Anda di halaman 1dari 17

316

2009,21(3): 316-332
DOI: 10.1016/S1001-6058(08)60152-3

NUMERICAL SIMULATION OF BUBBLE FLOW INTERACTIONS*


CHAHINE Georges L.
DYNAFLOW INC., 10621-J Iron Bridge Road, Jessup, Maryland 20794, USA,
E- mail: glchahine@dynaflow-inc.com

(Received October 18, 2008, Revised December 15, 2008)

Abstract: Bubble flow interaction can be important in many practical engineering applications. For instance, cavitation is a problem
of interaction between nuclei and local pressure field variations including turbulent oscillations and large scale pressure variations.
Various types of behaviours fundamentally depend on the relative sizes of the nuclei and the length scales of the pressure variations
as well as the relative importance of bubble natural periods of oscillation and the characteristic time of the field pressure variations.
Similarly, bubbles can significantly affect the performance of lifting devices or propulsors. We present here some fundamental
numerical studies of bubble dynamics and deformation, then a practical method using a multi-bubble Surface Averaged Pressure
(DF-Multi-SAP) to simulate cavitation inception and scaling, and connect this with more precise 3-D simulations. This same
method is then extended to the study of two-way coupling between a viscous compressible flow and a bubble population in the flow
field.
Key words: cavitation, bubble flow interaction, numerical simulation

1. Introduction 
Study of cavitation inception teaches us that
liquids rarely exist under a pure monophasic form and
that bubble nuclei are omnipresent. These nuclei are
very difficult to eliminate and are always in any
industrial liquid at least in very dilute concentrations.
However, most engineering studies, and rightfully so,
ignore the presence of this very dilute often invisible
bubble phase, and consider only the liquid phase to
conduct analytical or numerical simulations. This
applies to benign situations where pressure variations
are not significant and where cavitation, dynamic
effects, gas diffusion, and heat transfer do not result in
dramatic modification of the flow to warrant inclusion
in the modelling of both phases. In this contribution,
we are concerned with those conditions where it is
important to either evaluate whether cavitation may
occur and/or to model the flow in presence of bubbles
in significant local concentrations, sizes, or numbers
to play a significant role in the flow dynamics. Under
these conditions, interaction between the bubbles and


*Biography: CHAHINE Georges L. (1947-), Male, Ph. D.

the flow can be significant and need to be properly


addressed. Three families of situations can be
distinguished:
(1) Cavitation Inception in uniform or acoustic
flow fields.
(2) Cavitation Inception in vortical flow fields.
(3) Developed cavitation and bubbly flows.
In the first two cases, especially if cavitation
inception is determined acoustically (i.e., prior to
bubbles becoming visible), the interaction between
bubbles and liquid flow is 1-way, i.e., the liquid
dynamics affects the bubble dynamics, while bubble
effects on the hydrodynamic flow field are negligible.
In the third case, bubbles presence and behaviour is
influential enough to affect significantly the basic
flow field and implementation and use of 2-way
interaction models is warranted.
In this contribution, we will describe the work we
have conducted at DYNAFLOW over the recent years in
order to address the above aspects. This is not
intended to be a review of the research field, however,
references to other contributions will be done as
appropriate. The subdivision of the article follows the

317

three families of situations listed above.


2. Cavitation inception
Cavitation and bubble dynamics have been the
subject of extensive research since the early works of
Besant in 1859[1] and Rayleigh in 1917[2]. A host of
papers and articles and several books[3-10] have been
devoted to the subject. Various aspects of the bubble
dynamics have been considered and included one or
several of the physical phenomena at play such as
inertia, interface dynamics, gas diffusion, heat transfer,
bubble deformation, bubble-bubble interaction,
electrical charge effects, magnetic field effects, etc..
Unfortunately, very little of the resulting knowledge
has succeeded in crossing from the fundamental
research world to the applications world, and it is
uncommon to see bubble dynamics analysis made or
bubble dynamics computations conducted for example
by propeller or pump designers. This is due to the
perceived impracticality of using the methods
developed but for experts. However, with the
tremendous advances in computing resources, these
communities now commonly use Navier Stokes
solvers and CFD codes to seek better solutions[11-15].
The challenge is thus presently for the
cavitation/bubble community to bring its techniques to
par with the single phase CFD progress. Consideration
of bubble dynamics within a CFD computation should
also become common practice. Indeed, the tools
already developed are at the reach of all users and
should be adopted as more CFD tools to use for
advanced design[12,13].
2.1 Presence of cavitation nuclei
A traditional cavitation inception engineering
definition is: a liquid flow experiences cavitation if
the local pressure drops locally below the liquid
vapour pressure, Pv . This definition of cavitation
inception is only true in static conditions when the
liquid is in contact with its vapour through the
presence of a large free surface and is in practice
applicable for liquids saturated with bubbles. For the
more common condition of a liquid in a flow, or in a
rotating machine, liquid vaporization can only occur
through the presence of microscopic free surfaces or
microbubbles, also called cavitation nuclei. These
are micron and sub-micron sized bubbles in
suspension in the liquid or trapped in particles
crevices. Several techniques have been used to
measure these nuclei distributions both in the ocean
and in laboratory cavitation channels. These include
Coulter counter, holography, light scattering methods,
cavitation susceptibility meters, and acoustic bubble
spectrometers[16-22].
Therefore, any fundamental analysis of cavitation
inception has to start from the observation that, any

real liquid contains nuclei which when subjected to


variations in the local ambient pressure will respond
dynamically by oscillating and eventually growing
explosively (i.e., cavitating). Cavitation inception
cannot be defined accurately independent of this
liquid bubble population (sometimes characterized by
liquid strength[23]) and independent of the means of
cavitation detection. The cavitation inception is in fact
a complex dynamic interaction between the nuclei and
their surrounding pressure and velocity fields,
interaction that can be different between the small and
large scale of the same configuration. In addition, the
experimental means to detect and decide on calling
cavitation inception (practical threshold used by the
experimentalists) will affect the results and could be
different between laboratory experiments and full
scale tests.
2.2 Inception in quasi uniform flow: spherical model
When the pressure variations to which a bubble
is subjected are not slow compared to the bubble
response time, the nuclei cannot instantaneously adapt,
inertia effects become important, and one needs to
consider bubble dynamics effects. This is the case for
nuclei travelling through rotating machinery. The
nuclei /bubbles then act as resonators excited by the
flow field temporal and spatial variations. In the
case of a vortical flow field the strong spatial pressure
gradients (in addition to the temporal) strongly couple
with the actual bubble motion (i.e., position vs. time)
to result in a driving force that depends on the
resonator reaction. This makes such a case, considered
in the next section, much more complex than what
occurs for a travelling bubble about a foil where,
relatively speaking, the position of the bubble is less
coupled to its dynamics.
The flow field pressure fluctuations have various
time scales: e.g., relatively long for travelling
cavitation bubbles over a blade or captured in a
vortical region flow, or very short for cavitation in
turbulent strongly sheared flow regions. The
amplitudes of these fluctuations and the relationship
between the various characteristic times determine the
potential for cavitation inception.
For cavitation inception in uniform flow fields it
is common to use spherical bubble models to examine
the dynamics. This is justified as long as the local
velocity and pressure gradients around the bubble are
weak compared to the radial gradients, and as long as
proper account is made of the bubble time dependent
location in the flow field i.e., account for any slip
velocity relative to the liquid flow. Spherical models
have been extensively studied following the original
works of Rayleigh[1] and Plesset[10]. For instance, if
we limit the phenomena to be modelled to inertia,
small compressibility of the liquid, compressibility of
the bubble content, we obtain the Gilmore[24]

318

differential equation for the bubble radius R (t ) . We


modified this equation to account for a slip velocity
between the bubble and the host liquid, and for the
non-uniform pressure field along the bubble surface[25].
(This is further discussed later.) The resulting
Surface-Averaged Pressure (SAP) equation applied to
Gilmores equation[25,26] becomes:

R  3
R  1 R R d
RR
1
+
1



R = 1+ +
<
2 3c
U c c dt
c

(1)

where c is the sound speed, P is the liquid


viscosity, u is the liquid convection velocity and
ub is the bubble travel velocity.
Equation (5) degenerates to the classical
equation
for
negligible
Rayleigh-Plesset[10]
compressibility and slip velocity effects.
If in
addition, gas diffusion and heat transfer effects are
neglected and a polytropic law of gas compression is
assumed, the resulting modified SAP equation
becomes:
3k

3 2 1
R0

RR + R = pv + pg 0  Pencounter 
U
2
R

1 2J 4 P R u  ub
+

+
4
U R
R

the gas diffusion problem and the assumption that the


gas is an ideal gas[28].
The bubble trajectory is obtained using the
following motion equation[29]

dub 3
3
= P + CD u  ub u  ub +
dt
U
4
CL u  ub u u  ub +

2V
R
 4P +
pv + pg  pencounter 
R
R

u  ub

leads to a more realistic bubble dynamics. In general,


the gas pressure, p g , is obtained from the solution of

(2)

where k is the polytropic compression law constant.


In the Surface-Averaged Pressure (SAP) bubble
dynamics equation, we have accounted for a slip
velocity between the bubble and the host liquid, and
for a non-uniform pressure field along the bubble
surface. In this SAP method the definition of
Pencounter as the average of the liquid pressures over the
bubble surface results in a major improvement over
the classical spherical bubble model which uses the
pressure at the bubble center in its absence[25-27]. This
has serious implications for strongly non-uniform
flows as discussed in the next section. For instance, a
bubble does not always continuously grow once it is
captured by a vortex. Instead, it is subjected to an
increase in the average pressure once it grows and this

3
u  ub R
R

(3)

where the drag coefficient C D is given by an


empirical equation such as that of Haberman and
Morton[30]:

CD =

Reb =

24
1.38
(1+ 0.197 Reb0.63 + 2.6 u 104 Reb
),
Reb

2 U R u  ub

(4)

2.3 Inception in vortical flow fields: Axisymmetry


Spherical bubble models, as briefly described
above, can be efficient tools for studying cavitation
inception, scaling, bubble entrainment, and cavitation
noise. They can become more powerful if they are
provided with further intelligence based on more
precise non-spherical models which account for
bubble behavior near boundaries, in pressure gradients,
and in high shear regions, resulting in bubble
deformation, elongation, splitting, coalescence, and
non-spherical sound generation.
One such refinement consists of considering the
case of bubbles captured on a vortex axis. The bubble
then elongates along the axis and may split into two or
more sub-bubbles, and/or form jets on the axis. In
order to investigate this behaviour our boundary
element method axisymmetric bubble dynamics code
2DYNAFS[31-38] was exercised and was able to
simulate bubble dynamics through re-entrant jet
formation, jet break through, and bubble splitting. The
code can handle as input vortex flow fields obtained
from CFD viscous computations or from experimental
measurements.
2DYNAFS simulations of bubble dynamics on a
vortex axis under a large number of conditions lead to
the followings conclusions illustrated in Fig.1[32-38]:
(1) If the bubble is captured by the vortex far
upstream from the minimum pressure, it remains
spherical while oscillating at its natural frequency.

319

(2) When the bubble reaches the axis just


upstream of the minimum pressure, it develops an
axial jet on its downstream side which shoots through
the bubble moving in the upstream direction. Even at
this stage, the spherical model provides a very good
approximation because the bubble is more or less
spherical until a thin jet develops on the axis.
(3) The bubble behaviour becomes highly
non-spherical once it passes the minimum pressure
location. It elongates significantly and can reach a
length to radius ratio that can exceed 10. The bubble
then splits into two or more daughter bubbles emitting
a strong pressure spike followed later by other strong
pressure signals when daughter bubbles collapse. Two
axial jets originating from the split and a strong
pressure signal during the formation of the jets are
observed.

conclusion that has been preliminarily confirmed


experimentally[32,36,38]. Three types of tests were
conducted: spark generated bubbles, laser generated
bubbles, and electrolysis bubbles injected in vortex
lines. Figure 2 shows high speed photography and
acoustic signals of bubble splitting between two rigid
walls. A small but distinct pressure spike is formed at
splitting followed but a more significant spike during
the collapse of the sub-bubbles. The second set of
experiments was conducted in a vortex tube, where
bubbles generated by electrolysis were injected and
observed once captured by the vortex line. Figure 3
shows the elongated bubble dynamics and the
corresponding pressure signals[35]. The third set of
experiments was conducted at the University of
Michigan[36,38] using laser induced bubbles (in the
vortex and far upstream) and the flow field of a tip
vortex behind a foil. (Shown in Fig.4) Comparisons
between the observations and the 2DYNAFS
simulations showed very good correspondence as
shown in Fig.5. Detailed correspondence was harder
to establish because of difficulty of measurements of
all experimental microscale conditions.

Fig.1 Illustration of the acoustic pressure emitted by a bubble


in a vortex field as a function of the cavitation number.
Note that the bubble behaviour near and above the
cavitation inception is quasi-spherical[32]

Fig.3 High speed photos of an electrolysis bubble captured in


a line vortex, and resulting acoustic signal indicating
peak signals at splitting and collapse[32]

Fig.2 High speed photos of spark-generated bubble collapsing


between two solid walls, and resulting acoustic signal.
Peak signals at splitting and sub-bubbles collapses[32]
Fig.4

2.4 Experimental verification


This behaviour supports the hypothesis that the
noise at the inception of the vortex cavitation may
originate from bubble splitting and/or from the jets
formed after the splitting. This is an important

Bubble

behaviour

Rc = 4.51mm ,

in

vortex

* =0.2123 m2 /s ,

flow

field:

V =1.72 ,

U f =10m/s , R0 = 750Pm . 3-D view of the bubble just


before splitting predicted by 2DYNAFS and observed
at University of Michigan using laser-induced bubbles
at the center of the vortex[36]

320

2.6

Fig.5 Bubble behaviour in a vortex flow field[26]

2.5 Bubble splitting criteria


The numerical simulations resulted in the
following observations on the splitting of bubbles
captured in a vortex (see Fig.6), which we use in our
predictive models[31-35]:
(1) An explosively growing bubble splits into
two sub-bubbles after it reaches its maximum volume,
(equivalent radius, Rmax ) and then drops to 0.95
Rmax .
(2) The two resulting sub-bubbles have the
following equivalent radii: 0.90 Rmax and 0.52 Rmax .
(3) The locations of the two sub-bubbles after the
splitting are at 0.95 Rmax and 4.18 Rmax on the
vortex axis.
(4) The pressure generated by the subsequent
formation of reentrant jets in the sub-bubbles can be
approximated by a function of V [36].
Since the noise associated with the jet formation
appears to be much higher than the pressure signal
from the collapse of a spherical bubble, it is desirable
to include the splitting and the associated jet noise in
simulations with multiple bubble nuclei.

Fig.6 Sketch of the successive phases of a bubble behavior in a


vortex flow field

Inception in vortical flow fields: Fully


non-spherical
In order to study the full 3-D interaction between
a bubble and a complex flow field, two methods were
developed. The first, using our boundary element code,
3DYNAFS, enables study of full bubble deformations
during capture but neglects the effects that the bubble
may have on the underlying flow field. The second
method accounts for full 2-way bubble/flow field
interaction, and considers viscous interaction. This
model is embedded in an unsteady ReynoldsAveraged Navier-Stokes (RANS) code, DF-UNCLE 1,
with appropriate free surface boundary conditions and
a moving Chimera grid scheme[26]. This full 2-way
interaction non-spherical bubble dynamics model has
been successfully validated in simple cases by
comparing the results with base case results obtained
from the Rayleigh-Plesset equation and 3DYNAFS for
bubble dynamics in an infinite medium both with and
without gravity.

Fig.7

Comparison of the bubble radius versus time for the


spherical models (the conventional and the SAP model)
and the 3-D 1-way and 2-way UnRANS
computations[26]

As an illustration, Fig.7 shows results of a bubble


interacting with the tip vortex of an elliptical foil. The
bubble elongates once it is captured, and depending on
the cavitation number, either forms a re-entrant jet
directed upstream or splits into two sub-bubbles.
When 2-way interaction is taken into account, further
smoothing of the bubble surface is exerted by
viscosity resulting in a more distorted but overall
more rounded bubble. Figure 8 illustrates the various
stages of the interaction between a bubble and a tip
vortex flow.

1

DF_UNCLE, new dubbed 3DYNAFS-VIS is based on UNCLE


developed by Mississippi State University that we extended to
3-D cavity and free surface dynamics.

321

2.7 Validation of the SAP model


Here we compare the SAP spherical model and
the 3-D interaction non-spherical models to predict tip
vortex cavitation inception for a tip vortex flow
generated by a finite-span elliptic hydrofoil[26]. The
flow field was obtained by a RANS computation,
which provided the velocity and pressure fields for all
compared models: the classical spherical model, the
SAP model, the one-way interaction model where the
bubble deformed and evolved in the vortex field but
did not modify it, and finally the fully coupled 3-D
model in which unsteady viscous computations
included modification of the flow field by the
presence of the bubbles.

considered to be distributed randomly in a fictitious


supply volume feeding the inlet surface or release area
of the computational domain. The fictitious volume
size is determined by the sought physical duration of
the simulation and the characteristic velocity in the
release area as illustrated in Fig.9. The liquid
considered has a known nuclei size density
distribution function, n( R ) , which can be obtained
from experimental measurements[17-22] and can be
expressed as a discrete distribution of M selected
nuclei sizes. Thus, the total void fraction, D , in the
liquid can be obtained by
M

D = Ni
i =1

4SRi3
3

(5)

where N i is the discrete number of nuclei of radius


Ri used in the computations. The position and thus
timing of nuclei released in the flow field are obtained
using random distribution functions, always ensuring
that the local and overall void fraction satisfy the
nuclei size distribution function.

Fig.8

Comparison of the maximum bubble radius vs.


cavitation number between the spherical models (the
conventional and the SAP model) and the 3-D 1-way
and 2-way UnRANS computations[26]

Comparisons between the various models of the


resulting bubble dynamics history, and of the
cavitation inception values obtained from many tested
conditions, reveal the following conclusions,
illustrated in Fig.7 and Fig.8:
(1) The bubble volume variations obtained from
the full 2-way interaction model deviate significantly
from the classical spherical model due to the
interaction between the bubble and the vortex flow
field.
(2) Differences between the 1-way and 2-way
interaction models exist but are not major.
(3) Using the SAP scheme significantly improves
the prediction of bubble volume variations and
cavitation inception. SAP appears to offer a very good
approximation of the full interaction model.
2.8 Modelling of a real liquid nuclei field
We have developed the following general
approach to address bubble flow interactions in
practical conditions. In order to simulate the water
conditions with a known distribution of nuclei of
various sizes, as illustrated in Fig.9, the nuclei are

Fig.9

Illustration of the fictitious volume feeding the inlet to


the nuclei tracking computational domain (or release
area) and example of resulting nuclei size distribution
satisfying a given nuclei density distribution function

2.9 Effect of a propeller on a nuclei field


In order to simulate the effect of a propeller on a
nuclei flow field, the methods described above were
used. Gas diffusion into the bubbles was also

322

included, since it is very important in this case. For


the nuclei field, the fictitious volume size was
determined as the product of the release area, the
sought physical duration of the simulation, and the
characteristic inlet velocity as illustrated in Fig.10.
For the results shown below we have selected
characteristic values: size range of 10 to 200 m and
void fraction, D = 3.23 u 105 . Figure 11(a) shows the
selected nuclei size number density and Fig.11(b)
shows the numbers of discrete bubble sizes and band
widths selected.

necessary and the effects on the results were analyzed.


Figure 12 shows the time variation of the total number
of bubbles within the computational domain for
different number of repeats. The results clearly
indicate that a steady state is achieved after
t > 0.03ms with a minimum of 10 repeats
effectuated. Otherwise, the two-phase medium is in
a transient phase. Repeating the number of bubbles
more than 10 times increases the period of
quasi-steady state in the computational domain.

Fig.10

Fig.12

Effect of the bubble release duration on the total


number of bubbles within the computational domain.
Computations were conducted for 't = 3.7 u 10-3 s
and results were then repeated to cover much longer
release times

Fig.13

Front and side view of the bubbles in the computational domain at a given time during the
computation. Bubble sizes are to scale in Section b,
They are enhanced by a factor of 5 in Sections a and c

Illustration of the fictitious volume feeding the inlet to


the nuclei tracking computational domain (or release
area)

Fig.11 (a) The nuclei size distribution selected to resemble


the field measurement by Medwin[19] and (b) the
resultant number of nuclei released

In order to minimize CPU time, a time of release,


't , during which actual bubble computations are
conducted, was selected and corresponded to a
meaningful bubble population distribution in the
fictitious bubble supply volume. The resulting bubble
behavior was then assumed to be repeatable for the
next 't . This was repeated as many times as

Figure 13 illustrates the sizes and locations of


bubbles that became visible. Both front and side

323

views are shown at a given instant for = 1.75 . It is


noted that for the side view the bubbles in Section b
were illustrated with the actual sizes, while they were
amplified 5 times to make them more visible in print
in Sections a and c. Figure 13 illustrates the marked
increase in visible bubbles downstream of the
propeller, traveling bubble cavitation, tip vortex
cavitation and a resemblance of sheet cavitation.
In order to illustrate the results in a simplified
manner, a time dependent void fraction by section,
D ( x, t ) , is defined by integrating bubble volumes in
between x and x + dx :
N ( x ,t ) 4

i =1

D ( x, t ) =

SRi3

1
S D 2  d 2 'x
4

(6)
Fig.15 Comparison of the initial nuclei size distribution and
that resulting at x = 0.2m

where d is the shaft diameter, and N ( x, t ) is the


number of bubble within 'x at a given time, t . The
time-averaged void fraction of D ( x, t ) at t1 , over a
time period dt can also be defined as:

D=

1 t1+dt
D ( x, t )dt
dt t

x = 0.2m . It is seen that a large number of bubbles of


400 m radius are now present in the field, while all
bubbles upstream were smaller than 200 m.
The effect of the cavitation number, the initial
nuclei size distribution, and the initial dissolved gas
concentration, on the bubble nuclei entrainment and
dynamics were shown below.

(7)

Figure 14 shows the distribution of D along x .


It is seen that the void fraction increases significantly
as a result of cavitation on the blades (first hump) and
in the tip vortices (second hump). The void fraction
decreases as the encountered pressure by the bubble
increases. However, the downstream void fraction is
1.4 times that of upstream.

Cavitation number:
The time-averaged void fraction and downstream
nuclei size distribution obtained at three different
cavitation numbers, V = 1.5, 1.65 and 1.75, are
compared in Fig.16 and Fig.17. In the three cases the
void faction increases significantly near the propeller
then plateaus. As the cavitation number decreases, the
void fractions within the cavitation areas increase.
The relative importance of the tip vortices diminishes
relative to the propeller blade cavitation. With smaller
values of V , the bubbles retain a much larger size
downstream of the propeller due to enhanced net gain
of gas mass diffusing into the bubbles. The
downstream void fraction D at x = 0.2m is 50
times the upstream value for V = 1.5, 2.35 times for
V = 1.65 and 1.4 times for V = 1.75 . The largest
downstream bubble radii observed are 800 m for
V = 1.5 , 600 m for V = 1.65 and 400 m for
V = 1.75 , as compared to 200 m at release.

Fig.14 The time-averaged void fraction, D , distribution along


the axial direction for D = 3.23 u 105 and V = 1.75

The effect of propeller cavitation on the number


of observable bubbles is much more important than
on D . Figure 15 shows the bubble size distribution at

Fig.16

Comparison of the void fraction variations along x


for different cavitation numbers

Nuclei distribution:

324

The effect of the initial nuclei size distribution on


the results is shown in Fig.18. A bubble size
distribution ranging from 10-200 m is compared to
that of a uniform distribution of 100m bubbles with
the same initial void fraction. Higher bubble void
fractions are seen near the propeller blade region for
the initial uniform nuclei size case. However, this
does not result in any significant difference in the void
fraction downstream. Figure 19 compares the resulting
downstream nuclei size distributions at x = 0.2m . It
is seen that the largest bubble radius is 400 m for the
non-uniform nuclei distribution, while it is 300 m for
the uniform case.

in further increase in the void fraction downstream.


However, although the dissolved gas concentration
was doubled, the void fraction increased only 11%.

Fig.20 Comparison of the void fraction variation along the


streamwise direction for two initial gas concentrations

3. Modeling of bubbly flows


3.1 Bubbly flow and propeller/hydrofoil
In order to address problems where bubbles
significantly modify the flow, we have developed a
two-way coupled method between the viscous solver:
3DYNAFS-VIS and multi-bubble dynamics solver,
DF_MULTI_SAP. The coupling between the
two-phase continuum flow model and the bubble
discrete dynamics/tracking model is an essential part
of this method and can be described as follows:
(1) The bubbles in the flow field are influenced
by the local densities, velocities, pressures, and their
gradients in the medium. The dynamics of the
individual bubbles and their tracking are based on
these local flow variables.
The local properties of the mixture continuum are
determined by the bubble presence. The void fractions,
and accordingly the mixture local densities, are
determined by the bubble population and size, and
evolve as the bubbles move and oscillate. The flow
field is adjusted according to the modified mixture
density distribution while satisfying mass and
momentum conservation.
(2) The void faction is obtained by subdividing
the computational domain in D -cell where D is
computed and measuring the volumes of all bubbles in
such cells. An D -cell is large enough for the statistics
to be meaningful, i.e., it should contain a large enough
number of bubbles, and small enough, compared to
the flow field local characteristic length, such that
bubbles of equal size in the cell will behave
practically the same way.
This approach was applied to a NACA4412 foil
at zero angle of attack with various void fractions.
Figure 21 shows the void fraction distribution for a
hydrofoil operating in a bubbly flow with 5 % free
stream void fraction. It is seen that the void fraction
level decreases near the leading edge due to screening
effects. High concentrations of bubbles are observed
near the suction side surface, while the concentration
P

Fig.17 Bubble size distribution at


different cavitation numbers

x = 0.2m

for three

Fig.18 Comparison of the void fraction variation along the


streamwise direction for two initial nuclei size
distributions

Fig.19 Bubble size distribution at x = 0.2m for two initial


nuclei size distributions

Dissolved gas concentration:


The effect of the dissolved gas concentration on
the bubble dynamics is illustrated in Fig.20 by
doubling the dissolved gas concentration. This results

325

is very low in the wake region.

Fig.21 Void fraction distribution for a hydrofoil operating in a


bubbly flow

Figure 22 shows the effect of the void fraction on


the pressure coefficients on the foil surface. As D
increases, the pressure on the suction side increases,
which results in loss of lift compared to the single
phase flow. The drag has contributions from both skin
friction and pressure drag. Skin friction decreases
when bubbles are present near the hydrofoil surface,
the drag due to the pressure can increase. To illustrate
this, we plot the pressure coefficients with respect to
the vertical coordinate in Fig.23. The BC and CD
portions of the hydrofoil surface contributes to drag,
while the AB and AD portions contribute to thrust.
The total pressure drag is related to the area between
these curves, which increases as the void fraction
increases.

Fig.22 Pressure coefficient vs. x curves on the surface of the


hydrofoil for various void fractions

The changes of lift and drag from the single


phase flow are shown in Fig.24. It is seen that the
percentage of lift reduction is greater than the
percentage of void fraction increase. The total drag
resulting from both pressure drag and skin friction is
shown to increase as the void fraction is increased
because the increase of the drag due to the pressure is
greater that the reduction of drag due to the skin
friction.

Fig.23 Pressure coefficient vs. z curves on the surface of the


hydrofoil for various void fractions

Fig.24 Hydrofoil performance, the lift and drag changes as


functions of void fraction

3.2 Bubble augmented propulsion


Several publications[39-46] and patents[47-49] claim
that bubble injection in a propulsor or a waterjet can
significantly augment the thrust even at high speeds.
Experimental work has been conducted where fluid
enters the ramjet and is compressed by passing
through a diffuser (ram effect). High pressure gas is
then injected into the fluid via mixing ports and
constitute the energy source for the ramjet. The
multiphase mixture is then accelerated by passing
through a converging nozzle. Various prototypes have
been developed and tested in the past and have shown
a net thrust due to the injection of bubbles. However,
the overall propulsion efficiency was typically less
than that anticipated from mathematical models.
These results require further modelling and controlled
experimentation as we are presently undertaking at
DYNAFLOW.
First approximation of the multiphase flow field
inside the air augmented nozzle were obtained by
neglecting all non-uniformities in the cross-section
and considering only a quasi-one dimensional flow
field[50-54]. The fluid medium is considered as the
mixture of water and bubbles which satisfies the

326

following continuity and momentum equations:

wU m
+ < U m um = 0
wt

Um

(8)

Dum
2

= pm + 2 P mG ij  Pm <um
Dt
3

(9)

where the subscript m represents the mixture medium.


Here the G ij in the stress tensor term is the Kronecker
delta, the mixture density, U m , and the mixture
viscosity, P m , are functions of D :

Um = 1  D Ul + DU g

(10)

Pm = 1  D Pl + DP g

(11)

where the subscript l represents liquid, and the


subscript g represents the gas in the bubble. The
flow field has a variable density because the void
fraction, D , varies in space and in time. This makes
the overall flow field problem similar to a
compressible flow problem.
One approach we are developing is to use a
compressible single phase flow solver and replace the
equation of state which connects the density to the
pressure with a solution of the bubble phase using the
Lagrangian bubble tracking model described earlier,
MULTISAP, described earlier and which computes
bubble size and location and thus D ( x , t ) .
The 2-way coupling between the two is as
follows:
(1) The bubbles in the flow field are influenced
by the local density, velocity, pressure, and the
pressure gradient of the mixture medium. The
dynamics of individual bubbles are based on these
local flow variables as described in Eqs.(8) through
(11). This is achieved using our Discrete Bubble
Model (DBM).
(2) The mixture flow field is influenced by the
presence of the bubbles. The void fraction, and
accordingly the mixture density, is modified by the
migration and dynamics of the bubbles. The flow field
is adjusted according to the modified mixture density
distribution in a way that the continuity and
momentum are conserved through Eqs.(8) and (9).
The 2-way interaction described above is very
strong for this problem since D can vary from zero
near the inlet to as high as 70% at the nozzle exit.

Fig.25 The geometry of the nozzle inside wall as modeled.


Looking from the downstream of the nozzle exit

3.3 Example study


The tow pool experiment in Ref.[40] was chosen
as a test case because there are measured data
available. The nozzle is shown in Fig.25. The water
comes through the inlet, flows through a diffuser, then
an expansion with a bubble injector, and finally
through the converging exit nozzle. The tests were
performed at 8 m/s, and thrust, air flow, and pressure
were measured at four locations inside the nozzle.
The flow was modelled both in 3DYNAFS-VIS and
Fluent. Boundary conditions were set so that the inlet
has 8 m/s velocity and the exit has the pressure of 1
atm.

Fig.26

The axial velocity distribution on the center plane of


the nozzle

Fig.27

The pressure distribution on the center plane of the


nozzle

327

Figure 26 shows the velocity distribution on the


center plane of the nozzle. As expected, the velocity
inside the nozzle decreases as the cross sectional area
increases up to the bubble injector, and then increases
toward the exit as the cross section becomes gradually
narrower. The corresponding pressure as shown in
Fig.27 is the highest just downstream of the bubble
injector. The inlet pressure, which is predicted to be
84 kPa, matches with the experimental observation as
reported in Ref.[40].
3.4 Steady state 1-D approach
If one is interested in the steady state solution for
a relatively simple axisymmetric nozzle, a steady 1-D
approach can be used where the bubbles are assumed
to behave the same regardless of their locations in the
cross sections of the nozzle. One then obtains the void
fraction from tracking bubbles through the axial
pressure field of the nozzle and conducting an
iteration procedure for the 2-way interaction between
the bubbles and the flow. The procedure is as follows:
(1) Track the bubble in the flow field. At the first
step, the liquid flow is used as an initial guess. Later
in the iteration, the most recently updated flow field is
used.
(2) From the tracking, the bubble radius is found
as a function of axial coordinate along the nozzle.
(3) Deduce the medium density from the void
fraction using Eq.(10).
(4) Solve the flow field using the updated
medium density.
(5) Re-compute bubble motion and size as in step
1 and iterate until convergence.
The iteration described in the previous section
converges well for the simulated nozzle. In Fig.28, the
convergence of the bubble radius is shown indicating
good convergence as early as in the 3rd iteration.
Figure 29 shows a similar fast convergence of the
corresponding flow field solution, i.e., the pressure
distribution along the centerline of the nozzle.

Fig.28 The convergence of the tracked bubble radius

Fig.29

The convergence of the pressure along the centerline


of the nozzle

Fig.30 The bubble radius as it flows inside the nozzle

Fig.31

The void fraction distribution inside the nozzle. The


initial void fraction at the injection is 0.5 (the dashed
line)

Fig.32

The converged pressure distribution inside the nozzle.


The pressure along the centerline (top), and the
pressure distribution on the center plane (bottom)

328

A converged behavior of a 5 mm radius bubble


injected at a 2 atm pressure and 1.5 m downstream of
the inlet is shown in Fig.30. Since the pressure inside
the nozzle at the injection location is about 1 atm, the
bubble grows after the injection then oscillates as it
flows down and reaches an equilibrium radius. Figure
31 shows the converged void fraction distribution
inside the nozzle. It is zero from the inlet to the
injection location. Then it suddenly rises at the
injection point and overshoots due to the bubble
oscillations. Near the outlet the void fraction increases
slightly because the bubble grows due to the pressure
drop near the outlet. The pressure distributions inside
the nozzle are shown in Fig.32and Fig.33. The axial
velocity increases toward the outlet, and reaches as
high as 15 m/s which is much higher than the exit
velocity of the water only flow (Fig.26).

The thrust predicted from the steady 1-D model


is shown in Fig.34. As the void fraction increases,
there is a large gain in the momentum component of
the thrust, which is countered by a slight loss in the
pressure component. As a result, the total thrust
increases as the void fraction increases. At the initial
void fraction of 50%, the total thrust gain above the
water only flow is about 1100 N.

Fig.34 Predicted thrust for various initial void fractions

Fig.33 The converged axial velocity distribution inside the


nozzle. The axial velocity distribution on the center
plane (top), and the axial velocity along the centerline
(bottom)

The thrust of the nozzle can be computed by


integrating the pressure and the momentum flux over
a surface of a control volume that contains the nozzle.

T = pm + U mum2 dA

(12)

where u m is the axial component of the mixture


velocity. For the inside of the nozzle the thrust is
defined as follows:

T  po Ao  pi Ai + U o Ao umo 2  Ui Ai umi 2

(13)

where the subscripts i and o represent the inlet


and the outlet, respectively. The component of the
thrust described in the first parenthesis in Eq.(13) is
the contribution from the pressure difference between
the inlet and the outlet. The second parenthesis
represents the thrust due to the momentum change
between the inlet and the outlet.

Fig.35

Comparison of the pressure field for the quasi-steady


and Rayleigh-Plesset models for the quasi 1-D
problem

3.5 Results from quasi 1-D models


Both quasi-steady and Rayleigh-Plesset based
models were applied to the same configuration. The
operating conditions were determined such that the
exit pressure was atmospheric and the inlet velocity
was held constant at 8 m/s. The bubbles were injected
near the end of the injection section of the nozzle. For
the cases presented here, the injection was
implemented as a discontinuity in the void fraction.
The pressure was kept constant across it while the
velocity was suddenly increased in order to preserve
the mass flow rate. The pressures for a void fraction
injection of 20% and 50% are compared in Fig.35. It
is important to note that for a 20% void fraction
injection, the two models provided nearly identical
answers. However, at 50% void fraction injection, the
pressure field exhibited a marked difference between
the two models. It is important to note that the
quasi-steady approach predicts larger bubbles at the

329

nozzle exit. This is expected since the inertial effects


modelled by the Rayleigh-Plesset equations will retard
bubble growth.

Fig.36 Comparison of the total thrusts obtained by the


DBM+Fluent approach and two 1-D models

3.6 Comparison with two-way approach


The predictions using the DBM and Fluent solver
are compared to two other 1-D models described
earlier. The total thrust is compared in Fig.36. They
match well for low void fractions, but DBM deviates
from the others for higher void fractions. The pressure
distribution along the nozzle is compared in Fig.37.
The pressure distribution obtained by DBM+Fluent
approach and by the 1-D quasi-steady model matches
well for D = 20% , but is different for =50%. In these
computations, the outlet pressure is fixed at 1 atm.
In the case of D = 50% , the difference of the
pressure gradient is prominent in the exit nozzle.
Figure 38 shows the void fraction along the nozzle. It
is zero from the inlet to the injection (at 1.5 m). After
the injection, increases mildly for the injection
D = 20% , while it increases more rapidly for the
D = 50% . The large difference between DBM+Fluent
approach and 1-D quasi-steady model is observed near
the outlet. This is due to the bubble dynamics that
includes the slip of the bubble relative to the fluid.

Fig.38 Comparison of the void fraction along the length of the


nozzle predicted by DBM+Fluent approach and 1-D
quasi-steady model. The injection void fractions of 20%
and 50% are shown

3.7 3-D unsteady coupling


For full 3-D interactions, the void fraction is
defined in the 3-D space using the D -cells described
earlier. The bubbles and their volumes in each D -cell
are counted to compute the void fraction.
Improvement in the distribution of the void fraction
by introducing a smoothing over the neighbouring
bins was necessary to improve convergence of the
solutions. The DF_UNCLETM and DF_MULTI_SAP
coupled approach can run a simulation in unsteady
2-way coupled way.
The procedure for unsteady coupled run is as
follows:
(1) Initialize flow field, e.g., from water steady
state solution.
(2) Track the bubbles in the flow field following
injection.
(3) Based on the current bubble size and location,
compute the void fraction D x, y, z, t , using the
P

D -cells.
(4) Deduce the medium density U m x, y, z , t

from D x, y, z, t .
(5) Solve the flow field using the updated
medium density, and proceed to the next time step.
(6) Repeat from step 2 to step 5 until the desired
simulation time is reached.
Figure 39 shows an example simulation using a
full
unsteady
two-way
coupling
between
DF_UNCLETM and DF_MULTI_SAP. Three different
size bubbles were injected on a bottom port near the
inlet. It can be observed that the larger bubbles rise
faster due to the buoyancy than smaller bubbles and
the largest bubbles contribute predominantly to the
medium density. Figure 40 shows an instantaneous
snapshot of bubbles, medium velocity vectors, and
density contours for a distributed bubble size injection
at the widest section of the propulsor.
P

Fig.37 Comparison of the pressure distribution along the length


of the nozzle predicted by DBM+Fluent approach and
1-D quasi-steady model. The injection void fractions of
20% and 50% are shown

330

This model was finally coupled with a viscous


and compressible flow solver to model the two phase
flow in the case of bubble injection for thrust
augmentation. Work is on-going to transform this into
a generalized multi-scale two-phase flow model.

Fig.39 DF_Uncle and DF_Multi_SAP coupled simulation


for a bubbles mixture injected from a bottom port

Acknowledgements
The author acknowledges the sustained support
of the Office of Naval Research, Dr. Kim Ki-Han
monitor, and the contributions of many DYNAFLOW
colleagues, most particularly Dr. Hsiao Chao-Tsung
and Dr. Choi Jin-Keun who are responsible for most
of the numerical simulations presented here.
References
[1]
[2]
[3]

Fig.40

Snapshot of computed bubble distribution and mixture


medium density distribution on the center plain of the
bubble augmented propulsion test setup. Only the
portion of the nozzle between the injection and exit is
shown

These solutions are being compared against an


experimental fundamental study that we are
conducting in a Plexiglas quasi-2-D cross section of
the propulsor.

[4]
[5]
[6]

[7]
[8]

4. Conclusions
We have presented in this communication our
efforts to model bubble flow interactions under
various conditions and for several families of
engineering applications. These included modelling of
cavitation inception and development, estimation of
the effect of air bubbles on the drag and lift of a lifting
surface, and the modelling of purposely injected
bubbles on the thrust of a propulsor or nozzle.
The models developed were based on studies of
single bubble dynamics and deformation in complex
flow fields and used the acquired knowledge into a
simplified Rayleigh-Plesset like model using bubble
surface averaged quantities to compute the bubble
behaviour in the flow field. This SAP model enables
modelling in reasonable computation times of a real
nuclei field distribution and is presently available as a
User Defined Function (UDF) in Fluent.

[9]
[10]
[11]

[12]

[13]
[14]

BESANT W. H. Hydrostatics and hydrodynamics[M].


London: Cambridge University Press, 1859, 158,
RAYLEIGH Lord, On the pressure developed in a liquid
during collapse of a spherical cavity[J]. Phil. Mag.,
1917, 34: 94-98.
KNAPP R. T., DAILY J. W. and HAMMITT F. G.
Cavitation[M]. Yew York: McGraw Hill, 1970.
HAMMITT F. G. Cavitation and multiphase flow
phenomena[M]. Yew York: McGraw-Hill,1980.
YOUNG F. R. Cavitation[M]. Yew York: McGraw Hill,
1989.
FRANC J. P., AVELLAN F. and BELHADJI B. et al.
La cavitation. m l canismes physiques et aspects
industrielles, collection grenoble sciences[M]. Presses
Universitaires de Grenoble, 1995(in France).
BRENNEN C. E. Cavitation and bubble dynamics,
Oxford engineering sciences series 44[M]. Oxford:
Oxford University Press, 1995.
ISAY W. H. Kavitation, schiffahrts-verlag[M].
Hamburg: Hansa C. Shroedter and co., 1981(in
Germany).
LEIGHTON T. G. The acoustic bubble[M]. Acad.
Press, 1994.
PLESSET M. S. Bubble dynamics in cavitation in
real liquids[M]. Editor Robert Davies, Elsevier
Publishing Company, 1964, 1-17.
HSIAO C. T., PAULEY L. L. Numerical computation of
the tip vortex flow generated by a marine propeller[J].
ASME Journal of Fluids Engineering, 1999, 121(3):
638-645.
CHAHINE G. L. Nuclei effects on cavitation inception
and noise[C]. 25th Symposium on Naval
Hydrodynamics. St. Johns, Newfoundland and
Labrador, Canada, 2004.
CHOI J. K., CHAHINE G. L., Discrete bubble dynamics
modeling[J]. Fluent News, ANSYS, Inc., 2006, 15(3).
CHEN B., STERN F. Computational fluid dynamics of
four-quadrant marine-propulsor flow[C]. ASME
Symposium on Advances in Numerical Modelling of
Aerodynamics and Hydrodynamics in Turbomachinery. Washington, D. C., USA, 1998.

331

[15]

[16]

[17]
[18]
[19]

[20]

[21]

[22]

[23]
[24]
[25]

[26]

[27]

[28]

[29]

GORSKI J. J. Present state of numerical ship


hydrodynamics and validation experiments[J]. Journal
of Offshore Mechanics and Arctic Engineering,
124(2): 74-80.
MACINTYRE F. On reconciling optical and
acoustical bubble spectra in the mixed layer, in
Oceanic whitecaps[M]. New York: Monahan and
Macniocaill, Reidell, 1986,75-94.
OLDENZIEL D. M. A new instrument in cavitation
research: The cavitation susceptibility meter[J]. Journal
of Fluids Engineering, 1982,104: 136-142.
BILLET M. L. Cavitation nuclei measurements A
review[C]. ASME Cavitation and Multiphase Flow
Forum, 1985, FED-23: 31-38.
BREITZ N., MEDWIN H. Instrumentation for in situ
acoustical measurements of bubble spectra under
breaking waves[J]. J. Acoust. Soc. Am., 1989, 86:
739-743.
DURAISWAMI R., PRABHUKUMAR S. and
CHAHINE G. L. Bubble counting using an inverse
acoustic scattering method[J]. The Journal of the
Acoustical Society of America, 1998, 104(5):
2699-2717.
CHAHINE G. L., KALUMUCK K. M. and CHENG L.
Y. et al. Validation of bubble distribution measurements
of the ABS acoustic bubble spectrometer with high
speed video photography[C]. 4th International
Symposium on Cavitation. California, Pasadena, USA:
Institute of Technology, 2001.
CHAHINE G. L., KALUMUCK K. M. development of a
near real-time instrument for nuclear measurement: The
ABS acoustic bubble spectrometer[C]. Proceedings of
the 4th ASME-JSME Joint Fluids Engineering
Conference. Honolulu, Hawai, 2003
ARNDT R. E., KELLER A. P. Water quality effects on
cavitation inception in a trailing vortex[J]. ASME
Journal of Fluid Engineering, 1992, 114: 430-438.
GILMORE F. R. The growth and collapse of a spherical
bubble in a viscous compressible liquid[R]. California
Institute of Technology, Hydro. Lab. Rep., 26-4, 1952.
HSIAO C. T., CHAHINE G. L. and LIU H. L. Scaling
Effects on prediction of cavitation inception in a line
vortex flow[J]. Journal of Fluid Engineering, 2003,
125: 53-60.
HSIAO C. T., CHAHINE G. L. Prediction of vortex
cavitation inception using coupled spherical and
non-spherical
models
and
Navier-Stokes
computations[J]. Journal of Marine Science and
Technology, 2004, 8(3): 99-108.
HSIAO C. T., CHAHINE G. L. Scaling of tip vortex
cavitation inception noise with a statistic bubble
dynamics model accounting for nuclei size
distribution[J]. ASME Journal of Fluid Engineering,
2005, 127(1): DOI:10.1115/1.1852476.
CHAHINE G. L., KALUMUCK K. M. The influence of
gas diffusion on the growth of a bubble cloud[J]. ASME
Cavitation and Multiphase Flow Forum, Cineinnati,
Ohio, 1987, 50: 17-21.
JOHNSON V. E., HSIEH T. The influence of the
trajectories of gas nuclei on cavitation inception[C].
Sixth Symposium on Naval Hydrodynamics. 1966,
163-179.

[30] HABERMAN W. L., MORTON R. K. An experimental


investigation of the drag and shape of air bubbles rising
in various liquids[R]. Report 802, DTMB, 1953.
[31] CHOI J. K., CHAHINE G. L. Non-spherical bubble
behavior in vortex flow fields[J]. Computational
Mechanics, 2003, 32(4-6): 281-290.
[32] CHOI J. K., CHAHINE G. L. Noise due to extreme
bubble deformation near inception of tip vortex
cavitation[C]. FEDSM03, International Symposium
on Cavitation Inception, 4th ASME/JSME Joint
Fluids Eng. Conference. Honolulu, Hawaii, 2003.
[33] CHOI J. K., CHAHINE G. L. A numerical study on the
bubble noise and the tip vortex cavitation inception[C].
Eighth International Conference on Numerical Ship
Hydrodynamics. Busan, Korea, 2003.
[34] CHOI J. K., CHAHINE G. L. and HSIAO C. T.
Characteristics of bubble splitting in a tip vortex flow[C].
Fifth International Symposium on Cavitation,
CAV2003. Osaka, Japan, 2003.
[35] CHOI J. K., CHAHINE G. L. Noise due to extreme
bubble deformation near inception of tip vortex
cavitation[J]. Physics of Fluids, 2004, 16(7):
2411-2418.
[36]
REBOW M., CHOI J. and CHOI J. K. et al.
Experimental validation of BEM code analysis of bubble
splitting in a tip vortex flow[C]. 11th International
Symposium on Flow Visualization. Notre Dame,
Indiana, 2004.
[37] CHOI J. K., HSIAO C. T. and CHAHINE G. L. Tip
vortex cavitation inception study using the Surface
Averaged Pressure (SAP) model combined with a
bubble splitting model[C]. 25th Symposium on Naval
Hydrodynamics. St. Johns, Newfoundland, Canada,
2004.
[38] CHOI J., HSIAO C. T. and CHAHINE G. L. et al.
Growth, oscillation and collapse of vortex cavitation
bubbles[J]. Journal of Fluids Mechanics, 2009, 624:
255-279.
[39] ALBAGLI D., GANY A. High speed bubbly nozzle
flow with heat, mass, and momentum interactions[J].
International Journal of Heat and Mass Transfer,
2003, 46: 1993-2003.
[40]
MOR M., GANY A. Analysis of two-phase
homogeneous bubbly flows including friction and mass
addition[J].
Journal
of
Fluids
Engineering
Transaction of the ASME, 2004,126: 102-109.
[41] MOTTARD E. J., SHOEMAKER C. J. Preliminary
investigation of an underwater ramjet powered by
compressed air[R]. NASA Technical Note D-991, 1961.
[42] WITTE J. H. Predicted performance of large water
ramjets[C]. AIAA 2nd Advanced Marine Vehicles and
Propulsion Meeting. 1969, AIAA Paper No. 69-406.
[43] PIERSON J. D. An Application of hydropneumatic
propulsion to hydrofoil craft[J]. Journal of Aircraft,
1965, 2: 250-254.
[44] KOREN O. Water tank testing of a laboratory two-phase
marine ramjet engine[D]. Master Thesis, Technion
Institute of Technology, 2005.
[45] VALENCI S. Parametric sea trials of marine ramjet
engine performance[D]. Master Thesis, Technion
Institute of Technology, 2006.

332

[46] TANGREN R. F., DODGE C. H. and SEIFERT H. S.


Compressibility effects in two-phase flow[J]. Journal of
Applied Physics, 1949, 20(7): 637-645.
[47] SCHELL Jr. et al. The hydro-pneumatic ram-jet[Z]. US
Patent 3,171,379, 1965.
[48] VARSHAY H., GANY A. Underwater two phase ramjet
engine[Z]. US Patent 5,598,700, 1994.
[49] VARSHAY H., GANY A. Underwater two phase ramjet
engine[Z]. US Patent 5,692,371, 1996.
[50] MUIR T. F., EICHHORN R. Compressible flow of an
air-water mixture through a vertical two-dimensional
converging-diverging nozzle[C]. Proc. Heat Transfer
Fluid Mechanics Institute. Stanford University Press,
1963, 183-204.
[51] ISHII R., UMEDA Y. and MURATA S. et al. Bubbly

flows through a converging-diverging nozzle[J]. Physics


of Fluids, 1993, 5: 1630-1643.
[52] KAMEDA M., MATSUMOTO Y. Structure of shock
waves in a liquid containing gas bubbles[C]. IUTAM
Symposium on Waves in Liquid/Gas and
Liquid/Vapor Two Phase Systems. 1995, 117-126.
[53] WANG Y. C., BRENNEN C. E. One-dimensional
bubbly cavitating flows through a converging-diverging
nozzle[J]. Journal of Fluids Engineering, 1998, 120:
166-170.
[54] PRESTON A., COLONIUS T. and BRENNEN C. E. A
numerical investigation of unsteady bubbly cavitating
nozzle flows[C]. Proceedings of the ASME Fluid
Engineering Division Summer Meeting. Boston, 2000.

Anda mungkin juga menyukai