You are on page 1of 18

1 abstract 1

Sixth International Symposium on Cavitation


CAV2006, Wageningen, The Netherlands, September 2006

An attached cavity on a three-dimensional


hydrofoil

Evert-Jan Foeth Tom van Terwisga


Technical University of Delft Technical University of Delft &
Laboratory of Ship Maritime Research Institute of
Hydromechanics the Netherlands
+31-152786603
e.j.foeth@wbmt.tudelft.nl

1 abstract
A 3D twisted hydrofoil with a cavitation pattern closely related to propellers was
observed with a high-speed camera at the University of Delft Cavitation Tunnel.
Re-entrant ow coming from the sides of the cavity aimed at the center plane,
termed side-entrant ow, collided in the closure region of the cavity, pinching
o a part of the sheet resulting in a periodic shedding. The collapse of the
remainder of the sheet appears to be a mixing layer, originating in the closure
region moving upstream up to the leading edge. Collision of side-entrant jets in
the closure region of a cavity is identied as a second shedding mechanism, in
addition to re-entrant ow impinging the sheet interface at the leading edge.

2 Introduction
Fully developed sheet cavitation on ship propellers is a major cause of noise, vi-
bration and erosion. Although the nal evaluation of a propeller design is based
on model experiments, numerical methods for cavitation simulation are becom-
ing increasingly more important. Potential ow solvers are now the industry
standards (e.g., Young & Kinnas [19]), but in the last decades an increase in
both Euler (e.g., Choi & Kinnas , Sauer & Schnerr [4]) and RANS (e.g., Kunz
et.al. [15]) codes are observed. However, up to now these simulations are not
able to capture the pressures radiated by cavitation or to predict erosion loca-
tion and severity on propellers [12]. To improve the description of the cavity
behavior and especially the unsteady shedding in the form of cloud cavitation
and in support of the rapidly expanding eld of numerical simulation this ex-
perimental research was started with a threefold goal. 1st analyze the physical
mechanisms of the instability of the cavity 2nd : build a data set of simple cavi-
tating ows to be used as benchmark material for computational uid dynamics
3 Setup 2

validation, and 3rd , extend the insights gained to guidelines for propeller design.
Here we focus on the description of the ow eld around an attached cavity and
its shedding mechanism.
Cavitation has been extensively tested in the past on two-dimensional hy-
drofoils (i.e. Franc [10]. However, cavitation on ship propellers is distinctly
three-dimensional due to the propellers three-dimensional geometry, the radi-
ally increasing velocity and change in blade loading, and a periodic change of
inow conditions due to the wake behind a ships hull. As studying cavitation
on a rotating object is inherently more dicult, three-dimensional hydrofoils
were designed resulting in a cavitation topology closely related to propellers
with periodically changing inow conditions. This allows for observations of
the inuence of controlled three-dimensional eects. For this study a span-
symmetric 3D hydrofoil is chosen, creating an isolated sheet cavity around the
plane of symmetry. The hydrofoil is lightly loaded at the tunnel walls to avoid
any interaction of the cavity with the tunnel boundary layer.
Crimi [5] studied the eect of sweep (skew) and concluded the inception
velocity increased with an increase in the skew angle. Hart et.al. used an
oscillating three dimensional hydrofoil [11]. De Lange & de Bruin [8] studied
three transparent hydrofoils in the cavitation tunnel at the University of Delft.
De lange & de Brujn concluded that the re-entrant of the 2D hydrofoil was
directed upstream, but in the 3D case the re-entrant jet component normal to
the closure line was reected inward. As the pressure gradient is perpendicular
to the closure line, it is to be expected that the tangential component remains
unchanged. Laberteaux & Ceccio [16] studied a series of swept wedges. The
cavity plan form was signicantly changed and the re-entrant jet was directed
into the cavity allowing for a steady sheet which only shed cloud cavitation at
the far downstream edge. Dang & Kuiper [7] numerically studied the re-entrant
on a hydrofoil with a span wise varying angle of attack and found the re-entrant
jet direction to be strongly inuenced by the cavity topology. The change in
cavity shape was determined by loading and not by the sweep angle.
In this paper high speed recordings are presented with our interpretation
for the shedding behavior for two distinct cases, a cavity of roughly half the
chord length and a (near) super cavity. The shedding mechanism for both
cases diered from two-dimensional shedding - where the re-entrant jet reaches
the leading edge - but was governed by the three-dimensional topology of both
hydrofoil and sheet cavity. A brief description of the setup of the experiment is
given in section 3. The observations are described and interpreted in section 4
and conclusions are in section 6.

3 Setup
The experiments were performed in the University of Delft Cavitation Tunnel
(see gures 1 & 2), with an eective measuring channel 0.60 m in length with
a 0.3 m x 0.3 m cross section with optical access from all sides; velocities up
to 10 m/s per second can be attained and the local pressure can be reduced to
5,000 P a. The non-dimensional cavitation number is dened as
p0 pV
= 1 2
(1)
2 V0
3 Setup 3

or, the ratio of the pressure head to the vaporization pressure (pV ) and the
dynamic pressure located at the test section entrance.

Fig. 2: A close-up of the test sec-


tion showing the hydrofoil (suc-
tion side up, transparent hy-
Fig. 1: A sketch of the Delft Cavitation drofoil mounted upside down),
Tunnel consisting of two cylin- the camera location and eec-
drical and two square channels tive viewing area

The test object is a three-dimensional hydrofoil, previously used by Dang


[6], with a chord length of C = 150 mm, a span of S = 300 mm (spanning
the entire test section), and a span wise varying angle of attack (3). This
geometric angle of attack varies as sketched in 4, rotating the sectional prole
(NACA000) around the mid chord position. (See Foeth et.al. for specic details
[9]). Calculations by Koop. et.al. [13] indicate that the change in eective angle
of attack is only 2 degrees. The hydrofoil was manufactured in both anodized
aluminum and perspex. The perspex hydrofoil was lmed from the pressure
side so that the cavity could be viewed from within.
Geometric angle of attack

8
7
6
Angle [deg]

5
4
3
2
1
0
0 50 100 150 200 250 300
Span [mm]

Fig. 4: Distribution of the geomtetric


Fig. 3: Top, side, and front view of the angle of attack of the hydrofoil.
hydrofoil. The black outline in- The angle at the sides is taken
dicates the viewing area of g- as the reference angle for the
ure 6 & 11 whole geometry

At low Reynolds numbers the boundary layer near the minimum pressure
region will be laminar and no cavity sheet will appear. When the boundary layer
separates and a separation region is formed, a smooth and glassy cavity can
appear. With the limited Reynolds numbers typically present at small scales,
4 Observations 4

transition to turbulence does not occur unless the boundary layer is locally
disturbed. When it does occur natural transition to turbulence can temporarily
suppress leading edge detachment [10]. As a 3D hydrofoil is used at moderate
velocities, the sides of the attached cavity can be locally suppressed as the ow
remains laminar, so roughness elements of 120 m were applied at the leading
edge (4% chord length) as a turbulence tripping mechanism. The roughness
elements can lead to local streaks of cavitation appearing next to the main
cavity. At too low speeds, the entire detachment region near the leading edge
may resemble an agglomeration of such streaks which was observed at 5 m/s.
The gas content was measured to be less than 0.1%, , but the roughness will
supply the degassed ow with ample nuclei for sheet cavitation to develop [14];
incipient cavitation on roughness elements is typically observed when equals
the minimum pressure coecient [3] and the nuclei content of the ow is no
longer critical.
The camera used for the high-speed imaging is a Flowmaster HighSpeedStar
4 (Photron Ultima APX) with a 10 bit dynamic range, 1 M pix resolution at 2
kHz with a maximum acquisition frequency of 120 kHz (0.4% full resolution)
with 2.6 GB memory. The lens is a Nikon AF Nikkor 50 mm, used with a
f-stop of 2.8. A New Wave Pegasus dual-head, high repetition, diode pumped
Nd: YLF laser was used as a stroboscope, with a 180 ns duration with a 10
mJ/pulse power at 1,000 Hz.

4 Observations
The shedding process of the attached cavity is classied into three regimes. At
high cavitation numbers ( > 1.1), the attached cavity was short in length and
present over a wide range of the leading edge and hence mainly two-dimensional.
This cavity was shedding vortices intermittently, no large cloudy structures were
identied. Such a closure is termed open by Laberteaux & Ceccio [16]. At
moderate cavitation numbers, large structures were shed regularly. The middle
regime is dominated by the three dimensionality of the cavity. Lowering the
pressure further created an attached cavity reaching a length comparable to
the chord length of the hydrofoil ( < 0.80). Shedding was then intermittent
and irregular. The cavity spanned the entire foil and was once more mainly
two-dimensional.
Visual analysis of the high speed video recordings indicated that the Strouhal
number
fl
St = (2)
V
based on cavity shedding frequency f and cavity length l was around St = 0.185
for moderate cavitation lengths when 0.8 < < 1.1. For two-dimensional
hydrofoils, Strouhal numbers of St = 0.25 0.40 were reported or specied by
Arndt [1] as:

St = 1/4 1 + (3)
from which can be concluded that the resulting shedding of the sheet diered sig-
nicantly from a two-dimensional cavity shedding due to the three-dimensional
geometry of the foil, by Strouhal alone. In gure 5 the Strouhal numbers are
4 Observations 5

0.4

0.3

St

0.2

5 m/s =1
7 m/s =0
0.1 7 m/s =1
7 m/s =2
Arndt

0.0
0.40 0.50 0.60 0.70 0.80 0.90 1.00 1.10 1.20 1.30 1.40

Fig. 5: Strouhal numbers at two dierent velocities based on maximum cavity


length with an average of St = 0.185, signicantly lower than for a two-
dimensional foil. Several points at St = 0 are visible indicating either
irregular shedding ( < 0.80) or the absence of large structure shedding
resulting in an open cavity ( > 1.1). The continuous line is the relation
by Arndt [1] for 2D cavity shedding.

plotted versus the cavitation number, as well as eq. 3. There was no indica-
tion that the Strouhal number was dependent on the cavitation number for the
tested angles of attack.

4.1 Case study at a low sigma.


An example of a (super) cavity is given in gure 6 showing the shedding in
detail. In g. 6.1-6 the location of the front of the re-entrant jet is given by the
arrow on the top-center location of each frame. Although dicult to identify on
photographs, it is clearly seen to move slowly forward on the recordings. The
collapse of the sheet in g. 6.1-10 starts from the end, moving upstream.
The shed ow structure in g. 6.7-8 consists of primary span wise and sec-
ondary stream wise vortices, similar to the turbulent cavitating shear ow struc-
ture observed behind steps and other mixing layers [18]. Figure 7 shows the
formation of a span wise vortex from gure 6.5 (showing intermediate images as
well). Such a span wise vortex system can be a result of a Kelvin-Helmholtz in-
stability with a street of positive vortices (where the vorticity has the same sign
as the hydrofoils net circulation). A close-up of g. 6.6-9 is given in g. 8, includ-
ing all intermediate images. Bernal & Roskho describe this observed structure
of span wise and stream wise vortices in more detail using a helium-nitrogen
mixing layer [2]. Thus, the observed cloud cavitation is similarly structured and
greatly resembles mixing layer.
The stream wise vortices originated as a single span wise vortex warped around
the primary span wise vortices. The smaller scale vortices can be seen to be
stretched around the periphery of the span wise structures with an increase
4 Observations 6

Fig. 6.1: Fig. 6.2: Fig. 6.3: Fig. 6.4:

Fig. 6.5: Fig. 6.6: Fig. 6.7: Fig. 6.8:

Fig. 6.9: Fig. 6.10: Fig. 6.11: Fig. 6.12:

Fig. 6.13: Fig. 6.14: Fig. 6.15: Fig. 6.16:

Fig. 6: Visualization at 6.89 m/s 7.70%, =1, =0.49 28.4%, f =400 Hz.
Viewing area as outlined in gure 3, ow from top to bottom. Every 5th
frame is show. White outlines indicate areas enhanced in g. 7 & 8
4 Observations 7

of their vaporous cores. Normally in shear layers, cavitation inception is rst


observed in these streamwise vortices [18], but in the case of a sheet cavity
break-up vapor is trapped in the initial formation.
From gure 6.4-8 is visible that the front of the mixing moved forward and
the collapse is cascading toward the leading edge. The direction of the collapse
of the sheet starts out as radially divergent but the front draws parallel to the
span as it progressed. The front of the disturbance accelerated up to the mean
stream velocity when reaching the leading edge, as determined from frame-by-
frame analysis. The approximated location of the front at the center plane is
identied and plotted in g. 9. It appears that the collapse front accelerates
(roughly) constantly.

The re-entrant jet momentum depends on the pressure gradient in the closure
region (Le et.al. [17]. The increase in collapse speed may be explained as
following. At the start of the collapse cycle the cavity is a well-dened structure
and due to its 3D geometry with a symmetry plane, only a stagnation point
is present in the closure region (g 10.1). After the rst pinch-o the closure
region of the cavity has changed from a convex into a concave or at shape and a
short high pressure line element is present (g 10.2), widening with each further
pinch o, gradually changing from a point source into a line source (g 10.3) as
the cavity loses its three dimensionality. From observations at higher values of
- presented below - it is visible that the re-entrant jet on a three-dimensional
cavity diverges radially from the closure into the sheet when the cavity is fully
grown.

4.2 Case study as higher sigma


In gure 11 a full shedding cycle at 5 m/s is shown with the hydrofoil lmed
from above, ow top to bottom. The shedding was repeatable, constant in its
shedding frequency, and always followed the same macro structural collapse.
The shedding cycle of the cavity in gure 11 is broken up into 4 parts. Desta-
bilization, primary and secondary shedding, followed by growth into its initial
condition. Note that there is a short overlap between primary and secondary
shedding (and growth). The primary shedding is located at the mid plane of the
hydrofoil, the secondary shedding is visible at the sides of the primary shedding
as two distinct smaller vortices.

1 11.1-11.4 Initial disturbance


2 11.5-11.12 Primary Shedding (cavity center)
3 11.9-11.16 Secondary Shedding (cavity sides)
4 not presented Growth

4.3 Initial disturbance


Figures 11.1-11.4 show the convex cavity, here considered fully grown. The lower
part of the cavities interface is turbulent, while the cavity at the sides and near
the leading edge is glassy and transparent. It is in the closure region where the
cavity became turbulent rst, not near the leading edge, as is typical of large
structure shedding on two-dimensional hydrofoils. The reasons are two-fold.
4 Observations 8

Fig. 7: A close-up of gure 6 shows the formation of a large span wise vortex at
2,000 Hz. As the main sheet collapses, a trail of very small vortices is
created, merging in several distinct larger structures
4 Observations 9

Fig. 8: Close-ups of g. 6.6-9 including intermediate images. The stream wise


cavitating vortices originate from a small vortex generated at the cavity
collapse and are stretched around the span wise vortices.

Estimate Velocity Collapse Front Midplane

70 9

8
Distance from leading edge [mm]

60
7
50
6
Velocity [m/s]

40 5

30 4

3
20
Velocity 2
10 Location
1

0 0
0 1 2 3 4 5 6 7 8 9 10 11
Time [ms]

Fig. 9: The location and velocity of the visible collapse front as visible in gs
6.5-6.10 determined by frame to frame analysis. Error bars indicate a
10 pixel error in the location of the collapse front. The front seems to
accelerate linearly.
4 Observations 10

Fig. 10: The stagnation point behind the cavity may grow into a stagnation
line element during its collapse as it increasingly more resembles a two-
dimensional sheet

First, the closure region in a 2D ow would normally be succeeded by a


stagnation line (orthogonal to the 2D plane), here it was a stagnation point
at the mid plane weakening the local pressure gradient. As indicated above,
the momentum in the re-entrant jet depends on the pressure gradient in the
closure region, so the three-dimensionally resulted in a re-entrant jet diverging
radially into the cavity from its closure at the mid span position. Therefore,
its forward momentum is diminished as it progressed into the cavity. Second,
at the sides of the sheet the pressure gradient forces the ow over the sheet
into the cavity roughly mirroring the streamlines at the interface contour as
sketched in gure 12. De lange & de Brujn concluded that the re-entrant jet of
the 2D hydrofoil was directly upstream, but in the 3D case the re-entrant jet
component normal to the closure line was reected inward. As the (instationary)
pressure gradient is perpendicular to the closure line, it is to be expected that
the component tangential to that closure line remains unchanged. At the sides
of the cavity, re-entrant ow had a very small span-wise component and was
directed downstream. The span wise component was largest when the cavity
closure contour was roughly at 45 with the incoming ow where the velocity
component in downstream direction of the re-entrant ow was zero. At larger
angles, the re-entrant jet had an upstream component (g. 12). Flow from the
sides was not obstructed, nor was it directed at the leading edge. When the
sheet cavity was growing, it owed unobstructed to the center of the cavity.
To distinguish between various directions of the re-entrant ow, the term
side-entrant jet is introduced. This term refers to the that part of the re-
entrant jet that has a strong span wise velocity component directed into the
cavity originating from the sides (and may have a small up- or down stream
component). The term re-entrant jet is reserved for the ow originating from
that part of the cavity where the closure is more or less perpendicular to the
incoming ow and is thus mainly directed upstream (but may have a small
lateral velocity component). The re-entrant ow is thus split up in re-entrant
and side entrant components, even though at certain points of the ow both
terms may apply. The term side-entrant is introduced to emphasize the three-
dimensional character of the ow. For the case presented, the side-entrant jets
from both sides were owing into the closure region of the sheet where they
collided. Side-entrant jets of the re-entrant ow do not reach the leading edge
but may form an equally important source for the shedding.
Any uid ejected upward through the cavity interface creates a signicant
disturbance, isolating a small portion of vapor and creating a bubbly ow as the
jet entrained vapor. The velocity of a simple streamline at the cavity surface is
4 Observations 11

Fig. 11.1: Fig. 11.2: Fig. 11.3: Fig. 11.4:

Fig. 11.5: Fig. 11.6: Fig. 11.7: Fig. 11.8:

Fig. 11.9: Fig. 11.10: Fig. 11.11: Fig. 11.12:

Fig. 11.13: Fig. 11.14: Fig. 11.15: Fig. 11.16:

Fig. 11.17: Fig. 11.18: Fig. 11.19: Fig. 11.20:

Fig. 11: Visualization at 4.96ms 6.4%, = 1, = 0.66 7.94%, f = 285.7Hz


(every 7th frame of the 2,000 Hz recording). Flow from top to bottom.
4 Observations 12


measured at VV = V0 1 + [9]. Although the exact velocity of the re-entrant
ow is dicult to measure, the velocity of the jets is unlikely to be a order of
magnitude lower. Also, if we assume that during the shedding cycle the two
side-entrant jets were converging for about a third of the shedding cycle (15
Hz), the amount of uid through a square millimeter - taking a homogeneous
velocity distribution - at this velocity of 6.4m/s was about (6400 2)/(15 3)
=
285 mm3 /shedding per mm2 of jet . At this rate the cavity closure could be
collecting uid quickly even if the jets were thin.

Fig. 13: Estimate of the direction of


Fig. 12: Streamlines over the cavity di- re-entrant ow focusing in the
rected inward lobes causing a second pinch-
o

4.4 Primary Shedding (cavity center)


The primary shedding originated at the collision region in the center of the
sheet, see gs refg:largetable2.5-11.12. However, only a portion was broken
o from the main sheet and advected with the ow. Most of the cavity remained
attached. This structure could be seen to roll up quickly in gures 11.5 -11.8
by self-induction into a hairpin. This structure grew signicantly in height, on
the order of the cavity length [9]. The cavity closure after the cut-o of the
hairpin vortex was temporarily turbulent - shedding a large cloudy structure -
but reattached smoothly shortly thereafter. In order to visualize the re-entrant
ow more clearly, a series of images of the transparent hydrofoil is presented
in gs 14 & 15. The cavitation is lmed through the hydrofoil showing the
internal structure. The radially diverging re-entrant ow is clearly visible in
g. 14(denoted as A) as waves on the jet surface reected the laser light.
The re-entrant ow directed upstream in a 2D situation would be constrained
in its lateral movement. The vapor interface at the leading edge was not visibly
disturbed upon contact with this re-entrant ow; its weakness apparently did
not lead to immediate shedding. As the side-entrant jets were aimed at the
closure, it was here that the uid rst impinged on the interface. Therefore, the
re-entrant component did not seem to be the main cause for the detachment of
the main structure.
4 Observations 13

Fig. 14: The re-entrant ow was lmed through a transparent hydrofoil. The
images show the re-entrant jet after cleaning up the pictures (despeckle,
color & histogram enhancement) . These gures shows the radially
diverging re-entrant (A) emanating from the center of the foil at two
dierent shedding cycles as sketched in g. 17. The two horizontal lines
are holes for ink injection (not presented)

4.5 Secondary shedding (cavity sides)


The remaining topology of the sheet in g.s 11.5 to 11.12 was concave. The
locally convex regions of the cavity are seen to shed a series of larger vortices,
followed by a turbulent region. From observation, the secondary greatly re-
sembled the primary shedding. The secondary shedding disappeared when the
closure was no longer concave.
The re-entrant ow direction in the center was still directed radially out-
ward. The main side-entrant jets and radially diverging re-entrant jet were now
converging in both downstream lobes of the remaining cavity shape (g. 16 &
17). The secondary shedding was caused by the collision of these two ows.
Basically, the main shedding as visible on gs 11.5-12 was repeated next to the
center plane as visible in g. 11.9-18.
After the secondary shedding the remaining cavity had a reasonably convex
shape with two concave hollows (denoted H) (g. 18). Figure 15 shows the
cavity after the secondary shedding, corresponding to g. 18. The movement of
the front of the side-entrant ow (A) from the hollows at g. 18.H can be seen,
as re-entrant ow forced into the cavity collides with re-entrant ow from the
plane of symmetry (B) and a frothy turbulent region is created upon impact at
the lower corners of the cavity closure. Both situations clearly show the presence
and inuence of a lateral velocity component at the cavity closure. Side-entrant
ow also appeared either side of the hollows (H). At the outer side, a possible
tertiary shedding could have occurred as this ow collides with the ow from
the main cavity sides (C). However, with each subsequent shedding the scale
and hence jet momentum decreases while the inow from the sides of the cavity
remained constant.
Apparently, the momentum from the ow from the sides advected ow from
the hollows at H, allowing for the growth of the cavity into its original status. In
4 Observations 14

Fig. 15.1: Fig. 15.2:

Fig. 15.3: Fig. 15.4:


s

Fig. 15: This series shows the cavity at the end of its secondary shedding cycle.
The side-entrant jet is seen to develop at both corners of the sheet (A)
as visualized in g. 18. The re-entrant jet is visible near the leading
edge (B)
5 Cavity closure 15

Fig. 16: The streamlines at the side Fig. 17: Estimate of the direction of
planes in the concave part are re-entrant ow focusing in the
partly directed away from the lobes causing a second pinch-
center plane o

Fig. 19: Side-entrant jets converge in


the closure region and cut
o the rst vortical structure.
Fig. 18: The streamlines at the side The remaining cavity closure is
planes in the concave part are now open.
partly directed away from the
center plane.

a brief period of converging main side-entrant jets, the cavity grew back into its
initial crescent shape again but that was merely a period of respite before side-
entrant jets at the center plane collided once again. The cavity never reaches a
constant length.

5 Cavity closure
The shedding of the sheet cavity of the three-dimensionality is similar to a two-
dimensional shedding, having its origin an a disturbance of its interface, except
that the disturbance occurs at the aft part of the sheet. The uid impinging on
the interface isolates a region of vapor as sketched in g.
19. If the interface is
considered a streamline with a tangent velocity VV = V0 1 + , it is immedi-
ately apparent from boundary integral of that velocity over S, that circulation
is detached and adverted with the ow. The impingement and detachment of
this vapor structure is inertial in nature. The mixing layer with its region of
high shear and strong vortices visibly generates vorticity.
6 Conclusions 16

6 Conclusions
From this experiment follows that re-entrant ow from the sides dictates the
behavior of the shedding cycle and the ow from the sides depends on the cavity
shape. Thus the cavity topology largely dictates the re-entrant ow directions
and focal points of this ow. Its shape and motion governs its behavior and the
convex cavity shapes seem to be intrinsically unstable. Re-entrant ow reaching
the leading edge does not seem to be the only cause for shedding. The re-entrant
ow can be both moving upstream but also in span wise direction denoted as
a side-entrant component. For any convex cavity shapes, side-entrant compo-
nents of the re-entrant jet focus in the closure region of the sheet, creating a
disturbance causing a local break-o of the main sheet structure. Side-entrant
components may collide before the re-entrant ow reaches the leading edge or
the upstream directed re-entrant ow is too weak to cause shedding at the
leading edge, changing into a dierent shedding mechanism. Side entrant com-
ponent focusing is suggested as a second shedding mechanism for attached sheet
cavitation in addition to the re-entrant components reaching the leading edge.
The shedding mechanism observed after side-entrant jet collision at the cen-
ter plane is a pinch-o of a part of the attached cavity. The observed (cavi-
tating) vortices after the shedding lead to the conclusion that a mixing layer
is present with its characteristic span wise and stream wise vortices which are
clearly visible on the images presented for a large scale cavity.
Pressure measurements by Le et al. [17] indicated that thin cavities have a
smooth pressure recovery at the cavity closure generating re-entrant ow with
a minimum of momentum. In these experiments continuous mixing can be
present: the cavities of this experiment are thicker, but some cavities of signif-
icant length (over 50% chord) do not show shedding but have a open closure.
In the absence of shedding of large structures the closure of the sheet cavity re-
mained turbulent at all times, - shedding small vortices - but no large structure
is shed as side-entrant ow was continuously aimed at the same location of the
sheet. A cavity does not need to be thin to have an open closure if its closure
is continuously supplied with uid from the sides. If no appreciable change in
pressure is generated at the closure, re-entrant ow as visualized in g. 6 is not
present. Therefore, a long and thin sheet cavity could be fully stable if its clo-
sure consists of a shear layer and the weak adverse pressure gradient precludes
the formation of a (strong) re-entrant jet.
The alternating shedding seen on the presented hydrofoil results in distinct
cycle, but the two-dimensional hydrofoil lacks the symmetry plane in the center,
resulting in the seemingly random local shedding along its cavity closure. Any
disturbance at its closure will re-direct the re-entrant ow into side-entrant ow
resulting in focal points and subsequently into local shedding. As a result, the
two-dimensional cavity has a highly three-dimensional structure making it a
more dicult study object, either numerical or experimental.
For future experiments the eects of an unsteady inow will be investigated.
We intend to use a ow oscillator to stimulate the cavity with frequencies both
higher and lower than the natural shedding frequency and expand the range of
hydrofoils. We intend to use hydrofoils with a steeper angle of attack distri-
bution to enhance the 3D eects further and with dierent sectional proles to
compare the inuence of cavity thickness.
7 Acknowledgments 17

7 Acknowledgments
This research is funded by the Dutch Technology Foundation STW project
TSF.6170 and the Royal Netherlands Navy. See www.stw.nl for more details.

References
[1] R.E.A. Arndt, C. Ellis, and S. Paul. Preliminary investigation of the use of
air injection to mitigate cavitation erosion. Journal of Fluids Engineering,
117:498592, 1995.
[2] L.P. Bernal and A. Rosko. Streamwise vortex structure in plane mixing
layers. Journal of Fluid Mechanics, 170:449525, 1986.
[3] J.F Caron, M. Farhat, and F. Avellan. Physical investigation of the cavita-
tion phenomenon. In 6th International Symposium on Fluid Control, Mea-
surement and Visualization (Flucome 2000), Sherbrooke, Canada, 2000.
[4] J.K. Choi and S.A. Kinnas. A 3-d euler solver and its application on the
analysis of cavitating propellers. 25th American Towing Tank Conference,
1998.
[5] P Crimi. Experimental study of the eects of sweep on hydrofoil loading
and cavitation. Journal of Aeronautics, 4(1):39, 1970.
[6] J. Dang. Numerical simulation of unsteady partial cavity flows. PhD thesis,
Technical University of Delft, the Netherlands, 2000.
[7] J. Dang and G. Kuiper. Re-entrant jet modeling of partial cavity ow on
three-dimensional hydrofoils. Journal of Fluids Engineering, 121(4):781
787, 1999.
[8] D.F. de Lange and G.J. de Bruin. Sheet cavitation and cloud cavitation,
re-etrant jet and three -dimensionality. Applied Scientific Research, pages
91114, 1998.
[9] E. J. Foeth, C. W. H. van Doorne, T. van Terwisga, and B. Wienecke. Time-
resolved piv and ow visualization of 3d sheet cavitation. Experiments in
Fluids, 40:503513, 2006.
[10] J.P. Franc and J.M. Michel. Attached cavitation and the boundary numer-
ical treatment. Journal of Fluid Mechanics, 154:6360, 1985.
[11] D.P. Hart, C.E. Brennen, and W.L. Acosta. Observations of cavitation
on a three dimensional hydrofoil. In 25th Cavitation and Multiphase Flow
Forum, 1990.
[12] ITTC. Final report and recommendations to the 23rd ittc by the specialist
committee on cavitation induced pressures. 23rd International Towing Tank
Conference, Venice, Italy, 2:417459, 2002.
[13] A.H. Koop, H.W.M. Hoeijmakers, G.H. Schnerr, and E.J. Foeth. Design
of a twisted cavitating hydrofoil using a barotropic ow method. In 6th
International Symposium on Cavitation, 2006.
7 Acknowledgments 18

[14] G. Kuiper. Some experiments with specic types of cavitation on ship


propellers. Journal Fluid Engineering, 1:105114, 1982.
[15] R.F. Kunz, D.A. Boger, D.R. Stinebring, T.S. Chyczewski, and H.J. Gibel-
ing. A preconditioned navier-stokes method for two-phase ows with ap-
plication to cavitation prediction. American Institute of Aeronautics and
Astronautics, 99-3329, 1999. AIAA 99-3329.
[16] K.R. Laberteaux and S.L. Ceccio. Partial cavity ows pt 2. cavities form-
ing on test objects with spanwise variation. Journal of Fluid Mechanics,
431:4363, 2001.
[17] Q. Le, J.P. Franc, and J.M. Michel. Partial cavities: Global behavior and
mean pressure distribution. Journal of Fluids Engineering, 115:243248,
1993.

[18] T.J. OHern. An experimental investigation of turbulent shear ow cavita-


tion. Journal of Fluid Mechanics, 215:365391, 1990.
[19] Y.L. Young and S.A. Kinnas. A bem for the prediction of unsteady mid-
chord face, and/or back propeller cavitation. Journal of Fluids Engineering,
123(2):311319, 2001.

You might also like