phase transition and the presence of air. Air content exhibits a minimal impact on bubble motion prior to the initial bubble collapse, but plays a role in the bubble’s rebound thereafter.
... Zhang et al. [12] developed a theoretical model to predict the bubble dynamics near different structures. For cavitation bubble dynamics near complex structures (free surface and rigid wall), Zhang et al. [13] found that the bubble splits into two toroidal bubbles during the late collapse stage near the free surface, as well as at some distance from the rigid wall. Previously, for cavitation bubble dynamics near particles, our group [14] experimentally investigated the effects of spherical particles on the laser-induced cavitation bubble. ...
Article
Full-text available
The interaction between cavitation bubbles and particles is essential for the operational performance many kinds of fluid machineries. In the present paper, jet dynamics and shock waves induced by the cavitation bubble collapsing near two spherical particles are numerically investigated based on OpenFOAM. The numerical scheme is validated by the experimental data obtained based on our high-speed camera cavitation system. Our results reveal that bubble split induced by annular jet is the primary feature during bubble collapsing with four typical cases defined. For the jet formation, the localized high pressure produced at the bubble split point is the main reason and the split point also serves as the source of the shock waves shown by the numerical schlieren. Furthermore, the nondimensional bubble-particle distance is the most paramount parameter influencing the jet phenomenon (e.g. jet velocity).
Article
The influence of liquid viscoelasticity on the interaction between cavitation bubbles and free surfaces is of great practical significance in understanding bubble dynamics in biological systems. A series of millimeter cavitation bubbles were induced by laser near the free surfaces of the water and viscoelastic polyacrylamide (PAM) solutions with different concentrations. The effects of liquid viscoelasticity on the interactions of cavitation bubbles with free surfaces are analyzed from the perspectives of the evolution of free surface and bubble dynamics. The experimental results show that as the dimensionless standoff distance increases, the evolutions of free surface behaviors in all experimental fluids can be divided into six types of water mounds, i.e., breaking wrinkles, spraying water film, crown, swallowed water spike, hillock, and slight bulge. All the critical values of the dimensionless distance dividing different types decrease with increasing concentration. The evolutions of first four types of water mounds in PAM solutions differ from those in the water. Water droplets splashing in different directions are produced around the breaking wrinkles in the water. Meanwhile, the breaking wrinkles in PAM solution move with the “liquid filaments” towards the central axis. The water spike in the pattern of spraying water film in PAM solution is more stable than that in the water. As the solution concentration increases, the water skirt in the pattern of crown contracts earlier and faster, and the rate of increase in the height of the water skirt decreases. For swallowed water spike in PAM solution, the upper part of the newly formed water spike is not significantly thicker than the middle part, and thus the water waist structure does not form. Liquid viscoelasticity inhibits the bubble growth and collapse, and the bubble migration as well, especially in the second period. Shorter and thicker cavities are formed in PAM solutions with higher concentration, while slender and stable cavities formed in the water at the same dimensionless distance. The velocity and displacement of the tip of bullet jet both decrease as the solution concentration increases.
Article
Full-text available
The interaction of a laser-induced cavitation bubble with an elastic boundary and its dependence on the distance between bubble and boundary are investigated experimentally. The elastic boundary consists of a transparent polyacrylamide (PAA) gel with 80% water concentration with elastic modulus E = 0.25 MPa. At this E-value, the deformation and rebound of the boundary is very pronounced providing particularly interesting features of bubble dynamics. It is shown by means of high-speed photography with up to 5 million frames s−1 that bubble splitting, formation of liquid jets away from and towards the boundary, and jet-like ejection of the boundary material into the liquid are the main features of this interaction. The maximum liquid jet velocity measured was 960 m s−1. Such high-velocity jets penetrate the elastic boundary even through a water layer of 0.35 mm thickness. The jetting behaviour arises from the interaction between the counteracting forces induced by the rebound of the elastic boundary and the Bjerknes attraction force towards the boundary. General principles of the formation of annular and axial jets are discussed which allow the interpretation of the complex dynamics. The concept of the Kelvin impulse is examined with regard to bubble migration and jet formation. The results are discussed with respect to cavitation erosion, collateral damage in laser surgery, and cavitation-mediated enhancement of pulsed laser ablation of tissue.
Article
Full-text available
A limit value for the distance from the free surface to the center of the bubble reported to its radius is found. Under this limit the free surface is not disturbed before the end of the collapse in the first approximation. Only in this case, the method of images can be used and the free surface be replaced by an image-source, symmetrical with respect to the free surface, to the sink representing the bubble. Above this limit, precise measurements of bubble deformation and motion are given. Just after the collapse of the bubble begins, observations show a singular perturbation on the free surface, with the formation of a thin spike directed towards the air. In all cases, buoyancy has no time to take effect, and the bubble is repelled from the free surface while the re-entering jet, formed during collapse, is oriented away from it.
Article
Full-text available
The final stage of the collapse of a laser-produced cavitation bubble close to a rigid boundary is studied both experimentally and theoretically. The temporal evolution of the liquid jet developed during bubble collapse, shock wave emission and the behavior of the “splash” effect are investigated by using high-speed photography with up to 5 million frames/second. For a full understanding of the bubble–boundary interaction, numerical simulations are conducted by using a boundary integral method with an incompressible liquid impact model. The results of the numerical calculations provided the pressure contours and the velocity vectors in the liquid surrounding the bubble as well as the bubble profiles. The comparisons between experimental and numerical data are favorable with regard to both bubble shape history and translational motion of the bubble. The results are discussed with respect to the mechanism of cavitation erosion.
Article
Underwater explosion bubbles do severe damage to warship structures, and the particularity of underwater explosion and the quality of water bring great difficulties to get a clear picture. So lots of laboratories all around the world make use of discharge or laser to generate bubbles to replace real underwater explosion experiments. Based on the designation from Turangan et al and Abdolranman Dadvand, an electric circuit with a relatively low voltage 200 V and three shunt-wound 2200 μF capacitances is used to generate bubbles by discharge in a 500 mm × 500 mm × 500 mm water tank to study underwater bubble dynamics. The experiment setup is safe, stable, low in costs and simple to fix. It can generate bubbles with radius up to 12~15 mm. A device generating bubble by a relatively low voltage is designed to study bubble dynamics in detail under various boundary conditions in this paper. Main conclusions are obtained below by comparison and analysis: (1) A voltage of 200 V is adopted to generate bubble by discharge in water in this paper. The voltage is fairly low yet the bubble generated is relatively large in size. The pictures shot are pretty clear as well. (2) There are copper wires burning inside the bubble throughout the expansion process, causing the temperature inside bubble to rise apparently higher and result in the small calculated value of bubble pulsing period. So p v within the bubble is actually very large and the calculated p v of a single bubble pulsing in free field is 5.8 × 10 4 Pa. (3) Bubble's second pulsing energy is merely 5%~15% of the first pulsing energy and the pulsing energy after the second pulse can be ignored. (4) γ, the dimensionless standoff is defined between bubble and boundary. When γ ≥ 2.8, the influence of the wall to bubble can be neglected. When 1.95 ≤ γ ≤ 2.8, no jet is formed during the first cycle of bubble pulse. When γ ≤ 1.53, jet is formed during the first cycle of bubble pulse, and the smaller γ is, the larger bubble migration distance and the more fiercely the jet would be. (5) The bubble jetting velocity is minimum when the dimensionless standoff γ is around 0.8. (6) Bubble will interact strongly with boundaries when it moves near the shipboard and the free surface. It can be seen from the experiments that the bubble may produce jets that are towards two different directions when collapsing near the free surface and the shipboard, and this will certainly lead to the weakening effect of bubble attack to the shipboard.
Article
A perfect fluid theory, which neglects the effect of gravity, and which assumes that the pressure inside a cavitation bubble remains constant during the collapse process, is given for the case of a nonhemispherical, but axially symmetric cavity which collapses in contact with a solid boundary. The theory suggests the possibility that such a cavity may deform to the extent that its wall strikes the solid boundary before minimum cavity volume is reached. High speed motion pictures of cavities generated by spark methods are used to test the theory experimentally. Agreement between theory and experiment is good for the range of experimental cavities considered, and the phenomenon of the cavity wall striking the solid boundary does indeed occur. Studies of damage by cavities of this type on soft aluminum samples reveals that pressures caused by the cavity wall striking the boundary are higher than those resulting from a compression of gases inside the cavity, and are responsible for the damage.
Article
A detailed experimental study is made of behaviour of bubbles produced by a spark discharge in water. By using a high speed camera operated in framing and streaking modes, the motion of a collapsing bubble and propagation of shock waves generated by it are observed with changing distance between the electrodes and a solid wall. The effects of a solid wall on the time history of a bubble and the generated shock waves are experimentally clarified.Zusammenfassung Es wurde eine detaillierte experimentelle Untersuchung über das Verhalten von Blasen, die durch Funkenüberschläge zwischen Elektroden in Wasser erzeugt wurden, durchgeführt. Mit Hilfe einer Hochgeschwindigkeitskamera wurde die Bewegung einer kollabierenden Blase sowie die Ausbreitung der dabei entstehenden Stoßwellen beobachtet. Bei den Experimenten wurde der Abstand zwischen den Elektroden und einer festen Wand variiert. Der Einfluß der festen Wand auf das Verhalten der Blase und der Stoßwellen wurde durch das Experiment verdeutlicht.Sommaire Une décharge électrique dans l'eau produit une bulle unique dont on observe le comportement au moyen d'une caméra ultra-rapide. L'étude présentée utilise la méthode du réglage Imacon et la technique de la tomographie optique. On a pu ainsi suivre d'une façon continue l'évolution de la bulle depuis sa formation jusqu'a sa désintégration. On a mis en évidence les variations de cette évolution en fonction de la distance entre l'électrode de déclanchement et une paroi solide, ainsi que l'influence de cette paroi sur les ondes de choc produites par l'effondrement de la bulle.
Article
In acoustic cavitation the spatial variation and time–dependent nature of the acoustic pressure field, whether it is a standin or propagating wave, together with the presence of other bubbles, particles and boundaries produces gradients and asymmetrie in the flow field. This will inevitably lead to non–spherical bubble behaviour, often of short duration, before break–up int smaller bubbles which may act as nuclei for the generation of further bubbles. During the collapse phase, high temperature and pressures will occur in the gaseous interior of the bubble. This paper concentrates on the non–spherical bubble extensio to the earlier spherical–bubble studies for acoustic cavitation by exploiting the techniques that had previously been use to model incompressible hydraulic cavitation phenomena. Bubble behaviour near an oscillating boundary, jet impact and damag to boundaries, bubble interactions, bubble clouds and bubble behaviour near rough surfaces are considered. In many cases th key manifestation of the asymmetry is the development of a high–speed liquid jet that penetrates the interior of the bubble. Jetting behaviour can lead to high pressures, high strain rates (of importance to break–up of macromolecules) and toroida bubbles, all of which can enhance mixing. In addition it may provide a mechanism for injecting the liquid into the hot bubbl interior. Many practical applications such as cleaning, enhanced rates of chemical reactions, luminescence and novel metallurgica processes may be associated with this phenomenon.
Article
The motions of a gas bubble in proximity to a free surface with and without buoyancy force, as well as in shallow water are simulated based on a numerical time integration coupled with three-dimensional boundary integral spatial solution. The fluid is assumed to be inviscid, incompressible, and the flow irrotational. The unsteady Bernoulli equation is applied on the free surface and bubble surface as one of the boundary conditions of the Laplace equation for the potential. Improvements have been made in the mesh generation of the free surface and rigid boundary, the modeling of the toroidal bubble after the jet impact and the investigation into the combined effects on the motion of a bubble in the presence of the rigid bottom and free surface. The growth and collapse of a gas bubble together with the formation of the toroidal bubble after the jet impact are simulated. The shapes and positions of the bubble, the trajectories and velocities of the poles of the bubble as well as the pressure distributions in the fluid under different standoff distances and buoyancy parameters are obtained to better illuminate the mechanism underlying the motions of gas bubble and free surface. When a bubble is initiated sufficiently close to a free surface, the free surface spike and the second accelerating phenomenon of the free surface during the collapse phase can be observed. The buoyancy force has significant effects on the jet formation and development within the bubble and it may reverse the direction of the liquid jet when exceeding the effect of the Bjerknes force induced by the free surface. The large contortions in the shallow water and the formation of the high-pressure region between the bubble and the free surface are captured when the bubble is close enough to the rigid bottom and the free surface.
Article
Dynamics of explosion bubbles formed during underwater detonations are studied experimentally by exploding fuel (hydrogen and/or carbon monoxide)–oxygen mixture in a laboratory water tank. Sub-scale explosions are instrumented to provide detailed histories of bubble shape and pressure. Using geometric and dynamic scaling analyses it has been shown that these sub-scale bubbles are reasonable approximations of bubbles formed during deep sea underwater explosions. The explosion bubble undergoes pulsation and loses energy in each oscillation cycle. The observed energy loss, which cannot be fully explained by acoustic losses, is shown here to be partly due to the excitation of instability at the interface between the gaseous bubble and the surrounding water. Various possible mechanisms for the dissipation of bubble energy are addressed. The analysis of the experimental data gives quantitative evidence (confirmed by recent numerical studies) that the Rayleigh–Taylor instability is excited near the bubble minimum. The dynamics of the bubble oscillation observed in these experiments are in good agreement with experimental data obtained from deep sea explosions
Article
The nonlinear evolution of gas bubbles in the vicinity of a free surface is investigated numerically. The flow is assumed to be potential and a boundary-integral method is used to solve the Laplace equation for the velocity potential. The bubble content is described by an adiabatic gas law. For a bubble initiated at 1.5 maximum bubble radius from the quiescent free surface, three collapse patterns have been noticed: a downward directed Bjerknes jet for the case of weak buoyancy force; an upward directed buoyancy jet for the case of strong buoyancy force; and for the intermediate case with near-null Kelvin impulse state the bubble assumes an oblate form during collapse with no jetting. While the evolution patterns of a bubble may be different for different values of the buoyancy parameter δ, the variation of the bubble volume with time appears to be largely independent of δ. When the bubble is initiated at and less than 1.0RRm from the free surface, a much higher free surface spike is formed during the collapse phase. There are still three collapse patterns for the bubble depending on the buoyancy force imposed. The evolution of the bubbles in toroidal form is also simulated and examined. Finally, results are shown depicting the effect of initial ‘pressure’ on the evolution of the gas bubble near to the free surface.