You are on page 1of 7

Proceedings of FEDSM02

ASME 2002 Fluids Engineering Division Summer Meeting


Montreal, Quebec, Canada, July 1418, 2002

FEDSM2002-31011
EVALUATION OF CAVITATION MODELS FOR
NAVIER-STOKES COMPUTATIONS
Inanc Senocak and Wei Shyy
Department of Aerospace Engineering, Mechanics and Engineering Science
University of Florida
Gainesville, FL 32611
inanc@aero.ufl.edu, wss@aero.ufl.edu
ABSTRACT
The predictive capability of three transport equation-based
cavitation models is evaluated for attached turbulent, cavitating
flows. To help shed light on the theoretical justification of these
models, an analysis of the mass and normal-momentum
conservation at a liquid-vapor interface is presented. The test
problems include flows over an axisymmetric cylindrical body
and a planar hydrofoil at different cavitation and Reynolds
numbers. Proper grid distribution for high Reynolds number
cavitating flows is emphasized. Although all three models give
satisfactory predictions in overall pressure distributions,
differences are observed in the closure region of the cavity,
resulting from the differences in compressibility characteristics
handled by each model.
Keywords: CFD, cavitation, turbulence
INTRODUCTION
Navier-Stokes computations of turbulent cavitating flows
have received a growing attention due to advances in
computational capabilities and physical understanding for these
problems. The cavitating flow problem is complicated due to
high Reynolds number and multiphase nature of the flow with
large disparity between fluid properties of each phase. The
inception and concurrent development of a cavity into other
stages of cavitation, namely sheet, cloud, vortex and
supercavitation, is driven by the phase change due to
hydrodynamic pressure drop and bubble dynamics [1]. To date,
no computational model can offer comprehensive capabilities of
all these flow regimes based on first principles.
Due to changes in fluid properties and physical
mechanisms across the liquid-vapor boundaries, conventional
computational algorithms of single-phase incompressible flow
experience severe convergence and stability problems for
cavitating flows. To remedy this situation, improved numerical
methods have been proposed. In the context of density-based
methods, the artificial compressibility method has been applied

with special attention given to the preconditioning technique [25]. The preconditioning technique is guided by examining the
eigenvalues of system and may not be unique depending on the
forms of the governing equations [6]. Following the spirit of the
well-established SIMPLE algorithm [7], a pressure-based
method for turbulent cavitating flows has been developed [8]. It
is shown that the pressure correction equation for turbulent
cavitating flows shares common features of high speed flows
and a pressure-density coupling scheme is developed that
results in a unified incompressible-compressible formulation.
Upwinded density interpolation is adopted to improve mass and
momentum conservation in cavitating regions [8]. Highresolution non-oscillatory convection schemes have been
applied in both pressure-based and density-based methods [2-5,
8]. The use of such convection schemes is very important
especially in the vicinity of sharp density gradients. Both
methods, pressure-based and density-based, have been
successful to compute turbulent cavitating flows around
axisymmetric bodies and hydrofoils with comparable accuracy.
A common approach in cavitation modeling is to use the
homogeneous flow theory. In this theory, the mixture density
concept is introduced and a single set of mass and momentum
equations are solved. Different ideas have been proposed to
generate the variable density field; a review of such studies is
given in Ref. [9]. Some of the existing studies solve the energy
equation and determine the density through suitable equations
of state [3]. Since most cavitating flows are isothermal,
arbitrary barotropic equations have been proposed to
supplement the energy consideration [10]. Another popular
approach is the transport equation-based model (TEM) [2, 4-6,
8]. In TEM, a transport equation for either mass or volume
fraction, with appropriate source terms to regulate the mass
transfer between phases, is solved. Different source terms have
been proposed by different researchers, which will be discussed
in the next section. A recent experimental finding [11] helps
assess the adequacy of the above-mentioned physical models. It

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 07/13/2015 Terms of Use: http://asme.org/terms

Copyright 2002 by ASME

section. A recent experimental finding [11] helps assess the


adequacy of the above-mentioned physical models. It shows
that vorticity production is an important aspect of cavitating
flows, especially in the closure region [11]. Specifically, this
vorticity production is a consequence of the baroclinic
generation term of vorticity equation.
1
(1)
P

Clearly, if an arbitrary barotropic equation = f (P) is used,


then the gradients of density and pressure are always parallel;
hence the baroclinic torque is zero. This suggests that physical
models that utilize a barotropic equation will fail to capture an
experimentally observed characteristic of cavitating flows.
Likewise, solving an energy equation will also experience the
same situation if the flow is essentially isothermal. However in
TEM approach the density is a function of the transport
process. Consequently, gradients of density and pressure are
not necessarily parallel, suggesting that TEM can accommodate
the baroclinic vorticity generation.
Different modeling concepts have been introduced in TEM,
with inconsistent numerical treatments such as grid resolution
and no consideration of pressure-density dependency. As a
result, there is a lack of clear consensus of the capability and
relative merits of these models. The goal of the present study is
to assess the predictive capability of existing transport
equation-based models using a unified pressure-based method
for turbulent cavitating flows [8], with an emphasis on grid
resolution. Three models [(2, 12), 4, 13] are considered in this
study. A common characteristic of these models is the use of
user-defined empirical factors to regulate the mass transfer
process. Although these empirical factors may seem ad-hoc,
satisfactory results for different geometries and flow conditions
have been obtained with consistent modeling parameters [8, 14].
Nevertheless, it is desirable to gain insight into the meaning of
these empirical parameters based on the first principles. In this
study, an analysis based on the interfacial dynamics is
presented to shed light on the nature of the modeling
parameters.
In what follows, each cavitation model is shortly
summarized along with the empirical parameters used in the
computations. Following this, a derivation, starting from the
mass and momentum conservation at a liquid-vapor interface is
presented to explain the modeling concept. The first set of
results cover the empirical cavitation models tested for the
cases of flow over an axisymmetric cylindrical object with
hemispherical forehead and a NACA66MOD hydrofoil.
NUMERICAL METHOD
The present Navier-Stokes solver employs a pressurebased algorithm and the finite volume approach. The governing
equations are solved on multi-block structured curvilinear grids
[15, 16]. To represent the cavitation dynamics the transport
equation models described in the upcoming section are

adopted. The resulting system of equations for turbulent


cavitating flows is solved using the pressure-based method
developed by Senocak and Shyy [8]. As already mentioned, a
key feature of this method is to reformulate the pressure
correction equation into a convective-diffusive, instead of a
pure diffusive, equation. This modification is achieved through
the inclusion of a pressure-velocity-density coupling scheme
into the pressure correction equation. For details of the
numerical algorithm, the reader is referred to Ref. [8]. The
original k- turbulence model with wall functions is employed as
the turbulence closure [17].
SUMMARY
OF
TRANSPORT
EQUATION-BASED
MODELS (TEM)
Model-1 (from Refs. [2, 12])
Several researchers have adopted this model [2, 5, 12]. Both
volume fraction and mass fraction forms of it have been
adopted. Evaporation and condensation terms are both
functions of pressure. The liquid volume fraction form is
considered in this study.
L
r
C MIN ( PL PV , 0) L
+ ( L u) = dest L
+
t
V (0.50 LU 2 ) t
(2)

C prod MAX ( PL PV , 0)(1 L )


( 0.50 LU2 ) t
The empirical factors have the following values (Cdest=1.0,
Cprod=8.0x101). These parameters have been non-dimensionalized
with free stream values to have the correct dimensional form as
the convective terms. The same parameters can be used for
different geometries and flow conditions provided that they are
non-dimensionalized with the free stream values. The
discussion holds for the other cavitation models also.
Model-2 (from Ref. [4])
The liquid volume fraction is chosen as the dependent
variable in the transport equation. The evaporation term is a
function of pressure whereas the condensation is a function of
the volume fraction.
L
r
C MIN ( PL PV , 0) L
+ ( L u ) = dest V
+
t
(0.50 LU 2 ) L t
(3)

C prod L2 (1 L )

L t
The empirical factors have the following values, which have
been determined previously [8]. (Cdest=9.0x105, Cprod=3.0x104)
Model-3 (from Ref. [13])
The vapor mass fraction is the dependent variable in the
transport equation. Both evaporation and condensation terms
are functions of pressure. The model equations that are
presented here are slightly different from the ones in the original
paper [13], which considers only a non-condensable gas. No
such a limitation is imposed here.
( m fV )
r
(4)
+ ( m fV u) = ( m
& +m
& +)
t

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 07/13/2015 Terms of Use: http://asme.org/terms

Copyright 2002 by ASME

m& = Cdest

U
2 P P
L V V

3 L

if : P < PV

The final form reads the following after further arranging the
terms.
(5)

U
2 P PV
LV
if : P > PV

3 L
1
f
(1 fV )
(6)
= V +
m V
L
The empirical factors have the following values (Cdest/=1225,
Cprod/=36750).
m& + = C prod

DERIVATION OF AN EMPIRICISM-FREE CAVITATION


MODEL
The derivation starts by considering a liquid-vapor
interface. The mass and normal momentum conservation at such
an interface can be written as follows.
(7)
L (VL ,n VI ,n ) = V (VV ,n VI ,n )

V
V
1
1
PV PL = + + 2V V ,n 2 L L,n
(8)
n
n
R1 R2
+ L (V L ,n V I ,n ) 2 V (VV ,n V I, n ) 2
Figure 1 illustrates a typical representation of a liquid-vapor
interface based on the homogeneous flow theory. The mixture
density is defined as follows based on the liquid volume
fraction [18].
(9)
m = L L + V (1 L )
Assuming that a hypothetical interface lies in the liquid-vapor
mixture region, as illustrated in Fig. 1, and neglecting the surface
tension and viscosity effects, the mass and normal momentum
conservation conditions reduces to the following forms.
(10)
V (VV ,n VI ) = m (Vm ,n V I )

PV PL = m (Vm ,n V I )2 V (VV ,n V I )2

(11)

From the mass conservation condition, Eq. (10), the following


relation can be obtained:

(V m,n VI ) =

V (VV ,n V I )
m

(12)

The momentum conservation condition, Eq. (11), can be


rearranged to the following form by incorporating the mass
conservation condition, Eq. (10), and Eq. (12).

(13)
PV PL = V (VV ,n V I ) 2 V 1
m
At this point the definition of mixture density, given in Eq. (9), is
incorporated into the above equation that leads to the following
forms:

(14)
PV PL = V (VV ,n V I ) 2
1

(
1

)
L L
V
L

( V L ) L =

L =

( PV PL ) L L + ( PV PL ) V (1 L )
V (VV ,n V I ) 2

L ( PL PV ) L
( PL PV )(1 L ) (16)
+
2
V (VV ,n V I ) ( L V ) (VV ,n V I ) 2 ( L V )

The above equation defines the value of liquid volume fraction.


In the context of TEM, it is necessary to couple the above
interfacial condition as a source term into the transport equation
of liquid volume fraction. For this purpose, Eq. (16) is simply
normalized with a characteristic time scale so that a
dimensionally correct form that represents the rate of generation
of L appears.

L ( PL PV ) L
S& = L =
t
(VV ,n V I ) 2 ( L V ) t
1V44
44
4244444
3
(17)

( PL PV )(1 L )
+
2
(VV ,n V I ) ( L V ) t
1444424444
3

The above source term is coupled to the transport equation of


L as shown below.

L
r
L ( PL PV ) L
+ ( L u ) =
t
V (VV ,n V I ) 2 ( L V ) t

(18)
( PL PV )(1 L )
+
(VV ,n V I ) 2 ( L V ) t
No empirical constants appear in the above equation. The first
term on the right hand side, compared to the second term, is
scaled naturally by a factor of the nominal density ratio ( L/ V).
To utilize the above equation, the vapor phase velocity normal
to the interface (VV,n) and the velocity of the interface (VI) are
needed.
The derivation of the model is based on an existing
interface; hence, conditional statements are required on the
pressure terms in order to couple the model to the flow
computation. Cavitation inception condition suggests that
cavitation incepts once the hydrodynamic pressure drops below
the vapor-pressure value of the corresponding liquid. As seen
from Eq. (17), in the pure liquid phase (L=1) the second term
will be zero. Hence the inception condition is imposed as
minimum (MIN) function on pressure difference term of the first
term of Eq. (17). Then the desinent cavitation condition is
imposed as a maximum (MAX) function on the pressure
difference term of the second term. As a result the first term of
Eq. (17) is responsible for conversion of liquid phase to vapor
phase (evaporation) and the second term of Eq. (17) is
responsible for conversion of vapor phase back to liquid phase
(condensation). With these inclusions, the model equation to be
solved along with Navier- Stokes equations is the following:

(15)

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 07/13/2015 Terms of Use: http://asme.org/terms

Copyright 2002 by ASME

Eq. (19) forms a fundamental cavitation model for NavierStokes computations of cavitating flows. Unlike the existing
transport equation-based models in the literature, the proposed
model does not require empirical constants to regulate the mass
transfer process. It is general to handle the time-dependency
since the interface velocity (VI) is taken into account in the
model. The model requires that an interface be constructed in
order to compute the interface velocity for time-dependent
computations, as well as the normal velocity of the vapor phase.
However in sheet type of attached cavitation the cavity is often
modeled in a steady-state computation and the assumption can
be valid and does produce satisfactory results [4, 5, 8]. Hence
the interface velocity (VI) is zero for steady-state computations
and the normal velocity of the vapor phase can be computed by
taking the gradient of the liquid volume fraction [19, 20]. The
vapor phase normal velocity is the dot product of the velocity
and the normal vector.

L
(20)
VV ,n = u n
n=
L
RESULTS
Two flow configurations have been considered, namely, (i)
an axisymmetric cylindrical object with a hemispherical
forehead (referred to as hemispherical object) at a Reynolds
number of 1.36x105 and a cavitation number of 0.30, and (ii)
the NACA66MOD hydrofoil at an angle of attack of 4 with
Reynolds number of 2.0x106. It is experimentally observed that
sheet (attached) cavitation occurs for both of the geometries
under given conditions and time averaged experimental data of
pressure distribution along the surface is available [21, 22].
Hence steady-state computations are carried out in this study.
Figure 2 compares the performance of the cavitation
models for the hemispherical object at a cavitation number of
0.30.
All three models match the experimental data
satisfactorily. Differences in the performance are more
noticeable in the closure region, where the vapor phase
condenses. Figure 3 shows the corresponding density
distribution along the surface. As seen from density plots, the
liquid phase first expands and vapor phase appears uniformly
inside the cavity, then the vapor phase compresses, in a shock
like fashion, back to the liquid phase. The differences among
the three models in density profiles are significant. This implies
that each model generates different compressibility
characteristics, although they produce very similar steady-state
pressure distributions. This issue has important implications in
time dependent problems; which model produces the correct
compressibility is an open question and needs further
investigation. Figure 4 shows the distribution of density
throughout the cavity and the spanwise vorticity distribution.
The interface is captured sharply with Model-2 and Model-3
compared to Model-1, especially in the downstream region of
the cavity. As seen from the spanwise vorticity distribution, also
given in Fig. 4, there is additional generation of vorticity at the

closure region and this is primarily due to the baroclinic torque,


confirming the discussion given in Introduction.
The NACA66MOD hydrofoil case is computed at a
Reynolds number of 2x106. The turbulent boundary layer is
extremely thin at such high Reynolds numbers. Since the
original k- turbulence model along with the wall function is
adopted [17], it is important to offer spatial resolutions
consistent with the modeling requirement [23]. This requires
that the non-dimensional normal distance from the wall (y+), a
representation of the local Reynolds number, should be in the
log-law region. More details on the wall function formulation
are given in Ref. [23]. Once a cavity occurs on the surface, the
local Reynolds number decreases due to reduction in density
and the first grid point away from the wall may not be
positioned in the log layer. This issue has important
implications in both accuracy and convergence. It is found that
proper grid distribution is very important for satisfactory results
and convergence of cavitating flow computations. However
there is a trade-off between positioning the first grid point away
from the wall in the log layer verses having enough points to
discretize the cavity, especially for high Reynolds number
cases.
Two grids have been tested. Model-2 is used in the
computations. Grid-A is a three-block domain with
approximately 6.6x104 nodes whereas Grid-B is a six-block
domain with approximately 9.8x104 nodes. The grid resolution
is further clustered close to the suction side of the hydrofoil in
Grid-B compared to Grid A. Close up views of the same region
in both grids are shown in Fig. 5 to demonstrate the spatial
resolution. The extend of the cavity region is also highlighted
on the grids to give an idea of how many grid points have been
used to discretized the vapor cavity. As shown in Fig. 6 the
predictions with Grid-A are very poor compared to the case
with Grid-B. A very short cavity has been captured with Grid-A
as a result of poor resolution. The vapor cavity is so thin that
Grid-A can only accommodate two grid points in this region. In
Fig. 7 the y+ distribution of the first grid points away from the
wall is plotted along the surface. The law of the wall is also
highlighted on this plot. As can be observed in Fig.5, special
attention has been given during grid generation to place the
points of the Grid-B in the log layer. A higher resolution grid is
also utilized and it is found that if the y+ values of the cavitating
region are in the viscous sublayer, the computation is not stable
and convergent. This may be because in the wall function
formulation a linear velocity profile is imposed in the viscous
sublayer and such a profile may not be suitable to represent the
phase change dynamics. It should be emphasized that the
original k- turbulence model [17] is based on the equilibrium
assumption, and the rate of turbulence production and
dissipation are balanced in the log-layer, hence it is advisable to
place the near wall region in the log layer [24]. If higher
resolution is required a low Reynolds number two-equation
turbulence model should be the choice. Nevertheless Grid-B is
used for all other computations of NACA66MOD hydrofoil
case.
4

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 07/13/2015 Terms of Use: http://asme.org/terms

Copyright 2002 by ASME

[2] Merkle, C.L., Feng, J. Z. and Buelow, P.E.O., 1998,


Computational modeling of the dynamics of sheet
cavitation, Proc. Third Intern. Symp. on Cavitation,
Grenoble, France.
[3] Edwards, J.R. Franklin, R.K. and Liou, M.S., 2000, Lowdiffusion flux-splitting methods for real fluid flows with
phase transitions, AIAA J. 38(9), pp. 1624-1633.
[4] Kunz, R.F., Boger, D.A., Stinebring, D.R., Chyczewski, T.S.,
Lindau, J.W., Gibeling, H.J., Venkateswaran, S. and
Govindan, T.R., 2000, A preconditioned Navier-Stokes
method for two-phase flows with application to cavitation
prediction, Computer and Fluids, 29, p. 849.
[5] Ahuja, V., Hosangadi, A. and Arunajatesan, S., 2001,
Simulations of cavitating flows using hybrid unstructured
meshes, ASME J. of Fluids Engineering, 123, pp. 331-340.
[6] Venkateswaran, S., Lindau, J.W., Kunz, R.F and Merkle,
C.L., 2001, Preconditioning algorithms for the computation
of multi-phase mixture flows, AIAA 39th Aerospace
Sciences Meeting & Exhibit, AIAA-2001-0125.
[7] Patankar, S.V., 1980 Numerical Heat Transfer and Fluid
Flow, Hemisphere, Washington DC.
[8] Senocak, I. and Shyy, W., 2002, A pressure-based method
for turbulent cavitating flow simulations, J. of
Computational Physics, 176, pp. 1-22
[9] Wang, G., Senocak, I., Shyy, W., Ikohagi, T., and Cao, S.,
2001, Dynamics of Attached Turbulent Cavitating Flows,
Progress in Aerospace Sciences, 37, pp. 551-581.
[10] Delannoy, Y. and Kueny, J.L., 1990, Cavity flow
predictions based on the Euler equations, ASME Cavitation
and Multi-Phase Flow Forum.
[11] Gopalan, S. and Katz, J. 2000 Flow structure and modeling
issues in the closure region of attached cavitation, Physics
of Fluids, 12(4), p. 895.
[12] Singhal, A.K., Vaidya, N. and Leonard, A.D., 1997, Multidimensional simulation of cavitating flows using a PDF
model for phase change, ASME Fluids Engineering
Division Summer Meeting, ASME Paper FEDSM97-3272.
[13] Singhal, A.K., Li, N. H., Athavale, M. and Jiang, Y., 2001,
Mathematical basis and validation of the full cavitation
model, ASME Fluids Engineering Division Summer
Meeting, ASME Paper FEDSM2001-18015.
[14] Senocak, I., and Shyy, W. 2001, "Numerical Simulation of
Turbulent Flows with Sheet Cavitation," CAV2001 4th
International Symposium on Cavitation, Paper No.
CAV2001A7.002.
[15] Shyy, W., Thakur, S. S., Ouyang, H., Liu, J., and Blosch, E.,
1997, Computational Techniques for Complex Transport
Phenomena, Cambridge University Press, New York.
[16] Shyy, W., 1994, Computational Modeling for Fluid Flow
and Interfacial Transport, Elsevier, Amsterdam, The
Netherlands, (revised printing 1997).
[17] Launder, B. E., and Spalding, D. B., 1974, The Numerical
Computation of Turbulent Flows, Computational Methods
in Applied Mechanics and Engineering, 3, pp. 269-289.
[18] Wallis, G.B., 1969, One-dimensional Two-phase Flow,
McGraw-Hill, New York.
[19] Brackbill, J.U., Kothe, D.B. and Zemach, C., 1992, A
Continuum Method for Modeling Surface Tension, J. Of
Computational Physics, 100, pp. 335-354.
[20] Francois, M., 1998, A Study of the Volume of Fluid
Method for Moving Boundary Problems, M.Sc. Thesis,

In Fig. 8 and Fig. 9 the performances of the models have


been assessed for cavitation numbers of 0.91 and 0.84
respectively. For cavitation number of 0.84, the vaporous cavity
covers about 60% of the hydrofoil surface. All three models
produce satisfactory results at both cavitation numbers.
Differences in predictions are more pronounced, similar to the
hemispherical object case, at the closure region, which is due to
the different compressibility characteristics imposed by the
cavitation models. Model-2 and Model-3 do relatively a better
job to capture the pressure distribution at the closure region as
compared to the predictions of Model-1.
CONCLUSIONS
An improved pressure-based method for turbulent
cavitating flows is employed to evaluate the existing transport
equation-based cavitation models in literature. Three versions
of the transport equation-based model (TEM) are considered.
To help understand these models, a theoretical derivation is
presented starting from mass and normal momentum
conservation at a liquid-vapor interface. Unlike the existing
empirical cavitation models, the new model needs no user
defined empirical factors to regulate the mass transfer process.
It helps interpret the implication of the empirical parameters
required by the existing models.
Two flow configurations have been tested, including a
hemispherical projectile and the NACA66MOD hydrofoil. For
both geometries, all three models produce comparable and
satisfactory results in wall pressure distributions. However,
different density distributions, especially in the closure region
are predicted by the models, implying that the compressibility
effects are handled differently. These differences can
significantly impact the time dependent flow computations. It is
also shown that a wall-function consistent grid is required to
improve predictive capability for high Reynolds number cases.
The present results have further validated the generality of the
pressure-based algorithm for turbulent cavitating flows. The
numerical method have successfully accounted for the
compressibility effects induced by different cavitation models.
The present study suggests that the existing versions of the
TEM have merits and limitations. Further research is needed to
address the modeling uncertainties. Direct utilization of the
interfacial dynamics, such as that presented here, should be
further pursued to alleviate the need for empirically adjusting
modeling parameters with a sound theoretical foundation. We
will further pursue this path and report our findings in the
future.
ACKNOWLEDGMENTS
This study has been supported partially by ONR and NSF.
The first author would like to thank Marianne Francois for
helpful discussions on computation of normal velocities.
REFERENCES
[1] Knapp, R.T., Daily, J.W. and Hammitt, F.G., 1970,
Cavitation, McGraw-Hill, New York.

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 07/13/2015 Terms of Use: http://asme.org/terms

Copyright 2002 by ASME

[22]
[23]

[24]

Model-1
Model-2
Model-3

0.8

density

[21]

Hemispherical, Re=1.36x105, =0.30

Embry Riddle Aeronautical University, Daytona Beach FL,


USA.
Rouse, H., and McNown, J. S., 1948, Cavitation and
Pressure Distribution, Head Forms at Zero Angle of Yaw,
Studies in Engineering, Bulletin 32, State University of
Iowa.
Shen, Y., and Dimotakis, P., 1989, The Influence of Surface
Cavitation on Hydrodynamic Forces, Proceedings of 22nd
ATTC, St. Johns, pp.44-53.
He, X., Senocak, I., Shyy, W., Gangadharan, S.N. and
Thakur S., 2000, Evaluation of Laminar-Turbulent
Transition and Equilibrium Near Wall Turbulence Models,
Numerical Heat Transfer, Part A, 37, pp.101-112.
Versteeg, H.K. and Malalasekera, W., 1995, An
Introduction to Computational Fluid Dynamics, p. 73,
Longman, London.

0.6

0.4

0.2

0.5

1.5

2
s/D

2.5

3.5

Figure 3. Comparison of surface density distribution obtained


from empirical cavitation models. Experimental data is from
Ref. [21].

Model-1

Model-2

Figure 1. Representation of a vaporous cavity in homogeneous


flow theory.
Hemispherical, Re=1.36x105, =0.30

Model-3

1.2

Figure 4. Detailed view of the cavitating region. On the left is


the density distribution; on the right is the spanwise vorticity
distribution.

Model-1
Model-2
Model-3
Exp. data

1
0.8

Cp

0.6

0.04

0.04

GRID-A

0.4

GRID-B

0.035

0.035

0.03

0.03

0
-0.2
-0.4
0

0.2

0.025

0.025

Cavity edge
0.5

1.5

2.5

3.5

s/D

Figure 2. Comparison of surface pressure distribution obtained


from empirical cavitation models. Experimental data is from
Ref. [21]

0.02
-0.23

-0.225

-0.22

-0.215

-0.21

0.02
-0.23

-0.225

-0.22

-0.215

-0.21

Figure 5. Visual comparison of the grid distributions.

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 07/13/2015 Terms of Use: http://asme.org/terms

Copyright 2002 by ASME

NACA66MOD, Re=2.0x106, =0.91

1.5

NACA66MOD, Re=2.0x106, =0.84


1.2
1

Grid-A
Grid-B
Exp. data

Model-1
Model-2
Model-3
Exp. data

0.8
0.6

0.5

-Cp

-Cp

0.4
0.2

0
0
-0.2

-0.5
-0.4

-1

-0.6

0.2

0.4

0.6

0.8

-0.8

X/C

Figure 6. Effect of grid distribution on predictions. Model-2 is


used. Experimental data is from Ref. [22].
NACA66MOD, Re=2.0x106, =0.91

10

0.2

0.4

0.6

0.8

X/C

Figure 9. Comparison of surface pressure distribution obtained


from empirical cavitation models. Experimental data is from
Ref. [22].

Grid-A
Grid-B

defect layer

10

y+=400.0
2

Y+

10

log layer
y+=11.63

10

viscous sublayer
0

10

0.2

0.4

0.6

0.8

X/C

Figure 7. Distribution of the y+ values of the first grid points


away from the wall along the hydrofoil surface.
NACA66MOD, Re=2.0x106, =0.91
1.4
1.2

Model-1
Model-2
Model-3
Exp. data

1
0.8

-Cp

0.6
0.4
0.2
0
-0.2
-0.4
-0.6

0.2

0.4

0.6

0.8

X/C

Figure 8. Comparison of surface pressure distribution obtained


from empirical cavitation models. Experimental data is from
Ref. [22].

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 07/13/2015 Terms of Use: http://asme.org/terms

Copyright 2002 by ASME

You might also like