You are on page 1of 16

Dynamics of the supercavitating hydrofoil

with cavitator in steady flow field


Cite as: Phys. Fluids 32, 123307 (2020); https://doi.org/10.1063/5.0030907
Submitted: 26 September 2020 . Accepted: 16 November 2020 . Published Online: 04 December 2020

Chang Xu (徐畅), and Boo Cheong Khoo

ARTICLES YOU MAY BE INTERESTED IN

Experimental investigation of internal two-phase flow structures and dynamics of quasi-


stable sheet cavitation by fast synchrotron x-ray imaging
Physics of Fluids 32, 113310 (2020); https://doi.org/10.1063/5.0029963

Computational analysis of hydrodynamic interactions in a high-density fish school


Physics of Fluids 32, 121901 (2020); https://doi.org/10.1063/5.0028682

A lattice Boltzmann modeling of viscoelastic drops’ deformation and breakup in simple


shear flows
Physics of Fluids 32, 123101 (2020); https://doi.org/10.1063/5.0031352

Phys. Fluids 32, 123307 (2020); https://doi.org/10.1063/5.0030907 32, 123307

© 2020 Author(s).
Physics of Fluids ARTICLE scitation.org/journal/phf

Dynamics of the supercavitating hydrofoil


with cavitator in steady flow field
Cite as: Phys. Fluids 32, 123307 (2020); doi: 10.1063/5.0030907
Submitted: 26 September 2020 • Accepted: 16 November 2020 •
Published Online: 4 December 2020

Chang Xu (徐畅)a) and Boo Cheong Khoob)

AFFILIATIONS
Department of Mechanical Engineering, National University of Singapore, 119260, Singapore

a)
E-mail: e0348786@u.nus.edu
b)
Author to whom correspondence should be addressed: mpekbc@nus.edu.sg

ABSTRACT
Maintaining stability remains a crucial issue for the safety of underwater vehicles, especially during high-speed navigation where flow cavita-
tion may occur. Cavitators are small protrusions on the hydrofoil surface, which can be used to control the patterns of flow cavitation. In this
study, we investigate the effect of cavitators on supercavitating flow and hydrodynamic forces for high-speed hydrofoil. The volume of fluid
method and the large eddy simulation turbulence model with the Kunz cavitation model are used in the simulations in order to accurately
capture the interface between two phases (water and vapor) and the region of flow supercavitation. To validate the numerical method, the
time-averaged simulated results of the supercavitating flow over a wedge-shaped hydrofoil are compared and validated against the experi-
mental data by Kermeen (“Experimental investigations of three-dimensional effects on cavitating hydrofoils,” Technical Report No. 47-14,
Engineering Division of the California Institute of Technology, Pasadena, CA, 1960). Then, the numerical method is used to simulate the
supercavitating flow of two-dimensional and three-dimensional cases with and without the cavitator placed on the lower side of the hydrofoil.
The simulations were conducted for the cavitator located at the 1/32, 1/16, 1/8, and 1/4 chord of the hydrofoil at various angles of attack
1○ –12○ and cavitation numbers at 0.1, 0.2, and 0.4. A detailed analysis is made to examine the relations between cavitation pattern, hydro-
dynamic forces, and cavitator placement location. Compared to the cases without the cavitator, the supercavitating flow over the hydrofoils
controlled by the cavitator demonstrates a significant difference in lift, whereas the drag does not change much. The changes in lift are closely
related to the cavitator location. These findings can serve as a good guidance on how to improve the hydrodynamic performance and stability
condition of a high-speed hydrofoil/wing using the cavitator.
Published under license by AIP Publishing. https://doi.org/10.1063/5.0030907., s

I. INTRODUCTION hydrofoil surface in contact with water, it is necessary to maintain


good stability during navigation and takeoff. The lift generated by
Cavitation occurs when the submerged object moves suffi- the hydrofoil underwater will invariably change due to the large fluc-
ciently fast in water, and the flow cavitation can be one of the tuation especially in a high sea state and may cause the craft to veer
most important flow phenomena for the high-speed underwater violently or even flip over.7,8
object/vehicle. Based on the development of cavitation shape, the The interest on supercavitating flows is to a large extent linked
cavitation phenomenon can be broadly classified into incipient cav- to the development of increasing high-speed underwater vehicles
itation, sheet cavitation, cloud cavitation, and flow supercavitation and the requirement to overcome the drag.9 The friction drag
as the cavitation number changes from high to low.1,2 The unsteadi- reduction in the supercavitating flow over the underwater vehi-
ness of the supercavitating/cloud cavitating flow with phenomena cle is provided by replacing liquid water with water vapor and/or
such as cavity breaking, shedding, and collapsing may cause dam- air in the supercavity, which may envelop the underwater vehi-
age to the model structure, such as noises, erosion, vibration, and cle wholly or partially to enable a large reduction in drag. It is
instability during cruise.3,4 For a high-speed underwater vehicle with difficult, however, to maintain a steady natural supercavitating
submerged hydrofoil or even Wing-In-Ground (WIG) craft with the flow, so artificial ventilation8 and cavitator9 are employed to help

Phys. Fluids 32, 123307 (2020); doi: 10.1063/5.0030907 32, 123307-1


Published under license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

maintain the supercavity. Typically, this may lead to bubble layer


drag reduction (BLDR),10,11 air layer drag reduction (ALDR),12–14
and, of course, artificially ventilated supercavitation.15 The passive
control method in supercavitating flow employs different cavitators
to control the cavitation phenomena.16–20 The generated supercav-
ity is highly affected by the shape and the location of the added
cavitator. For the artificial ventilation, non-condensable gas injected
into the cavity can also weaken or eliminate the unsteady phenom-
ena such as cavity shedding or collapsing to protect the structure of
the vehicle. Although supercavitating flow is much more stable than
cloud cavitating flow, it also involves complex unstable factors such
as gas leakage, cavity shedding, and collapsing at the cavity closure, FIG. 1. Schematic of the supercavitating flow around the NACA0012 foil with the
which need to be addressed21,22 for eventual implementation. cavitator placed at the lower side. O is the reference point (0, 0), c is the chord
The cavitating flow that surrounds a model can be studied by length, α is the angle of attack, U∞ is the free stream velocity, l is the cavity
various experimental and numerical approaches. The most com- length, and t is the maximum cavity thickness. The cavitator on the lower side of
the hydrofoil is a 0.02 m × 0.005 m rounded rectangular protrusion (exaggerated
monly used experimental systems are the water tunnel and water and not to scale).
tank launching test, with the data acquisition system such as high-
speed camera, particle image velocimetry (PIV),23–25 and x-ray.26–28
The computational fluid dynamics (CFD) approach is also widely
employed to study this phenomenon by solving the Navier–Stokes
simulations,48 which serves to validate the accuracy of the numerical
(NS) equations with the cavitation and turbulence model. Cavita-
method.
tion models such as Kunz, Schnerr–Sauer, and Zwart are deemed
In addition, to be able to compensate the changes in hydrody-
the key factors introduced to calculate the mass transfer between
namic forces during high-speed navigation in high sea states, there is
liquid water and water vapor phases29–31 in the cavitation study.
an inadvertent need of flow control. To achieve this, we propose the
Since cavitation occurs at a high speed, the flow is usually highly
employment of the cavitator on the lower side of hydrofoil to control
turbulent. The Reynolds-averaged Navier–Stokes (RANS) equations
the formation of the supercavity enveloping the vehicle. The changes
and large-eddy simulation (LES) are most commonly used meth-
in hydrodynamic forces due to cavitator are closely related to the
ods in dealing with turbulence. To accurately estimate the turbu-
location of placement along the chordwise direction of the hydrofoil.
lent viscosity inside the cavity, physical modifications are usually
We evaluate the efficacy and dynamics of the cavitator placed at var-
necessary for RANS approaches, such as modified renormalization-
ious distances at the 1/32, 1/16, 1/8, and 1/4 chord of a NACA0012
group k–ε turbulence model,32–36 filter-based k–ε model,37 and par-
foil along the span. The shape of the cavitator is a 0.02 m long
tially averaged Navier–Stokes method.38–40 Compared with RANS,
× 0.005 m high rounded rectangular protrusion (as shown in
LES performs better, in general, in capturing the considerable
Figs. 1 and 2, exaggerated and not to scale). The reference point
details of vortex-related cavity structures in the flow field with rea-
O and the parameters include the location of the cavitator, the
sonably high accuracy and is adopted more prevalently in recent
cord length (c), the angle of attack (α), and the free stream veloc-
years.38,41–46
ity (U ∞ ), which are shown in Fig. 1. The supercavitating flow
Cavitators are small bumps/protrusions on hydrofoils. The pat-
around the hydrofoil with and without the cavitator is investigated.
tern of cavitating flow and the hydrodynamic forces can change
severely in the presence of cavitator. However, there is a dearth of
work on the relation between changes in hydrodynamic forces and
arrangement/location of the cavitator for supercavitating flow. In
this work, we propose a new flow control method by introducing
the flow cavitator to be placed specifically on the lower side of the
hydrofoil. In order to establish the relation between flow supercavi-
tation, hydrodynamic forces, and location of the cavitator, numeri-
cal simulations are conducted for hydrofoils with various cavitator
configurations. Therefore, by studying how the shape of the cav-
itating region changes in the presence/absence of the introduced
cavitator on the hydrofoil surface affecting the incipient cavitation
and its extent is the focus of the present study to evaluate its effi-
cacy as a flow control approach/means. The findings can serve as a
good guidance on how to improve the hydrodynamic performance
and stability condition of the high-speed hydrofoil/wing with the
application of the cavitator. The volume of fluid (VOF) method
and the LES turbulence model with the Kunz cavitation model
are used in the simulation to capture the detailed flow patterns
FIG. 2. A 3D exaggerated view of the cavitator placed at the lower side of the
and the interface of the multi-phase flow. The simulation results
hydrofoil (not to scale). The chord length is 1 m.
are compared with the previous experimental data47 and previous

Phys. Fluids 32, 123307 (2020); doi: 10.1063/5.0030907 32, 123307-2


Published under license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

The dynamics of the flow supercavitation with and without the cav- In this work, the phase change between liquid and vapor is consid-
itator are carefully analyzed at different cavitation numbers (σ) (0.1, ered, and the mass transfer rate ṁ = ṁ+ − ṁ− is solved by the Kunz
0.2, and 0.4) and angles of attack (AOA) (1○ –12○ ) while always mass transfer model,14
keeping in view the overall effect on the profile of the cavitating
Cv ρv αmin[0, p̄ − pv ]
flow and the total lift/drag coefficient. It is found that the cavita- ṁ+ = , (7)
tor could significantly change the pressure distribution on the lower (1/2ρl U 2 ∞ )t∞
surface of the hydrofoil and thus a corresponding change in the cav-
ity length and force coefficients. In a certain range, the lift coefficient Cc ρv α2 (1 − α)
has strong correlation with the position of the cavitator, whereas the ṁ− = , (8)
t∞
drag coefficient almost remains unchanged.
The rest of this paper is organized as follows. Section II presents where ṁ+ and ṁ− are the mass transfer rate of evaporation and
the numerical method: Sec. II A presents the governing equations condensation. U ∞ is the speed of the hydrofoil. t ∞ = c/U ∞ is the
and LES approach, Sec. II B presents the numerical setups, and characteristic time, where c is the chord length of the hydrofoil. In
Sec. II C presents the validation of the numerical methods. this model, the empirical constants for the different phase trans-
Section III discusses and analyzes the simulated results: Sec. III A fer rate are set as Cv = 1000 and Cc = 1000. These settings have
presents the force analysis of the two-dimensional (2D) supercav- been discussed and found to work well for a variety of fluids and
itating hydrofoil, including the hydrodynamic coefficient changes devices.50–52
with the cavitator position. Section III B presents the discussion of The LES model53 is applied to solve for the turbulent flow in this
the mechanics of the cavitator and its effect on the force coefficients. study. The instantaneous Navier–Stokes equations are solved for the
Section III C presents the simulated results at other higher cavitation large-scale eddies, and the sub-grid scale (SGS) model reflects the
numbers. Section III D presents the hydrofoil with or without the small-scale eddies. LES equations are derived from Eqs. (1)–(3) by
cavitator under the three-dimensional (3D) effect. The concluding applying a Favre-filtering operation,
remarks are found in Sec. IV.
1 1
∇ ⋅ ⃗v = ( − )ṁ, (9)
ρl ρv
II. NUMERICAL METHOD

A. Governing equations and Kunz cavitation model (ρ⃗v) + ∇ ⋅ (ρ⃗v ⃗v) = −∇p̄ + ∇ ⋅ (S̄ − B). (10)
∂t
The incompressible Navier–Stokes equations of the mixture Here, B is the influence of the small-scale eddies or the sub-grid
flow are extensively used for solving the liquid water/water vapor stress tensor, which is defined as
multiphase flow problems. The continuity and momentum equa-
tions are established as follows: B = ρ(⃗v⃗v − ⃗v ⃗v). (11)
1 1
∇ ⋅ ⃗v = ( − )ṁ, (1) Based on the Boussinesq equation,54 the sub-grid dynamic viscosity
ρl ρv μsgs is considered. Then, the sub-grid stress can be computed from

∂ B = −2μsgs D̄. (12)


(ρ⃗v) + ∇ ⋅ (ρ⃗v⃗v) = −∇p + ∇ ⋅ S, (2)
∂t
As such, the entire viscous term becomes (S̄ − B) = 2(μ + μsgs )D̄ and
where the sub-grid dynamic viscosity μsgs needs to be solved. In the k − μ
S = 2μD. (3) sub-grid scale model, it is modeled by
3
Here, ρl and ρv are the liquid water and water vapor density, p ksgs
2
∂ksgs μ + μsgs μsgs
is the pressure, ṁ is the mass transfer rate (from vapor to liq- + ∇ ⋅ (ksgs v̄) = ∇ ⋅ [ ∇ksgs ] + 2 D̄D̄ − Ce , (13)
uid water), which is modeled by mass transfer cavitation models, ∂t ρ ρ Δ̄
D = 12 (∇⃗v + ∇⃗vT ) is the strain rate tensor, and μ is the dynamic √
laminar flow viscosity of the mixture, which can be defined by using μsgs = Ck Δ̄ ksgs , (14)
the VOF approach with the liquid volume fraction,49 √
where ksgs is the SGS turbulence kinetic energy, Δ̄ = 3 Δx Δy Δz is the
μ = (1 − αv )μl + αv μv , (4) spatial filter size, Ce = 1.048, and Ck = 0.094.
where α is the volume fraction of the different phases. The subscripts
l and v represent the liquid water and water vapor, respectively. The B. Simulation setups
density ρ of the mixture is defined as
In this study, we conducted a series of simulations for 2D and
ρ = (1 − αv )ρl + αv ρv . (5) 3D hydrofoil with different cavitator arrangements. The hydrofoil
The transport equation of the water vapor volume fraction is employed is the NACA0012 foil with the cavitator incorporated
on the lower side of the foil along the chordwise direction. The
∂αv ṁ chord length of the foil is 1 m. For 3D foil, the aspect ratio is 4.
+ ∇ ⋅ (⃗vαv ) = . (6)
∂t ρv Figures 3(a) and 3(b) show the 28 × 10 m2 in 2D and 28 × 10 × 12 m3

Phys. Fluids 32, 123307 (2020); doi: 10.1063/5.0030907 32, 123307-3


Published under license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

so as to be able to capture the interface between the liquid water and


vapor phases accurately. The total mesh size is about 3 × 106 with
good orthogonality.
The pressure-based solver interPhaseChangeFoam of the open
source toolbox OpenFOAM55 is employed in this study, which con-
siders the phase change between two incompressible, isothermal,
and immiscible fluids by using a PIMPLE algorithm. The transient,
first order Euler scheme was chosen. During the calculation, the
value of the y+ parameter on the foil wall is set to be less than 5. The
time step is determined by the maximum Courant number, which is
set to be less than 1.

C. Validation of the numerical method:


Supercavitating wedge-shaped hydrofoil
The numerical method employed is validated and verified in
comparison with the water tunnel experimental results of Ker-
meen.47 A high-speed water tunnel experiment investigation on
the characteristics of the supercavitating flow over a wedge-shaped
hydrofoil at different σ, AOA, aspect ratio, and shape of the mod-
els are described in Ref. 47. In the experiment, the wedge-shaped
hydrofoil was tested in a water tunnel with one end fixed on the
side wall. Figure 5 shows that the half span of the hydrofoil (bw ) is
FIG. 3. Computational domain and boundary conditions include velocity-inlet, 0.0762 m and the chord length (cw ) is 0.0381 m. The wedge angle
pressure-outlet, symmetry, sides (slip-wall), and foil (nonslip-wall) for (a) 2D and is 6○ . Based on Kermeen’s experimental data, Vernengo and Bon-
(b) 3D cases.
figlio48 presented a new viscous lifting-line numerical method for
simulating two- and three-dimensional supercavitating hydrofoils
by introducing variable positions of the collocation points where
the body boundary condition is enforced. The original mesh size
in 3D computational domains for the numerical calculation. Bound-
is about 1.5 × 106 . We also conducted the simulation cases with a
ary conditions have been imposed at the computational domain
refined mesh size of about 4.5 × 106 to verify the suitability of the
walls perpendicular to the hydrofoil span as shown in this figure with
original mesh size for the simulation. Both the original and refined
uniformed velocity-inlet, pressure-outlet, symmetry, sides, and foil.
meshes use the same mesh plan as mentioned in Sec. II B. The cl ,
A non-slip wall condition has been used at the hydrofoil surface.
cd , and lift to drag ratio (L/D) results with the original and refined
Constant atmospheric pressure and zero velocity gradient condi-
mesh plan at σ = 0.2 with angles of attack varying from 1○ to 12○
tions have been used at the outlet boundary. Slip wall conditions
are compared with the experiment of Kermeen47 and the simulation
have been imposed at the sides of the boundary.
results of Vernengo et al.,48 as shown in Figs. 6–8, respectively. The
The typical mesh layout of the flow field surrounding the foil
is shown in Fig. 4. An unstructured tetrahedral mesh is used in the
calculation. The minimum and maximum face size of the mesh is
set as 2.5 × 10−3 m and 1 m, respectively. The first layer height is
2 × 10−4 m with 12 layers of inflation. The mesh is refined near the
leading edge and trailing edge of the foil and the cavitator. A higher
mesh resolution around the foil and downstream wake is established

FIG. 5. Wedge model in the water tunnel experiment (side view/plan view). Here,
FIG. 4. Mesh plan of the flow field surrounding the foil. The mesh layout near the cw = 0.0381 m is the cord length and bw = 0.0762 m is the half span. The wedge
leading edge and cavitator is enlarged in this figure. angle is 6○ . The angle of attack convention is shown.

Phys. Fluids 32, 123307 (2020); doi: 10.1063/5.0030907 32, 123307-4


Published under license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

Figures 6–8 show that the simulated and experimental results


agree well with each other in both trend and magnitude, which
validate and verify the present numerical method. After verifying
the mesh independence of the simulation method, the simulation
results of the original mesh are used for further analysis and dis-
cussion. Results show that the mesh plane is fairly stable. The sim-
ulation results of small cell sizes remain fine and are used in the
following contents. Here, cl and cd are the lift and drag coefficients
defined as
L
cl = 1 2
, (16)
2
ρU ∞ bw cw

FIG. 6. Comparison of the numerical results and the experimental data of the lift D
coefficient vs angles of attack at σ = 0.2.
cd = 1
. (17)
2
ρU∞ 2 bw cw
The results of Vernengo et al. show relatively higher discrepancies
at high AOA particularly on the cl at AOA >8○ . It may be noted that
at these higher AOA, the vortex wake that surrounds the hydrofoil
tip is quite strong. With the validation and verification, we proceed
to calculate further 2D and 3D cases of the flow over the NACA0012
foil with and without the presence of the cavitator placed at different
locations along the chord length.

III. RESULTS AND DISCUSSION


A. Force analysis on the 2D supercavitating hydrofoil
with/without cavitator
The force analysis on the 2D supercavitating hydrofoil with or
without the cavitator is conducted to show how the force coeffi-
FIG. 7. Comparison of the numerical results and the experimental data of the drag cients vary with the location of the cavitator on the lower side of the
coefficient vs angles of attack at σ = 0.2. NACA0012 foil. The cavitator is a rounded rectangular section mea-
suring 0.02 m long × 0.005 m high (protrusion), and the nominal
chord length of the NACA0012 foil is 1 m. The 2D NACA0012 foil
cavitation number σ can be calculated using the following formula: at AOA = 7○ is calculated first with no cavitator, and then separately,
p∞ − pv the cavitator is placed at the 1/32, 1/16, 1/8, and 1/4 chord at σ = 0.1
σ= 1
, (15) and 0.2. Figures 9 and 10 show the simulated results of the lift/drag
2
ρU∞ 2
coefficient with respect to the placement location of the cavitator.
where p∞ is the standard atmospheric pressure, pv = 2300 Pa is the Here, x indicates the location of the cavitator, and x/c = 0 stands
saturated vapor pressure (SVP), ρ = 1000 kg/m3 is the liquid water for the no-cavitator case. The results show that the lift coefficient
density, and U ∞ is the free stream velocity.

FIG. 8. Comparison of the numerical results and the experimental data of the L/D FIG. 9. Lift coefficient varies with the location of the cavitator placed at the lower
ratio vs angles of attack at σ = 0.2. side of the 2D NACA0012 foil at σ = 0.1 and 0.2. AOA = 7○ .

Phys. Fluids 32, 123307 (2020); doi: 10.1063/5.0030907 32, 123307-5


Published under license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

FIG. 10. Drag coefficient varies with the location of the cavitator placed at the lower FIG. 12. Drag coefficient varies with the location of the cavitator placed at the lower
side of the 2D NACA0012 foil at σ = 0.1 and 0.2. AOA = 7○ . side of the 2D NACA0012 foil at AOA = 1○ , 3○ , 5○ , 7○ , and 12○ . σ = 0.2.

is significantly affected by the cavitator placed at the lower side of up to, say, 1/32 (or even 1/16) chord, cl indicates a relatively mild
the foil, whereas the change in the drag coefficient is relatively much decreasing trend to a minimal change. Beyond the 1/16 chord, as
small and limited to the range of less than 40%, as compared to the the cavitator moves toward the trailing edge, cl exhibits a markedly
case without the cavitator. For both σ = 0.1 and 0.2, the trends of increasing trend with a peak value realized when the cavitator is
cl and cd with respect to the location of the cavitator are reasonably placed at the 1/4 chord. The mentioned peak value is found to range
similar. The lift coefficient slightly decreases when the cavitator is from 1.5 to 9.5 times higher than the corresponding case without
placed very near the leading edge of the hydrofoil. Then, cl increases the cavitator at the same AOA. Whatever, the change in cl at the
almost linearly as the cavitator is moved closer to the trailing edge cavitator location of the 1/4 chord is indeed a very drastic increase
of the hydrofoil and reaches the peak value at the cavitator location in comparison to the no-cavitator arrangement, all in the midst of a
of the 1/4 chord. Compared to the case without the cavitator, cl has rather limited change in cd .
increased by over 200% from the vicinity of 0.04 up to 0.10. After analyzing the simulation results for the cases at various
Further results of the simulation on the 2D NACA0012 foil at cavitator numbers and AOAs, we find that the general trend of the
the AOA of 1○ , 3○ , 5○ , 7○ , and 12○ are plotted in Figs. 11 and 12 cl and cd changes with the cavitator location is similar. Figures 9
to examine how the force coefficient changes at the different AOA. and 11 show that the cl will first decrease when the cavitator is placed
Here, σ is set at 0.2. Once again, cd has exhibited a somewhat lim- near to the leading edge (1/32 or 1/16 chord), then it will increase to
ited change to the magnitude in comparison to the no-cavitation the peak value when the cavitator placed at the 1/4 chord as the cav-
case for the different AOA. It is interesting to note that as for the itator move toward the trailing edge. Compared to the changes in cl ,
no-cavitator arrangement, as the AOA increases, cd increases mono- Figs. 10 and 12 show that cd only varies in a relatively small range.
tonically as expected. Just that it is observed that there is a limited This is deemed very significant as a large increase in cl is not accom-
variation of cd to about ±40% in the presence of the cavitator in com- panied by a correspondingly large increase in cd to negate potential
parison to the no-cavitator case. For cl , there appears to be two fairly engineering applications.
distinct trends. At the cavitation location nearer to the leading edge
B. Mechanics of cavitator and effect
on force coefficients
To study how the location of the cavitator affects the force coef-
ficient, the cavity profile and the pressure distribution that surround
the 2D NACA0012 foil at AOA = 7○ and σ = 0.2 are shown in Figs. 13
and 14 for the cases without or with the cavitator added at the 1/16
and 1/4 chord. The cavitating flow and pressure distribution near
the leading edge of and foil and the cavitator are enlarged to show
the detailed cavity formation and pressure distribution. The cavity
length is longer for the cases with the cavitator as the pressure distri-
bution around the leading edge of the hydrofoil is highly affected by
the cavitator placed nearby. It is clear from Fig. 13 that on the bottom
surface of the hydrofoil, flow cavitation commences at the location
of the cavitator vis-à-vis the distance of about 0.24 chord for the nat-
urally occurring flow supercavitation in the absence of the cavitator
FIG. 11. Lift coefficient varies with the location of the cavitator placed at the lower (see Fig. 14). The pressure distribution in Fig. 14 shows that the pres-
side of the 2D NACA0012 foil at AOA = 1○ , 3○ , 5○ , 7○ , and 12○ . σ = 0.2.
sure on the upper side of the hydrofoil remains almost unchanged in

Phys. Fluids 32, 123307 (2020); doi: 10.1063/5.0030907 32, 123307-6


Published under license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

FIG. 13. The cavity profile over the 2D


NACA0012 foil without the cavitator and
with the cavitator placed at the 1/16 and
1/4 chord. σ = 0.2. AOA = 7○ . l and t
are shown. The cavitating flow near the
leading edge of the foil and the cavitator
is enlarged for convenience. The small
arrows indicate the location of the cavita-
tor and where the cavitation on the lower
surface starts.

FIG. 14. The pressure distribution sur-


round the 2D NACA0012 foil without the
cavitator and with the cavitator placed at
the 1/16 and 1/4 chord. σ = 0.2. AOA
= 7○ . The pressure distribution near the
leading edge of the foil and the cavitator
is enlarged for convenience. The small
arrows indicate the presence of the cav-
itator and where the cavitation on the
lower surface starts.

the absence/presence of the cavitator. Then for the lower side, the Lift depends on the magnitude of the pressure difference and the
pressure distribution in the region between the leading edge of the extent of the high pressure region (see Fig. 15). Generally, the upper
hydrofoil and the cavitator is found to be much higher than the case surface pressure distribution shown in these figures has exhibited
without the presence of cavitator, and clearly, the said distribution is minor differences among the cases with and without the cavitator.
affected by the presence and location of the cavitator. The extent of
the high pressure (lhigh ) and low pressure region (llow ) on the lower
surface of the hydrofoil is defined in Fig. 15. To show quantitatively
how the pressure distribution on the hydrofoil is affected by the cavi-
tator, the pressure coefficient cp along the chordwise direction on the
upper and lower surfaces of the hydrofoil is presented in Figs. 16–18
for the cases without the cavitator and with the cavitator added at
the 1/16 and 1/4 chord, respectively. The pressure coefficient cp is
defined as
FIG. 15. Schematic of the supercavitating foil with the cavitator placed at the lower
p − p∞ side (exaggerated and not to scale). l high is the extent of the high pressure region;
cp = 1
. (18)
ρU∞ 2 l low is the extent of the low pressure region.
2

Phys. Fluids 32, 123307 (2020); doi: 10.1063/5.0030907 32, 123307-7


Published under license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

FIG. 16. The pressure coefficient along the chord direction on the upper and lower
surfaces of the 2D NACA0012 foil without the cavitator. σ = 0.2. AOA = 7○ . The
zoomed-in view shows that the pressure reaches the SVP of water at about 0.24c FIG. 18. The pressure coefficient along the chord direction on the upper and lower
on the hydrofoil lower surface. surfaces of the 2D NACA0012 foil with the cavitator placed at the 1/4 chord.
σ = 0.2. AOA = 7○ .

For the case without the cavitator, the pressure difference between
the upper and lower surfaces starts near the leading edge of the the no-cavitator case at lhigh ∼ 0.24c; thus, the overall lift is lower
hydrofoil. As mentioned, the natural flow supercavitation occurs at than the case without the cavitator. For the case with the cav-
about 1/4 chord on the bottom surface of hydrofoil, so lhigh ∼ 0.24c itator placed at the 1/4 chord (see Figs. 13 and 18), the pres-
(see Figs. 13 and 16). When the cavitator placed at the 1/16 chord, cp sure distribution between the leading edge of the hydrofoil and
changes drastically near the cavitator, and hence, the vicinity of the the cavitator is much higher in magnitude than the case with-
leading edge of the hydrofoil (see Fig. 17). In essence, a high pressure out the cavitator. The large pressure difference between the upper
region is generated starting from the leading edge of the hydrofoil and lower surfaces caused by the cavitator and extending over the
to the cavitator. The sharp decrease and increase in cp resulted in a region at lhigh = 1/4c. This lhigh is almost the same value as the
relatively high pressure magnitude on the lower surface. However, no-cavitator case, which coupled with the mentioned large pres-
for this case, as the cavitator is placed very close to the leading edge sure difference between the upper and lower surfaces therefore
of the hydrofoil and lhigh = 1/16c is nearly four times smaller than gives rise to an overall lift, and hence, cl becomes much larger
than the no-cavitator case (see Fig. 9). It is interesting and appro-
priate to note that broadly beyond the position of the cavitator
toward the trailing edge, the pressure takes on the SVP value that
is similar to the supercavitating region on the upper hydrofoil
surface and hence makes no or little contribution to the overall
lift.
In general, in supercavitating flow, the viscous drag is deemed
much smaller than pressure drag as most of the hydrofoil surface is
enveloped by the supercavity. By placing a cavitator on the hydro-
foil, we aim to change the pressure distribution on the lower surface
to control the lift. At the same time, we do not want to induce a
large flow separation, which will cause a drastic increase in the pres-
sure drag. Hence, as mentioned in Sec. I, the size of the cavitator we
have introduced in this study is deliberately kept small compared to
the foil. In Fig. 19, the viscous drag coefficient shows a perceptible
decrease with an increase of cavitating flow coverage on the lower
surface of the hydrofoil, whereas the drag coefficient contributed
by pressure dominates. Similar to the lift, the pressure drag highly
depends on the magnitude and distribution of the pressure on the
top and bottom surface of the hydrofoil. As the intent of the intro-
FIG. 17. The pressure coefficient along the chord direction on the upper and lower duced cavitator is to continue to maintain a streamlined flow over
surfaces of the 2D NACA0012 foil with the cavitator placed at the 1/16 chord. the hydrofoil, it is perhaps logical to examine closely the extent of
σ = 0.2. AOA = 7○ .
the cavitating region and wake following the hydrofoil for a possible

Phys. Fluids 32, 123307 (2020); doi: 10.1063/5.0030907 32, 123307-8


Published under license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

side of the hydrofoil along the chordwise direction (see Figs. 9 and
10). For cl , there must be a situation where we place the cavitator at a
certain position on the lower surface of the hydrofoil, the increased
pressure between the leading edge of the hydrofoil and the cavitator
is the same as the decreased pressure after the cavitator compared
to the no-cavitator case such that the overall lift coefficient of the
hydrofoil remains unchanged. In another word, if we place the cavi-
tator nearer the leading edge of the hydrofoil where high pressure
region occurs, then due to the change in increased pressure near
cavitator will be less dramatic (compare to the no-cavitator case)
and coupled to a smaller region of large pressure, cl will be smaller
than the case without the cavitator (see Fig. 9). If we place the cav-
itator further away from the hydrofoil leading edge (or where the
high pressure region occurs), the high pressure distribution over the
larger region between the leading edge of the hydrofoil and the cavi-
tator will therefore tend to lead to an overall increase in cl compared
to the no-cavitator case. For cd , the pressure drag coefficient domi-
nates. As the change in pressure drag coefficient is correlated with t,
FIG. 19. The pressure and viscous drag coefficient varies with the location of the which can then be considered as a proxy of the wake region behind
cavitator placed at the lower side of the 2D NACA0012 foil at AOA = 7○ . σ = 0.2. the hydrofoil, cd will increase as the cavitator moves closer to the
trailing edge of the hydrofoil. As such, by careful understanding on
how the lift and drag coefficients change with the location of the
cavitator on the lower side of the hydrofoil at different cavitation
numbers or AOAs, one may be able to device some degree of con-
relation to the pressure drag. (It is fair to say that if the cavitat- trol over the lift (and drag) of an underwater vehicle for engineering
ing region and wake become akin to that of the flow over a bluff applications.
body, then the pressure drag would have markedly increased.) From By placing the cavitator between the leading edge and the nat-
Fig. 13, t increases monotonically as the cavitator moves closer to the ural separation point on the lower surface of the hydrofoil (where
trailing edge of the hydrofoil, which is consistent with the monotonic cavitation starts for the no-cavitator case, ∼0.24c), we are able to
changes in pressure drag coefficient shown in Fig. 19. As long as the move the “separation” point closer to the leading edge. On the other
changes in t with the cavitator location is limited, the pressure drag hand, if we place the cavitator after the natural occurring separa-
coefficient only changes in a relatively small range. It is also very tion point, the separation point can be moved further away from the
interesting to note the close correlation between the pressure drag leading edge of the hydrofoil. Here, we add two more cases with the
coefficient and hence the drag coefficient to t. cavitator placed on the 1/5 (before the natural separation point) and
It is clear that the general behavior of the lift and drag coeffi- 1/3 (after the natural separation point) chord to verify the general
cients changes with the location of the cavitator placed on the lower trend of cl and cd change with the cavitator location as summa-

FIG. 20. The cavity profile and pres-


sure distribution surrounding the 2D
NACA0012 foil without the cavitator and
with the cavitator placed at the 1/5 chord
and 1/3 chord. σ = 0.2. AOA = 7○ . l and
t are shown.

Phys. Fluids 32, 123307 (2020); doi: 10.1063/5.0030907 32, 123307-9


Published under license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

separation point is about 0.24 chord). Figure 20 shows the com-


parison between no-cavitator case and the case with the cavita-
tor placed at the 1/3 chord. With the cavitator, a high pressure
region is generated between the leading edge and the cavitator (see
Fig. 21). As lhigh ≈ 31 c and t = 0.43c increases with x, both cl and cd
increase compared to the previous cases. In this case, cl = 0.116 and
cd = 0.061. Hence, the results are consistent with our discussion
above on the mechanics of the cavitator on the force coefficients
associated with the supercavitating flow. However, it is timely to
note the expected flow control by placing the cavitator after the nat-
ural separation point is still dependent on the size of the cavitator
introduced and, of course, the shape of the hydrofoil leading to the
supercavity enveloping the latter.

C. Cavitating hydrofoil with/without cavitator


at moderate cavitation number, σ = 0.4
The time averaged cavity profiles and lift coefficient of the cavi-
FIG. 21. The pressure coefficient along the chord direction on the upper and lower tating flow over the 2D NACA0012 foil without and with the cavita-
surfaces of the 2D NACA0012 foil with the cavitator placed at the 1/3 chord. tor placed at the 1/16 and 1/4 chord vs σ = 0.1, 0.2, and 0.4 at AOA
σ = 0.2. AOA = 7○ . 7○ are shown in Figs. 22–24. There is a significant difference in the
cavity length at different cavitation numbers. The cavity becomes
longer in the presence of the cavitator added on the lower side of
rized in our above discussions. We found that when the cavitator the hydrofoil. As the saturated pressure inside the cavitating flow is
is placed at the 1/5 chord, cl = 0.087 and cd = 0.056, which is con- 2300 Pa, the pressure distribution could therefore be highly affected
sistent with the earlier discussion. Next, we show the results of the by the location of the cavitator. For the case at σ = 0.4, the cavity over
case with the cavitator placed at the 1/3 chord (the natural occurring the hydrofoil changes from supercavitating flow to cloud cavitating

FIG. 22. The time averaged cavitating


profile and lift coefficient vs σ of the 2D
NACA0012 foil without the cavitator at
σ = 0.1, 0.2, and 0.4, AOA = 7○ . l and
t are shown.

FIG. 23. The time averaged cavitating


profile and lift coefficient vs σ of the 2D
NACA0012 foil with the cavitator placed
at the 1/16 chord at σ = 0.1, 0.2, and 0.4,
AOA = 7○ . l and t are shown.

Phys. Fluids 32, 123307 (2020); doi: 10.1063/5.0030907 32, 123307-10


Published under license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

FIG. 24. The time averaged cavitating


profile and lift coefficient vs σ of the 2D
NACA0012 foil with the cavitator placed
at the 1/4 chord at σ = 0.1, 0.2, and 0.4,
AOA = 7○ . l and t are shown.

flow. It may be noted that the cloud cavitating flow is unsteady with As discussed, the changes and trend of the lift coefficient with
evolving cavitation shape, while the supercavitating flow is steady the location of the cavitator at a σ of 0.4 are quite different from
with a stable cavitation shape.56,57 The instability of the cloud cavi- the cases at the smaller σ = 0.1 and 0.2: the cloud cavitating flow
tating flow such as cavity shedding and collapsing may induce large associated with σ = 0.4 surrounding the hydrofoil is unsteady. The
lateral force or even cause damage to the structure. cavitating flow given in Fig. 27 for the 3D NACA0012 foil with-
The force coefficients at σ = 0.4 are, however, quite different out the cavitator and with the cavitator placed at the 1/16 chord at
in terms of trend and quantity from the results of σ at 0.1 and 0.2 σ = 0.4 shows clearly the cavity growing and shedding after the
as cloud cavitating flow occurs. The cl and cd for unsteady cloud hydrofoil at t = 0.04 s, 0.12 s, and 0.2 s. This kind of unsteady phe-
cavitation cases at σ = 0.4 are time averaged quantities in Figs. 25 nomenon induced by the re-entrant jet inside the cavity is observed
and 26. For the case of comparison of Figs. 9 and 10 for cl and cd at during the cavity evolution, including the phenomenon such as cav-
σ = 0.1 and 0.2, these are replotted to indicate the different trend and ity shedding and collapsing.58–60 As the σ is large, the cavity hardly
magnitude. For the cases without the cavitator, cl of the hydrofoil grows continuously from the cavitator to envelop wholly the sur-
with cloud cavitating flow at σ at 0.4 is much larger than the hydro- face of the hydrofoil all the time. Hence, it is a timely and cau-
foil with supercavitating flow at σ = 0.1 and 0.2. The lift coefficient, tionary reminder that the major findings thus far as pertaining to
in general, decreases as the cavitator at the lower side of the hydro- the flow supercavitation at a lower cavitation number may not be
foil moves closer to the trailing edge of the hydrofoil and reaches the fully/directly applicable for the cloud cavitation flow of the moderate
lowest value of about 0.08 (which is 50% compared to the case with- cavitation number. Based on the mechanism analyzed in this paper,
out the cavitator) when the cavitator is placed at the 1/8 chord (see such a flow control may be achieved better by other active control
Fig. 25). In comparison to the cloud cavitating flows at σ is 0.4, the methods such as artificial ventilation to facilitate and ensure a fuller
cd of supercavitating flows at a σ of 0.1 and 0.2 is lower when the cavity envelope over the hydrofoil under the moderate cavitation
cavitator is placed at the same location (see Fig. 26). number regime.

FIG. 25. Lift coefficient varies with the location of the cavitator placed at the lower FIG. 26. Drag coefficient varies with the location of the cavitator placed at the lower
side of the 2D NACA0012 foil at σ = 0.1, 0.2, and 0.4. AOA = 7○ . Time averaged side of the 2D NACA0012 foil at σ = 0.1, 0.2, and 0.4. AOA = 7○ . Time averaged
value is shown for the unsteady cloud cavitation case at σ = 0.4. value is shown for the unsteady cloud cavitation case at σ = 0.4.

Phys. Fluids 32, 123307 (2020); doi: 10.1063/5.0030907 32, 123307-11


Published under license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

FIG. 27. The supercavitating flow over


the 3D NACA0012 foil without the cavi-
tator and with the cavitator added at the
1/16 chord. σ = 0.4. AOA = 7○ . t = 0.04 s,
0.12 s, and 0.2 s. The cavity shape is
shown by the iso-surface at the volume
fraction of water equals to 0.5. The small
arrows indicate the flow direction.

D. Supercavitating hydrofoil with/without cavitator surface, cavitation starts at the location of the cavitator (similar to
under 3D effect the 2D arrangement). By comparing the results of the foil without
In this section, the numerical results of the 3D NACA0012 foil the cavitator vis-à-vis cavitator placed at the 1/4 chord and 1/16
at σ = 0.2 and AOA 7○ are discussed to analyze the effect of the cavi- chord, it is found that the presence of the cavitator affects the cavity
tator on the cavity under the 3D scenario. The cavity shape of the 3D formation on the lower surface of the hydrofoil and thus the pressure
cavitating flow over the NACA0012 foil on the upper and lower sides distribution. Due to the tip vortex under the 3D effect, the curvature
is shown in Fig. 28. The aspect ratio of the foil is set at 4. It can be of the cavity decreases from the middle-section to the tip along the
clearly seen that the cavitating flow over the hydrofoil starts close spanwise direction. Both the maximum l and t increase with x, but
to the leading edge on the upper surface, whereas on the bottom the values are smaller than the 2D cases (see Fig. 28).
The lift and drag coefficients of the 2D and 3D NACA0012 foil
with and without the cavitator on the lower surface are shown in
Figs. 29 and 30. The results show that the trend of the cl of the 3D
cases varies with the position of the cavitator vis-à-vis no cavitator
case and somewhat fairly similar to the 2D results, but the ampli-
tude of variation is far smaller: the lift coefficient may increase or
decrease compared to the case without the cavitator when the cavi-
tator is placed at different positions, whereas drag only changes in a
relatively smaller range with the position of the cavitator (of course,

FIG. 28. The supercavitating flow over the 3D NACA0012 foil without the cavitator
and with the cavitator added at the 1/16 chord and 1/4 chord. σ = 0.2. AOA = 7○ .
Maximum l and t are shown. The cavity shape is shown by the iso-surface at the FIG. 29. Lift coefficient varies with the location of the cavitator placed at the lower
volume fraction of water equals to 0.5. The small arrows indicate the flow direction. side of the 2D and 3D NACA0012 foil at AOA = 7○ . σ = 0.2.

Phys. Fluids 32, 123307 (2020); doi: 10.1063/5.0030907 32, 123307-12


Published under license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

numbers of σ = 0.1 and 0.2 due to the unsteady cloud cavitating flow
over the hydrofoil for the former. It is found that the presence of
the cavitator used is not so effective at all times in greatly enhancing
the lift coefficient for the cloud cavitating flow as the cavity cannot
grow continuously at the cavitator to envelop the hydrofoil. It may
be thus necessary to apply other active control methods such as arti-
ficial ventilation to ensure a fuller and steady cavity envelope of the
hydrofoil.
Overall, the results show that the cavitator can have a significant
effect on the hydrodynamic forces of hydrofoil, since the existence
of the cavitator can alter the cavitation shape and further affect the
pressure distribution around the hydrofoil. A detailed analysis is
made to investigate how these changes are related to the position
of the placed cavitators, which is important to evaluate the efficacy
FIG. 30. Drag coefficient varies with the location of the cavitator placed at the lower
side of the 2D and 3D NACA0012 foil at AOA = 7○ . σ = 0.2. and effectiveness of the cavitator as a flow control approach/means.
The findings in this paper can serve as a good guidance on how to
improve the hydrodynamic performance and stability condition of
high-speed hydrofoils/wings with the application of the cavitator.
if the aspect ratio of the 3D hydrofoil is greatly increased from the
present 4.0, the results tend toward the 2D configuration).
NOMENCLATURE
bw wedge model half span
IV. CONCLUSIONS c chord length
In this study, we have investigated the effect of the cavitator cd drag coefficient
on supercavitating flow and hydrodynamic forces for high-speed cl lift coefficient
hydrofoils through a numerical approach. The numerical approach cp pressure coefficient
is based on the finite volume method for the NS equation with multi- cw wedge model chord length
phase models, in particular, the Kunz model is for phase change D drag
in cavitation, VOF for interfaces, and LES for details in cavitating g gravity
flow. The numerical method is validated by comparing with previous L lift
experimental data for the supercavitating flow over a wedge-shaped l cavity length
hydrofoil. After which, a series of numerical simulations are con- lhigh extent of the high pressure region
ducted to analyze the supercavitating flow that envelops a hydrofoil llow extent of the low pressure region
with/without the cavitator placed at different locations along the p pressure
chordwise direction on the bottom surface of hydrofoil. The effect p∞ standard atmospheric pressure
of the cavitator placement location on the cavity profile as well as pv saturated vapor pressure
the force coefficient is discussed. t maximum cavity thickness
The force analysis on the 2D supercavitating hydrofoil with or U∞ free stream velocity
without the cavitator is conducted first. The results show that the ρl liquid water density
cavitator placed at the lower side of the foil largely affect cl , while cd ρv water vapor density
only changes in a relatively small range as compared to cl . For the σ cavitation number
cases in supercavitating flow at σ = 0.1 and 0.2, the general trend
of the force coefficient changes with the cavitator position is simi-
lar. A high pressure region is generated between the leading edge of DATA AVAILABILITY
the hydrofoil and the cavitator, while after the cavitator, the pres-
sure takes on the value of the saturated vapor pressure of water. cl The data that support the findings of this study are available
depends on the magnitude and distribution of the pressure on the from the corresponding author upon reasonable request.
hydrofoil surfaces and broadly increases as the cavitator is placed
further away from the leading edge and toward the trailing edge. On REFERENCES
the other hand, cd is found to be closely correlated with the max- 1
imum thickness of the cavitating wake flow, which can serve as a J. P. Franc and J. M. Michel, Fundamentals of Cavitation (Springer Science &
Business Media, 2004).
proxy for the drag coefficient. For the 3D supercavitating foil with 2
G. Wang, I. Senocak, W. Shyy, T. Ikohagi, and S. Cao, “Dynamics of attached
an aspect ratio of 4, the general trend of cl varying with the position turbulent cavitating flows,” Prog. Aerosp. Sci. 37(6), 551–581 (2001).
of the cavitator is similar to the 2D results, whereas the amplitude of 3
C. E. Brennen, Cavitation and Bubble Dynamics (Oxford University Press, USA,
change is found to be smaller. 1995).
The changes of the lift coefficient are different at a moderate 4
R. T. Knapp, J. W. Daily, and F. G. Hammitt, Cavitation (McGraw-Hill, 1970),
cavitation number of σ = 0.4 in comparison to the smaller cavitation pp. 8:197–198:207.

Phys. Fluids 32, 123307 (2020); doi: 10.1063/5.0030907 32, 123307-13


Published under license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

5 30
Z. C. Pan, C. J. Lu, Y. Chen, and S. L. Hu, “Numerical study of periodically forced- C. L. Merkle, J. Feng, and P. E. O. Buelow, “Computational modeling of the
pitching of a supercavitating vehicle,” J. Hydrodyn. 22(1), 856–861 (2010). dynamics of sheet cavitation,” in Proceedings of the 3rd International Symposium
6 on Cavitation, Grenoble, France, 2001.
K. P. Yu, G. Zhang, J. J. Zhou,Z. O. U. Wang, and Z. W. Li, “Numerical study
31
of the pitching motions of supercavitating vehicles,” J. Hydrodyn., Ser. B 24(6), A. K. Singhal, M. M. Athavale, H. Li, and Y. Jiang, “Mathematical basis and
951–958 (2012). validation of the full cavitation model,” J. Fluids Eng. Trans. ASME 124(3), 617–
7 624 (2002).
S. L. Ceccio, “Friction drag reduction of external flows with bubble and gas
32
injection,” Annu. Rev. Fluid Mech. 42(1), 183–203 (2010). O. Coutier-Delgosha, F. Deniset, J. A. Astolfi, and J.-B. Leroux, “Numer-
8 ical prediction of cavitating flow on a two-dimensional symmetrical hydro-
Y. Jiang, S. Shao, and J. Hong, “Experimental investigation of ventilated super-
cavitation with gas jet cavitator,” Phys. Fluids 30(1), 012103 (2018). foil and comparison to experiments,” J. Fluids Eng. Trans. ASME 129(3), 279
9
E. Kadivar, M. V. Timoshevskiy, M. Y. Nichik, O. el Moctar, T. E. Schellin, and (2007).
33
K. S. Pervunin, “Control of unsteady partial cavitation and cloud cavitation in O. Coutier-Delgosha, R. Fortes-Patella, and J. L. Reboud, “Evaluation of the tur-
marine engineering and hydraulic systems,” Phys. Fluids 32(5), 052108 (2020). bulence model influence on the numerical simulations of unsteady cavitation,”
10
C. L. Merkle and S. Deutsch, “Microbubble drag reduction in liquid turbulent J. Fluids Eng. Trans. ASME 125(1), 38 (2003).
34
boundary layers,” Appl. Mech. Rev. 45(3), 103–127 (1992). O. Coutier-Delgosha, B. Stutz, A. Vabre, and S. Legoupil, “Analysis of cavitating
11
X. Shen, S. L. Ceccio, and M. Perlin, “Influence of bubble size on micro-bubble flow structure by experimental and numerical investigations,” J. Fluid Mech. 578,
drag reduction,” Exp. Fluids 41(3), 415–424 (2006). 171–222 (2007).
12 35
B. R. Elbing, S. Mäkiharju, A. Wiggins, M. Perlin, D. R. Dowling, and S. L. L. Zhou and Z. Wang, “Numerical simulation of cavitation around a hydrofoil
Ceccio, “On the scaling of air layer drag reduction,” J. Fluid Mech. 717(2), 484–513 and evaluation of a RNG κ-ε model,” J. Fluids Eng. Trans. ASME 130(1), 011302
(2013). (2008).
13 36
B. R. Elbing, E. S. Winkel, K. A. Lay, S. L. Ceccio, D. R. Dowling, and O. Coutier-Delgosha, J. L. Reboud, and Y. Delannoy, “Numerical simulation of
M. Perlin, “Bubble-induced skin-friction drag reduction and the abrupt transition the unsteady behaviour of cavitating flows,” Int. J. Numer. Methods Fluids 42(5),
to air-layer drag reduction,” J. Fluid Mech. 612(4), 201–236 (2008). 527–548 (2003).
14 37
X. Yu et al., “Experiment and simulation on air layer drag reduction of high- J. Wu, G. Wang, and W. Shyy, “Time-dependent turbulent cavitating flow
speed underwater axisymmetric projectile,” Eur. J. Mech. B: Fluids 52, 45–54 computations with interfacial transport and filter-based models,” Int. J. Numer.
(2015). Methods Fluids 49(7), 739–761 (2005).
15 38
E. Kawakami and R. E. Arndt, “Investigation of the behavior of ventilated B. Huang, Y. Zhao, and G. Wang, “Large eddy simulation of turbulent vortex-
supercavities,” J. Fluids Eng. Trans. ASME 133(9), 091305 (2011). cavitation interactions in transient sheet/cloud cavitating flows,” Comput. Fluids
16
B.-K. Ahn, T.-K. Lee, H.-T. Kim, and C.-S. Lee, “Experimental investigation of 92, 113–124 (2014).
39
supercavitating flows,” Int. J. Naval Arch. Ocean Eng. 4(2), 123–131 (2012). B. Ji, X. Luo, R. E. A. Arndt, and Y. Wu, “Numerical simulation of three dimen-
17
M.-R. Pendar and E. Roohi, “Investigation of cavitation around 3D hemi- sional cavitation shedding dynamics with special emphasis on cavitation-vortex
spherical head-form body and conical cavitators using different turbulence and interaction,” Ocean Eng. 87, 64–77 (2014).
40
cavitation models,” Ocean Eng. 112, 287–306 (2016). B. Huang, G.-y. Wang, and Y. Zhao, “Numerical simulation unsteady cloud
18
D. Yang, Y. L. Xiong, and X. F. Guo, “Drag reduction of a rapid vehicle in cavitating flow with a filter-based density correction model,” J. Hydrodyn. 26(1),
supercavitating flow,” Int. J. Naval Arch. Ocean Eng. 9(1), 35–44 (2017). 26–36 (2014).
41
19
E. Kadivar, E. Kadivar, K. Javadi, and S. M. Javadpour, “The investigation of R. E. Bensow and G. Bark, “Implicit LES predictions of the cavitating flow on a
natural super-cavitation flow behind three-dimensional cavitators: Full cavitation propeller,” J. Fluids Eng. Trans. ASME 132(4), 041302 (2010).
42
model,” Appl. Math. Modell. 45, 165–178 (2017). L. Li, Q. Jia, Z. Liu, B. Li, Z. Hu, and Y. Lin, “Eulerian two-phase modeling
20
E. Roohi, M.-R. Pendar, and A. Rahimi, “Simulation of three-dimensional cav- of cavitation for high-speed UUV using different turbulence models,” in 2015
itation behind a disk using various turbulence and mass transfer models,” Appl. IEEE International Conference on Cyber Technology in Automation, Control, and
Math. Modell. 40(1), 542–564 (2016). Intelligent Systems (CYBER) (IEEE, 2015), pp. 1247–1252.
21 43
W. Zou, K.-P. Yu, R. E. A. Arndt, G. Zhang, and Z.-W. Li, “On the shedding of N. Dittakavi, A. Chunekar, and S. Frankel, “Large eddy simulation of turbulent-
the ventilated supercavity with velocity disturbance,” Ocean Eng. 57(2), 223–229 cavitation interactions in a venturi nozzle,” J. Fluids Eng. Trans. ASME 132(12),
(2013). 121301 (2010).
22 44
Z. Wang, B. Huang, G. Wang, M. Zhang, and F. Wang, “Experimental and E. Roohi, A. P. Zahiri, and M. Passandideh-Fard, “Numerical simulation of
numerical investigation of ventilated cavitating flow with special emphasis on gas cavitation around a two-dimensional hydrofoil using VOF method and LES turbu-
leakage behavior and re-entrant jet dynamics,” Ocean Eng. 108, 191–201 (2015). lence model,” Appl. Math. Modell. 37(9), 6469–6488
23
M. Dular, R. Bachert, B. Stoffel, and B. Širok, “Experimental evaluation of (2013).
45
numerical simulation of cavitating flow around hydrofoil,” Eur. J. Mech. B: Fluids X. Yu, C. Huang, T. Du, L. Liao, X. Wu, Z. Zheng, and Y. Wang, “Study
24(4), 522–538 (2005). of characteristics of cloud cavity around axisymmetric projectile by large eddy
24
B. Stutz and J. L. Reboud, “Experiments on unsteady cavitation,” Exp. Fluids simulation,” J. Fluids Eng. Trans. ASME 136(5), 051303 (2014).
46
22(3), 191–198 (1997). A. Gnanaskandan and K. Mahesh, “A numerical method to simulate turbulent
25 cavitating flows,” Int. J. Multiphase Flow 70, 22–34 (2015).
B. Stutz and J.-L. Reboud, “Measurements within unsteady cavitation,” Exp.
47
Fluids 29(6), 545–552 (2000). R. W. Kermeen, “Experimental investigations of three-dimensional effects on
26 cavitating hydrofoils,” Technical Report No. 47-14, Engineering Division of the
V. Aeschlimann, S. Barre, and S. Legoupil, “X-ray attenuation measurements in
a cavitating mixing layer for instantaneous two-dimensional void ratio determi- California Institute of Technology, Pasadena, CA, 1960.
48
nation,” Phys. Fluids 23(5), 055101 (2011). G. Vernengo, L. Bonfiglio, and S. Brizzolara, “Supercavitating three-
27 dimensional hydrofoil analysis by viscous lifting-line approach,” AIAA J. 55,
I. Khlifa and O. Coutier-Delgosha, “Velocity measurements in cavitating flows
using fast x-ray imaging,” in Congrés Français de Mécanique, France, 2013. 4127–4141 (2017).
28 49
S. A. Mäkiharju, C. Gabillet, B. G. Paik, N. A. Chang, M. Perlin, and M. Passandideh-Fard and E. Roohi, “Transient simulations of cavitating flows
S. L. Ceccio, “Time-resolved two-dimensional X-ray densitometry of a two- using a modified volume-of-fluid (VOF) technique,” Int. J. Comput. Fluid Dyn.
phase flow downstream of a ventilated cavity,” Exp. Fluids 54(7), 1561 22(1-2), 97–114 (2008).
50
(2013). Y. Wang, L. Liao, T. Du, C. Huang, Y. Liu, X. Fang, and N. Liang, “A
29
R. F. Kunz et al., “A preconditioned Navier–Stokes method for two-phase study on the collapse of cavitation bubbles surrounding the underwater-launched
flows with application to cavitation prediction,” Comput. Fluids 29(8), 849–875 projectile and its fluid–structure coupling effects,” Ocean Eng. 84, 228–236
(2000). (2014).

Phys. Fluids 32, 123307 (2020); doi: 10.1063/5.0030907 32, 123307-14


Published under license by AIP Publishing
Physics of Fluids ARTICLE scitation.org/journal/phf

51 55
Y. Wang, C. Huang, X. Fang, X. Yu, X. Wu, and T. Du, “Cloud cavitat- C. J. Greenshields, OpenFOAM-The Open Source CFD Toolbox-User Guide
ing flow over a submerged axisymmetric projectile and comparison between (OpenFOAM Foundation, Ltd., 2015), Vol. 2.
two-dimensional RANS and three-dimensional large-eddy simulation methods,” 56
X. Cheng, X. Shao, and L. Zhang, “The characteristics of unsteady cavitation
J. Fluids Eng. Trans. ASME 138(6), 061102 (2016). around a sphere,” Phys. Fluids 31(4), 042103 (2019).
52 57
C. Yu, Y. Wang, C. Huang, T. Du, C. Xu, and J. Huang, “Experimental and O. Supponen, P. Kobel, D. Obreschkow, and M. Farhat, “The inner world of
numerical investigation on cloud cavitating flow around an axisymmetric projec- a collapsing bubble,” Phys. Fluids 27(9), 091113 (2015).
tile near the wall with emphasis on the analysis of local cavity shedding,” Ocean 58
Y. Long, X. Long, B. Ji, and T. Xing, “Verification and validation of Large
Eng. 140, 377–387 (2017). Eddy Simulation of attached cavitating flow around a Clark-Y hydrofoil,” Int. J.
53
F. Nicoud and F. Ducros, “Subgrid-scale stress modelling based on the Multiphase Flow 115, 93–107 (2019).
square of the velocity gradient tensor,” Flow, Turbul. Combust. 62(3), 183–200 59
H. Y. Cheng, X. R. Bai, X. P. Long, B. Ji, X. X. Peng, and M. Farhat, “Large eddy
(1999). simulation of the tip-leakage cavitating flow with an insight on how cavitation
54
F. G. Schmitt, “About Boussinesq’s turbulent viscosity hypothesis: Historical influences vorticity and turbulence,” Appl. Math. Modell. 77, 788–809 (2020).
60
remarks and a direct evaluation of its validity,” C. R. Méc. 335(9-10), 617–627 T. Du, Y. Wang, L. Liao, and C. Huang, “A numerical model for the evolution
(2007). of internal structure of cavitation cloud,” Phys. Fluids 28(7), 077103 (2016).

Phys. Fluids 32, 123307 (2020); doi: 10.1063/5.0030907 32, 123307-15


Published under license by AIP Publishing

You might also like