You are on page 1of 11

Cav03-GS-9-001

Fifth International Symposium on Cavitation (Cav2003)


Osaka, Japan, November 1-4, 2003

A Numerical Study of an Unsteady Turbulent


Wake Behind a Cavitating Hydrofoil

Qiao Qin Charles C.S. Song Roger E.A. Arndt


St. Anthony Falls Laboratory, St. Anthony Falls Laboratory, St. Anthony Falls Laboratory,
University of Minnesota University of Minnesota University of Minnesota
qinx0007@umn.edu csssong@aol.com arndt001@umn.edu

Abstract along the suction surface and collapses near the trailing edge
A detailed numerical study of an unsteady turbulent wake and the collapsed cavity in the form of a large eddy is
behind a cavitating hydrofoil is carried out. A virtual single- transported into the wake. Further reducing σ/2α (between 4
phase cavitation model with barotropic flow assumption is and 6) results in the generation of sheet cavitation near the
used to analyze the unsteady structures in the far wake leading edge of the suction side. This cavity grows and
behind a cavitating hydrofoil. The model automatically breaks off to form a cloud cavity due to the reentrant jet.
becomes weakly compressible in the region where there is When σ/2α falls below 4, the flow pattern is changed
no cavitation. Time-averaged quantities are in good significantly. The maximum cavity length exceeds the chord
agreement with those from experiments, which mean that length and the cavity oscillates between partial cavity and
the model can capture the main dynamics of complex super cavity. In this case the sheet cavity again starts from
cavitating flows. The cavity flow simulated is highly the leading edge but grows to the full chord length and
unsteady, which strongly affects the wake flow. Five types eventually sheds to the wake.
of unsteady vortex shedding mechanisms are identified in A selected case of a 2D NACA 0015 hydrofoil at 8 degree
the wake of a cavitating hydrofoil. angle of attack, cavitation number 1.0 (σ/2α=3.6), Reynolds
Number 106 will be used to investigate the details of
Introduction unsteady structures in the wake. In the case of sheet/cloud
The unsteady turbulent wake behind bluff bodies such as cavitation, at least five types of discrete vortex systems are
cylinders and flat plates has been extensively investigated identified in this research. First a large negative vortex
numerically as well as experimentally (Berger, 1972; structure containing numerous bubbles forms, develops,
Okajima, 1982; Huerre et al, 1990; Jackson, 1987; Kahawita breaks down and eventually sheds into the wake from the
and Wang, 2002). The turbulent wake generated by suction side. This is the primary vortex directly related to
streamlined bodies such as hydrofoils has been found to be the cloud cavity, which is found to be rather periodic
more interactive with the boundary layers formed on the causing a sharp peak in frequency spectrum. The second
surfaces of hydrofoil (Kornilov, 2002). It is very unsteady if vortex system is a positive vortex that is drawn from the
boundary layer separation occurs on the suction surface of a pressure side of the foil by the primary negative vortex as it
hydrofoil. But the details of flow near the trailing edge are passes the trailing edge. These two vortices form a vortex
still poorly understood (McCroskey, 1982). The unsteady pair in the wake and moves mainly in the downstream
turbulent wake behind a cavitating hydrofoil is even more direction but with a slight upward velocity component. A
complex. This paper primarily focuses on the numerical third vortex system is a secondary negative vortex sheds
investigation of the unsteady structures in the wake of a from the suction side, which appears to be weaker in
cavitating hydrofoil, especially various types of vortex strength and less regular than the primary negative vortex.
shedding mechanism. Time-averaged results will also be Its mechanism is not clearly understood. Whenever the
presented and compared to those from experiments. secondary negative vortex passes by the trailing edge it also
Cavitation can be defined as the formation of the vapor attracts a positive vortex from the pressure side to form a
phase in a liquid. The formation of individual bubbles and secondary vortex pair, which also moves downstream like
subsequent development of attached cavities, bubble clouds, the primary vortex pair. The fifth and last vortex system is
etc., makes the unsteady structures in the wake of a due to the instability of the positive vortex sheet originating
cavitating hydrofoil fundamentally different from those in in the pressure side of the hydrofoil. This last vortex system
the wake of a non-cavitating hydrofoil. A detailed consists only of positive vortices occurring between the
description and analysis of cavitation phenomena with vortex pairs, appearing like half the Karman vortex street.
different cavitation number and angle of attack can be found The mathematical model for non-cavitating and cavitating
in Arndt et al (2000) and Song and Qin (2001). It is well flows and their relationship, Large Eddy Simulation (LES)
known that cavitation inception occurs at composite and Sub-Grid Scale (SGS) modeling, and the numerical
parameter (σ/2α) of about 8.5, where σ is the cavitation method will first be summarized. The numerical results
number and α stands for angle of attack. When σ/2α is contain two parts: time-averaged quantities, which are
slightly lower than 8.5 (between 6 and 8), a bubble cavity mainly used for validation purpose and the unsteady
will first appear near the leading edge of the suction side characteristics, which are the main objective of this work. It
where there is a minimum pressure. This bubble slides down is found that the lift coefficients collapse quite nicely if

1
individual lift coefficient is normalized with its own  p  ρa 2 u 
corresponding non-cavitating lift coefficient and is plotted  ρu   2 
versus the composite parameter σ 2α , regardless of    ρu + p − τ xx 
U= , Q1 = ,
 ρv   ρuv − τ 
cavitation number and angle of attack. Finally conclusions    xy

are drawn based on these results.  ρw  ρuw − τ xz 

Governing Equations  ρa 2 v   ρa 2 w 
   
Song (1996) showed that the compressibility effect may  ρvu − τ yx   ρwu − τ zx  ,
dominate the dynamics of highly unsteady flows even when Q2 =  2  , Q =
3  ρwv − τ 
Mach number is small. Using the following equation for  ρv + p − τ yy   2 zy

sound speed a,  ρvw − τ yz  
 ρw + p − τ 
  zz
∂p  ∂a 2
= a2 , (1) ∂a 2 ∂a 2 
∂ρ  ρu + ρv + ρw 
 ∂x ∂y ∂z 
the equation of continuity for compressible fluid may be
S= 0 . (7)
written as  
 0 
∂p r  
r 0
+ V • ∇p + ρa 2∇ • V = 0 . (2)
∂t
By dimensional analysis, it can be shown that the first term Large-Eddy Simulation Approach
is proportional to M 2 St while the second term is Turbulent flows are composed of eddies. The scales of
eddies ranges from those comparable to the domain of
proportional to M 2 , where M is the Mach number and St is interest down to the Kolmogorov dissipation scale. To
the Strouhal number. The second term is negligible if Mach resolve all the scales in the flow of engineering interest
number is small, but the first term may not be negligible if remains impractical with current digital resources, and likely
the Strouhal number is large. When Mach number is small, to be for the next decade. There is evidence that large scales
both density and sound speed are nearly constant so that the of motion have vigorous interaction with the mean flow and
equation of continuity may be written as they strongly depend on the flow geometry and the nature of
∂p r the flow, while the small scales (or Sub-grid Scale, SGS) of
+ ∇ • ρ 0 a02V = 0 , (3)
∂t the motion as the production of a nonlinear interaction of
where the subscript “0” represents a reference quantity. large eddies tend to be more universal and homogeneous
Equation (3) is called the equation of continuity for weakly and have less influence on mean flows. The essence of
compressible flow, which is valid for steady as well as Large-Eddy Simulation (LES) is to explicitly resolve the
highly unsteady flows as long as the Mach number is small. large scales of the motion and, in the meanwhile, model the
Cavitation is usually highly unsteady and noisy; the liquid small scales (SGS) of the motion in the flow. The LES
flow should be regarded as weakly compressible rather than approach is becoming increasingly popular in the
incompressible. community of CFD because of this nature.
When cavitation occurs, the density drops sharply as soon as After applying a Favre-filtering operation to equation (6),
the pressure is below the critical pressure. To simulate this and neglecting the non-linearity of the viscous stresses
vaporization phenomenon, Song and He (1998) formulated a (Piomelli, 1999), an extra non-linear term occurs,
so-called virtual single-phase flow model as follows τ Sij = ρ (ui u j − ui u j ) , (8)
5
ρ = ∑ Ai p i for pε < p < pc . (4) which are called Sub-grid Scale (SGS) stresses and must be
i =0 modeled. Here quantities with overbars stand for filtered
The coefficients Ai are so chosen that the resulting pressure- quantities, or large scale quantities. The overbars will be
density curve has a desirable shape. Equation (4) smoothly dropped hereafter for convenience.
joins with the weakly compressible flow pressure-density A popularly used SGS model is the eddy-viscosity model of
relationship at the critical pressure pc the form
p − p0 = a0 (ρ − ρ 0 ) for pc ≤ p . (5) τ Sij = −2 ρυ t S ij , (9)
Since the Mach number of the flow inside a cavity may not that relates the SGS stresses τ Sij to the Favre-filtered strain-
be small, the fully compressible flow model with barotropic
assumption is used. Evaporation requires large amount of rate tensor Sij.
latent heat so that the energy equation may be needed if the In most cases the equilibrium assumption is made to further
thermodynamic aspect is of interest. However, as we focus simplify the problem and an algebraic model for the eddy
on the cavitation characteristics in the macro sense at this viscosity is obtained
time, thermodynamic effects remain excluded. υ t = (C s ∆ ) 2 | S | S ij , (10)
The general conservation form of equations of continuity where ∆ is the grid size and Cs is the Smagorinsky constant
and motion can therefore be summarized as follows and takes values between 0.1 and 0.23, depending on the
∂U r flow. This model is also known as the Smagorinsky model.
+∇•Q= S , (6)
∂t However, the Smagorinsky constant Cs must be decreased in
where, the presence of shear, near solid boundaries or in transitional
flows. This can be done by incorporating another damping

2
function. Moin and Kim (1982) proposed the following where Cr is the Courant stability factor, a numerical value
modification that is between zero and one.
υ t = (C s D f ∆ ) 2 | S | S ij , (11)
Computational Setup and Boundary Conditions
where Df is a nondimensional damping factor. This damping A three dimensional LES code is used to simulate a two
factor will only affect the turbulent viscosity near the solid dimensional NACA 0015 hydrofoil, at 6 and 8 degrees angle
wall. of attack, respectively. Only three grids are used in the span
Despite the lack of universality of the Smagorinsky constant wise direction. The flow configuration is shown in Fig. 1.
and some other defects, the eddy viscosity model is still a Since the flow close to the hydrofoil surface and the wake
very useful model for LES. To dynamically obtain the behind the foil are the main interest in the current study, the
Smagorinsky coefficient, Germano (1991) proposed a mesh structure of the computational domain deliberately
dynamic eddy viscosity model. In the dynamic model, the reflects this concern by heavily clustering the mesh close to
Smagorinsky coefficient is determined as the calculation the solid surface of the hydrofoil so that the fine mesh
progresses, based on the energy content of the smallest encloses the foil and covers the core of the wake behind the
resolved scale. This is accomplished by introducing a foil (Fig. 2).
secondary test filter whose width is larger than the grid filter The commonly used far-field boundary conditions are
width. This is accomplished by introducing a secondary test implemented at upstream and downstream boundaries. In
filter whose width is larger than the grid filter width. Two addition, the inflow stream wise velocity is assigned and the
drawbacks, however, exist behind the dynamic model. One downstream cross-sectional-averaged pressure is controlled
is that the filtering process of the second test filter is as the reference pressure. These conditions are summarized
considerably more expensive. The other is that the dynamic as follows
model is proved not dissipative enough near the wall while Upstream
Smagorinsky model is too dissipative. The dynamic model, ∂u i ∂P
therefore, goes to the other extreme, compared to the = = 0 (i=1,2,3), and u1=U0, (14)
Smagorinsky model, in this sense. A more sophisticated ∂x ∂x
SGS model needs to be investigated later on. Downstream
∂u i ∂P
= = 0 (i=1,2,3) and P = Preference . (15)
Numerical Method ∂x ∂x
The finite volume method with a 2nd order accurate A full-slip boundary condition is used at both the upper and
MacCormack’s predictor-corrector numerical scheme is lower walls as well as at the artificial boundaries at the ends
used. By integrating equation (6) over each finite volume, of the span to avoid the occurrence of viscous boundary
and using the divergence theorem, one can get layers. Although the mesh is heavily clustered near the solid
∂U 1 r r
∫∫
+ n • QdA = S . (12) surface, it is still not fine enough to resolve the thin
∂t ∀ A boundary layer in the non-cavitating scenario. A partial-slip
Here U is volume averaged quantity placed at the center of boundary condition (Song 1999) is therefore applied. When
the volume. To meet the numerical stability requirement, the cavitation occurs, however, the boundary layer becomes
time increment is limited by the following: thicker and hence the partial-slip automatically becomes no-
∀ slip. In either case, the pressure condition on foil solid wall
∆t ≤ C r r r , (13) is specified in the following way
| u • A | +a | A |
∂P
=0. (16)
∂n

y
5c

c
4c

Upstream Downstream
0 x
1.5c

Fig. 1 Computational flow configuration

3
(a) The grid system enclosing the hydrofoil

(b) (c)

Fig. 2 Grid system surrounding the NACA 0015 hydrofoil at 6 degree angle of attack
(b) A close-up near the leading edge, (c) A close-up near the trailing edge

of the composite parameter ( σ 2α ), it is found that the lift


Simulation Results coefficients collapse markedly, regardless of the cavitation
Time-averaged computational results number and the angle of attack (fig. 4).
Lift Coefficient. The lift coefficient of a 2D hydrofoil is
defined as
L
Cl = , (17)
1 2
ρU 0 c
2
where L equals the net lift force per unit span working on
the hydrofoil, U0 is the free stream velocity and c is the
hydrofoil chord length.
Different representations of calculated time-averaged lift
coefficients at 6 and 8 degree angles of attack, respectively,
are presented in figures 3 and 4. Figure 3 simply shows lift
coefficient as a function of cavitation number. Non-
cavitating lift coefficients are calculated quite accurately and
found agree very well with experimental data indicating that
the weakly compressible flow model works quite well for a σ
true single phase flow, as long as the Mach number is small. Fig. 3 Lift Coefficient as a function of Cavitation Number
Some deviations exist between calculated and measured lift Pressure Distribution. Pressure coefficient is defined as
coefficients under cavitating conditions, however. It is P − P0
interesting and somewhat surprising to note that two lift Cp = , (18)
1
coefficients collapse as long as cavitation occurs, or in ρU 02
another words, the lift coefficient is independent of the 2
angle of attack after hydrofoil cavitates. The composite where P0 is the upstream reference pressure. And the
parameter ( σ 2α ) appears to be valid for hydrofoils of cavitation number is simply defined as the negative of the
finite thickness although it is originally from cavitating flat pressure coefficient, e.g. σ = −C p .
plate theory. When the lift coefficient is normalized by its
own non-cavitating lift coefficient and plotted as a function

4
Wake Velocity Distribution. Mean velocity profiles of the
wake at various angles of attack at different downstream
locations under both non-cavitating and cavitating
conditions are presented in Figures 7 through 11. White
(1974) demonstrates that the growth of the width b of both
laminar and turbulent wakes is proportional to the square
root of the distance downstream of the trailing edge, while
the decay of the minimum velocity in the wake is
proportional to x-1/2. That is
b( x ) ∝ x , (19)
1
U ref − u ( y ) ∝ , (20)
x
σ/2α where x is the distance from the trailing edge and Uref is the
reference velocity. To follow the White’s idea, plots 7
Fig. 4 Normalized lift coefficient as a function of σ 2α through 10 are organized with the following definitions
The time averaged pressure coefficient under non-cavitating y
condition is plotted in Figure 5, together with the theoretical Y'= , (21)
data (Abbott et al, 1959). One point that needs to be stressed xc
here is that the pressure coefficient is one at the trailing edge (U ref − u ( y )) x
in Abbott, which turned out to be an error. Batcher (1967) U'= . (22)
U ref c
proved that the pressure at the trailing edge is approximately
zero instead of one. The calculated data agrees quite well Figures 7 and 8 represent typical viscous wake behind a
with theoretical data. Figure 6 is the time averaged pressure non-cavitating hydrofoil. The mean velocity distributions at
coefficients under cavitating condition. Because of the different downstream locations indeed collapse quite nicely.
nature of this cavitation model (a rather simplified density- The spread of the wake due primarily to viscosity is quite
pressure relationship instead of true equation of state), the narrow at 6 and 8 degrees angle of attack. When the
time averaged quantities, such as lift and pressure hydrofoil cavitates and large vortical structure containing
coefficients, are only good in a qualitative sense. numerous bubbles of different sizes is shed into the wake,
the spread of the wake becomes much wider than those from
non-cavitating wakes (figures 9 and 10). This is because
these clouds of bubbles extend much further in the cross
stream direction than the viscous wake associated with non-
cavitating flow (Kjeldsen et al, 2000). The experimental
data at 7 degree angle of attack are also included in figures
11 and 12 for validation. Note that the definition of positive
y experimentally is the negative direction computationally.
It is easy to observe that the mean velocity distribution
becomes systematically narrower as the flow develops
further downstream in a cavitating wake, and also the
magnitudes of the velocity deficit are greater than that of
experimental data. This is believed to be the effect of
dissolved gas. A new model that is able to account for the
dissolved gas effect has been developed and the results are
Fig. 5 Time averaged pressure distribution on the suction presented in Qin et al (2003b).
side for non-cavitating flow

Fig. 7 Calculated mean velocity profile at 6 degree angle of


Fig. 6 Time averaged pressure distribution on both sides of attack without cavitation
the foil at 8 degree angle of attack, cavitation number 0.5

5
(b) without cavitation
Fig. 11 Measured mean velocity profile at 7 degree angle of
attack (Reprinted from Arndt et al, 2002)
Fig. 8 Calculated mean velocity profile at 8 degree angle of
attack without cavitation Unsteady Structures
Unsteady Vortical Structures in the Wake and their
Evolution. To facilitate better analysis, the upstream-
specified stream-wise velocity has been removed from the
whole velocity field. Figure 14 illustrates the instance of a
shed large negative (defined as the normal direction into the
paper) vortex structure containing large amount of bubbles
in the wake near the trailing edge of the suction side (named
as type A vortex) and a positive (defined as the normal
direction out of the paper) vortex from the pressure side
(named as type B vortex) that is being induced by type A
vortex. A few hundred time steps later, a vortex pair
containing type A and type B vortices is formed in the wake
right after the trailing edge (fig. 15). Further analysis shows
that the moving velocity of this vortex pair in the stream
Fig. 9 Calculated mean velocity profile at 6 degree angle of wise direction is slower than the free stream velocity. This
attack, cavitation number 0.5 velocity difference depends on a number of factors, such as
cavitation number, angle of attack, etc. Since type A vortex
is located downstream of type B vortex, the induced
velocities tend to move the vortex pair upward in the cross
stream wise direction. As a result, the vortex pair moves
diagonally at the speed roughly equal to the free-stream
speed. But, since a finite width channel is considered, the
vortex pair eventually will lose its upward speed as it
approaches the upper wall. Another observation is that the
type A vortex is stronger than the type B vortex so that the
relative position of the pair tends to become more parallel to
the free stream as they move. The above observation is
consistent with the wider spread of cavitating wake. Figure
16 shows this trend (exactly the same scale as in figure 15).
There is also a secondary negative vortex (named type C)
formed above the suction side near the trailing edge in
Fig. 10 Calculated mean velocity profile at 8 degree angle of figures 15 and 16. This type C vortex is weaker in strength
attack, cavitation number 0.5 than type A vortex, but it also induces a corresponding
secondary positive vortex from the pressure side and
eventually forms a secondary and weaker vortex pair in the
wake (figure 17). Another observation is that the strength of
vortices becomes weaker when it moves further downstream
as shown by figure 18. As will be analyzed later, the
primary shedding vortex pair of type A and type B is rather
periodic while the secondary vortex pair of type C and type
D is less regular. The mechanism behind it is not well
understood. Between two periods of primary vortex
shedding, there is certain time period in which only positive
vortices (named type E) shed from pressure side (figure 19).
(a) With cavitation number of 1.2 This is somewhat similar to half Karman Vortex Street. Also

6
type E vortex is weaker in strength than the primary vortices, but occurs at higher frequency.

Type A

Type B

Fig. 14 Generation of large negative vortex structure on the suction side


and its induced positive vortex from the pressure side

Type A

Type B

Fig. 15 A formed primary vortex pair (type A and type B) in the wake

Type C

Fig. 16 The evolution of the vortex pair in the wake and a type C vortex

7
Type C

Type D

Fig. 17 A formed secondary vortex pair (type C and type D) in the wake

Type B

Type A

Fig. 18 An instance of weakened primary vortex pair (type A and type B)


further downstream in the wake

Type E

Fig. 19 Type E positive vortex chain formed in the wake

Type E

Fig. 20 A close-up of type E vortex chain


Spectral Characteristics. Negative vortex shedding due to its induced positive vortex shedding from the pressure side
the sheet/cloud cavitation on the suction side (type A) and (type B) is highly periodic. This is clearly reflected in the

8
time series of lift coefficient. The time series of lift
coefficient at 6 and 8 degrees angle of attack are presented
in figures 21 and 23. Their corresponding results from
spectral analysis are presented in figures 22 and 24. It is
easily seen that Strouhal number is around 0.2, if the
frequency is normalized by the cavity length and the
upstream velocity
fl
St = cav , (23)
U
where lcav is the cavity length. What is not so clear in
Figures 22 and 24 but is evident in figures 21 and 23 is that
there are secondary lift fluctuations superimposed in both
lift coefficients. This is type E vortex shedding frequency
due to the instability of positive vortex sheet. If this
frequency is normalized by the projected chord length Fig. 23 Time series of lift coefficient at 8 degree angle of
normal to the stream wise direction attack, cavitation number 0.5
fc sin α
St = , (24)
U
the Strouhal number is in the order of 0.2, which is very
close to the flow separation phenomenon found in the wake
of circular cylinder. Therefore the type E vortex can be
regarded as a half Karman Vortex Street.
The composite plot of peak frequency at various angles of
attack with different cavitation numbers are presented in
figure 25. These results are in a good agreement with those
from experiments (Kjeldsen et al, 2000).

Fig. 24 Frequency spectrum of the lift coefficient,


8 degree angle of attack, cavitation number 0.5

Fig. 21 Time series of lift coefficient at 6 degree angle of


attack, cavitation number 0.5

Fig. 25 Composite plot of peak frequency at 2 different


angles of attack
The spectral characteristic of the shedding vortex pair can be
illustrated in another way here. Each figure (26 and 27)
contains instantaneous pair of pressure/vorticity fields. The
dark color in the pressure plot stands for low pressure while
the bright color for high pressure. For the vorticity plot dark
color stands for the negative vorticity while the bright color
for the positive vorticity. The color upstream of the leading
edge pretty much stands for zero vorticity. See the legend on
the right side of each figure for detailed reference.
Fig. 22 Frequency spectrum of lift coefficient, Figure 26 illustrates an instance where the newly generated
6 degree angle of attack, cavitation number 0.5 large negative vortex structure from the suction side and its
induced positive vortex from the pressure side are about to

9
shed into the wake while the previous shed vortex pair is cavitation about 0.2, which is in a very good agreement with
ready to leave the computational domain. Clearly positive those from the experiments (Kjeldsen et al, 2000). In
vortex follows the negative vortex but both of them have addition, the vortex shedding due to the instability of the
negative pressure so that they are hardly distinguished in the positive vortex sheet is also apparent between two major
pressure plot. An instance similar to figure 26 at 6 degree vortex pair cycles, which is mainly viscous effects and very
angle of attack is presented at figure 27. similar to the wake behind the circular cylinder.
The distance between two cycles is roughly five times the
chord length, making the vortex shedding due to sheet/cloud

Fig. 26 A new large vortex structure formed after previous shedding


Upper half is the instantaneous pressure distribution and the lower part is
the corresponding vorticity, 8 degree angle of attack, cavitation number 0.5

Fig. 27 A new large vortex structure formed after previous shedding


Upper half is the instantaneous pressure distribution and the lower part is
the corresponding vorticity, 6 degree angle of attack, cavitation number 0.5

Conclusions angle of attack normalized with the value of non-cavitating


The unsteady lift and vortex structure behind a cavitating condition nearly collapses on a straight line when plotted
hydrofoil have been investigated using the virtual single- against σ / 2α .
phase cavitating model with a barotropic flow assumption. Five types of unsteady vortex shedding are identified. A
The time averaged quantities are in good agreement with pair of primary vortices consisting of a cloud cavity (type A
experimental data and theoretical result indicating that this negative vortex) and an induced positive vortex (type B) are
model can capture the major dynamics of cavitating flows. shed periodically into the wake region. Between two pairs of
A closer look at the mean velocity distributions in the primary vortices, there are secondary pairs of vortices shed
cavitating wake reveals that, without consideration of the in less periodical manner. There are also strings of positive
dissolved gas, the mean velocity distribution becomes vortices formed between primary and secondary pairs of
systematically narrower than that indicated by experimental vortices. This fifth type of vortices appears like a half
data. The time averaged lift coefficients for two different Karman vortex street.

10
The lift coefficient of the cavitating foil oscillates in a [15] Piomelli, U., 1999, Large-Eddy Simulation:
manner highly correlated with the 5 types of vortices. The Achievements and Challenges, Progress in Aerospace
primary mode of lift oscillation correlates with the primary Sciences 35(4), 335-362
pair of vortex shedding. Its Strouhal number based on the [16] Okajima, A., 1982, Strouhal Numbers of Rectangular
averaged cavity length is approximately equal to 0.2. The Cylinders, J. Fluid Mech., 123, 379-398
lift coefficient also oscillates periodically with Strouhal [17] Song, C.C.S., 1996, Compressibility boundary layer
number approximately equal to 0.2, if the projected foil theory and its significance in computational hydrodynamics,
width is used as the reference length. This oscillation Journal of Hydrodynamics, Series B, Vol. 8, No. 2, pp. 92-
correlates perfectly with the positive vortex string or the half 101.
Karman vortex street. Lift oscillation due to the secondary [18] Song, C.C.S., He J., Zhou F. and Wang G., 1997,
pair of vortices is less regular but quite significant. Numerical simulation of cavitating and non-cavitating flows
over a hydrofoil, SAFL project report no. 402, University of
Acknowledgements Minnesota
The National Science Foundation (Dr. Michael Plesniak) [19] Song, C.C.S., Qin, Q., 2001, Numerical Simulation of
and the Office of Naval Research (Dr. Kam Ng) sponsor this Unsteady Cavitating Flows, 4th International Symposium on
project. The Minnesota Supercomputing Institute (MSI), Cavitation, Pasadena, California
University of Minnesota, generously provides the [20] Song C.S.S. and Yuan, M., 1988, A weakly
computational resources. compressible flow model and rapid convergence methods,
Journal of Fluids Engineering, Vol. 110
References [21] White, F.M., 1974, Viscous Fluid Flow, McGraw-Hill,
[1] Abbott, I.H. and Doenhoff, A.E.V., 1959, Theory of New York
Wing Sections, Dover Publications, Inc. New York [22] Wu, T.Y., 1972, Cavity and Wake Flows, Annual
[2] Arndt, R.E.A., 2002, Cavitation in Vortical Flows, Review of Fluid Mech., 4, 243-284
Annual Review of Fluid Mechanics, 34, 143-175 [23] Yuan, M., 1988, Weakly Compressible Flow Model and
[3] Arndt, R.E.A., Kjeldsen, M., Song, C.C.S., Keller, A., Simulations of Vortex-shedding Flows about a Circular
2002, Analysis of Cavitation Wake Flows, Proceedings of Cylinder, PhD thesis, University of Minnesota
the Hydraulic Machinery and the Systems 21st IAHR
Symposium, September 9-12, Lausanne
[4] Arndt, R.E.A, Song, C.C.S., Kjeldsen, M., He, J., Keller,
A., 2000, Instability of Partial Cavitation: a
Numerical/Experimental Approach, 23rd Symposium on
Naval Hydrodynamics, September, Rouen, France
[5] Batchelor, G.K., 1967, An Introduction to Fluid
Dynamics, Cambridge University Press
[6] Berger E., 1972, Periodic Flow Phenomena, Annual
Review of Fluid Mech., 4, 313-340
[7] Bourgoyne D.A., Ceccio S.L., Dowling D.R., Park S.J.J.,
Brewer W., Pankajakshan R., 2000, Hydrofoil Turbulent
Boundary Layer Separation at High Reynolds Numbers,
Naval Hydrodynamic Symposium
[8] Huerre, P. and Monkewitz P., 1990, Local and Global
Instabilities in Spatially Developing Flows, Annual Review
of Fluid Mech., 22, 473-537
[9] Jackson C.P., 1987, A Finite-element Study of the Onset
of Vortex Shedding in Flow past Variously Shaped Bodies, J.
Fluid Mech., 182, 23-45
[10] Kahawita, R. and Wang P., 2002, Numerical Simulation
of the Wake Flow Behind Trapezoidal Bluff Bodies,
Computers & Fluids, 31, 99-112
[11] Kjeldsen, M., Arndt, R.E.A., Effertz, M., 2000,
Spectral Characteristics of Sheet/Cloud Cavitation, Journal
of Fluids Engineering, 122, 481-487
[12] Kornilov, V.I., Pailhas, G. and Aupoix, B., 2002,
Airfoil-Boundary Layer Subjected to a Two-Dimensional
Asymmetrical Turbulent Wake, AIAA Journal, 40(8), 1549-
1558
[13] Kubota, A., Kato H. and Yamaguch, H., 1992, A New
Modeling of Cavitating Flows: A Numerical Study of
Unsteady Cavitation on a Hydrofoil Section, J. Fluid Mech.,
240, 59-96
[14] McCroskey, W.J., 1982, Unsteady Airfoils, Annual
Review of Fluid Mechanics, 14, 285-311

11

You might also like