You are on page 1of 14

Ocean Engineering 108 (2015) 336–349

Contents lists available at ScienceDirect

Ocean Engineering
journal homepage: www.elsevier.com/locate/oceaneng

An experimental study on flow separation control of hydrofoils


with leading-edge tubercles at low Reynolds number
Zhaoyu Wei a, T.H. New a,n, Y.D. Cui b
a
School of Mechanical and Aerospace Engineering, Nanyang Technological University, 50 Nanyang Avenue, Singapore 639798, Singapore
b
Temasek Laboratories, National University of Singapore, Engineering Drive 1, Singapore 117411, Singapore

art ic l e i nf o a b s t r a c t

Article history: Hydrodynamic characteristics of hydrofoils with leading-edge tubercles were experimentally investi-
Received 6 January 2015 gated in a water tunnel at a Reynolds number of Re¼1.4  104. Particle image velocimetry measurements
Accepted 4 August 2015 and particle-streak visualizations reveal that the tubercles improve flow separation behaviour. In
Available online 29 August 2015
particular, hydrofoils with larger wave amplitudes and smaller wavelengths tend to perform significantly
Keywords: better in flow separation control. Cross-stream flow measurements indicate that streamwise counter-
Hydrofoil rotating vortex pairs are generated over the tubercles and mitigate flow separation. Analysis confirms
Passive flow control that the tubercles function as vortex generators, due to their comparable heights relative to the
Leading-edge tubercles boundary layer thickness. The vortex pairs meander and interact with adjacent flows, causing the flow
Flow visualization
separation behaviour to be occasionally unstable, thus leading to variable flow separation region sizes.
Particle-image velocimetry
This suggests that measures may have to be taken to ensure the stability of the counter-rotating vortex
NACA634-021
pairs for more persistent and predictable improvements.
& 2015 Elsevier Ltd. All rights reserved.

1. Introduction their control surfaces. Lower Reynolds number flows have thicker
boundary layers and more prone to flow separations, thus often
1.1. Research background leading to poor performance.

In the past decades, drawing inspirations from nature for solutions


to engineering problems have seen increased interests from a number 1.2. Literature review
of new research areas. Lately, a novel passive flow control technique
derived from the morphology of humpback whales has attracted The idealized humpback whale flipper model with leading-
increasing attention of scientists from different areas (Fish et al., edge tubercles was tested by Miklosovic et al. (2004) in a wind
2011; Bolzon et al., 2015). The humpback whale is capable of perform- tunnel at Re¼5  105. They found that the wing with leading-edge
ing complex underwater manoeuvres to catch prey, including sharp, tubercles increases the maximum lift by 6% and delays the stall
high-speed, banked turns, as well as loops and rolls. This has been angle-of-attack by about 40%, as compared to one without tuber-
attributed to their pectoral flippers, where several tubercles are cles. Murray et al. (2005) showed that the three-dimensional
distributed along the leading-edge (Fish and Battle, 1995; Fish and flipper models could increase maximum lift values by 9% and 4%
Lauder, 2006). Since whales use pectoral flippers as hydrodynamic for sweep angles of 151 and 301 respectively, as compared to one
control surfaces (with a Reynolds number of about Re¼ 106), questions without leading-edge tubercles. Johari et al. (2007) investigated
are being raised as to whether these leading-edge tubercles could be the effects of tubercle wavelength and amplitude on the flow
applied on marine engineering devices. In fact, potential marine separation behaviour of full-span NACA634-021 models in a water
applications of leading-edge tubercles have been reviewed by Fish tunnel at Re¼ 1.83  105. It was found that wings with tubercles
et al. (2011). Some of these applications include rudders (Weber et al., reduce the maximum lift coefficient. However, the wings stall
2010), dive planes, stabilizers, surfboard skegs and propellers. Another more gently and the lift coefficient could be raised by as much as
potential application of these tubercles is on unmanned underwater 50% in the post-stall regime.
vehicle (UUV) control surfaces. These vehicles often cruise at low Flow visualizations using surface tufts and dye showed that the
speeds and tend to have lower Reynolds number flows associated with flow behind tubercle peaks remains attached to part of the wing
surface in the post-stall regime of the baseline hydrofoil (Johari et
al., 2007; Custodio, 2007). Similar trends for the lift and drag of a
n
Corresponding author. Tel.: þ 65 6790 4443; fax: þ 65 6792 4062. full-span NACA 0020 aerofoil with leading-edge tubercles were
E-mail address: dthnew@ntu.edu.sg (T.H. New). also observed by Miklosovic and Murray (2007) at Re¼2.74  105.

http://dx.doi.org/10.1016/j.oceaneng.2015.08.004
0029-8018/& 2015 Elsevier Ltd. All rights reserved.
Z. Wei et al. / Ocean Engineering 108 (2015) 336–349 337

In addition, using dye visualization technique, Custodio (2007) technique and evaluated the aerodynamic characteristics of wings
showed that each tubercle behaves like the leading-edge of a small with leading-edge tubercles at Re¼1.6  105, while the case with
delta wing and a counter-rotating vortex pair is created behind Reynolds number up to 3  106 was also considered by Dropkin
each tubercle. The analogy of a tubercle as a small delta wing was et al. (2012).
later confirmed by Stanway (2008). This study also measured the These above CFD simulations demonstrate that they can
lift and drag of the whale flipper model with and without provide information on the surface pressure distributions and 3D
tubercles at Re¼4.4  104 to 1.2  105, where it was found that flow structures which are not readily available from experimental
there is no significant improvement on the maximum lift except and theoretical investigations. Numerical simulations were con-
for the highest Reynolds number. In addition, the flow field around ducted by Rostamzadeh et al. (2014) to investigate the formation
the flipper model with tubercles was measured by particle image mechanism of streamwise vortices and surface pressure distribu-
velocimetry (PIV) technique in that study, where delays in flow tions were also captured in wind tunnel testing. They found that a
separation and formations of streamwise vortices were shown. skew-induced mechanism accounts for the generation of stream-
Furthermore, force characteristics of the whale flipper were tested wise vortices and the development of these vortices is accompa-
under dynamic conditions (i.e., as flapping foils). nied by flow separation in delta-shaped regions close to the
Later, Ozen and Rockwell (2010) applied leading-edge tuber- trailing-edge. More interestingly, they made use of critical-point
cles onto a flapping wing and PIV results showed that the theory to depict and explain the flow topology above the wing
tubercles could attenuate the positive and negative spanwise surface. Skillen et al. (2015) used large-eddy simulation (LES) and
flows along the wing surface, as well as demonstrating the onset studied infinite-span wings with leading-edge undulations at
and developments of the large-scale concentrations of positive Re ¼1.2  105 and the results showed that a secondary flow toward
and negative streamwise vortices at inboard locations. Goruney the regions behind the troughs are driven by the spanwise
and Rockwell (2008) have also implemented them on a delta pressure gradient. The stall delay properties behind the peak were
wing in an attempt to look into their flow effects on the flow explained by the entrainment of higher-momentum fluid and the
structures and topology along the wing surface. Hansen et al. re-energization of the boundary layer.
(2011) investigated the influence of leading-edge tubercles on the
aerodynamic characteristics of two NACA aerofoils. They sug-
gested that the tubercles behave in a manner similar to conven-
tional vortex generators and that the optimum tubercle 1.3. Current research motivation
amplitude and wavelength bear strong resemblances to the
optimum vortex generator spacing and height. Zhang et al. It should be noted that although much work has been done on
(2014) investigated the effects of leading-edge tubercles on the the implementation of leading-edge tubercles on wings, the flow
aerodynamic performance of a full-span NACA634-021 aerofoil in separation behaviour still has not been entirely clarified. Further-
a wind tunnel for a wide range of angles-of-attack. Particle image more, most of the previous studies on either aerodynamic or
velocimetry results confirmed the formation of streamwise hydrodynamic applications were performed at relatively high
counter-rotating vortex pairs, though it should be highlighted Reynolds numbers. It should be mentioned that in an earlier
that only one measurement location was considered. Further- study by the authors (New et al., 2014b), a full-span NACA634-
more, Delgado et al. (2014) tested a unity aspect-ratio wing with 021 wing with leading-edge tubercles has been investigated in a
a sinusoidal leading-edge at Re ¼1.4  105 using stereoscopic PIV water tunnel at Re ¼1.4  104, where PIV and particle-streak
to aid their numerical modelling. visualization techniques were applied to reveal some preliminary
More recent experimental studies include that by Custodio understanding on the flow separation behaviour. In this work
et al. (2015). They conducted a systematic testing of the hydro- however, the hydrodynamic characteristics of the hydrofoils with
dynamic forces on finite-span wings with leading-edge tubercles leading-edge tubercles are considered at a similar Reynolds
at Reynolds numbers up to Re¼4.5  105, as well as making use of number but with a wider variety of tubercle amplitudes and
three different planforms and a wide range of tubercle geometries. wavelengths. The aim is to reveal the fundamental flow beha-
Besides the above experimental works, Van Nierop et al. (2008) viour of tubercles and to relate them to the resulting flow
developed a mathematical model based on lifting-line theory to separation behaviour. It should also be kept in mind that the
investigate the stall characteristics of semi-span wings with flow-fields associated with these tubercles will be highly three-
tubercles. The control mechanism has been attributed to modifi- dimensional. Hence, to avoid the limitation of two-dimensional
cations to the pressure gradient on the suction side of the wing. PIV and particle-streak photography techniques, three laser sheet
Pedro and Kobayashi (2008) used detached-eddy simulation (DES) positions along the tubercle trough, peak and a location mid-way
to simulate the flow field over a flipper model at Re¼5.0  105 and between them were included during the streamwise experi-
they noticed that the tubercles could reduce the flow separation ments. For cross-stream experiments, the laser sheet planes were
near the tip of the flipper. Similar simulation work was also aligned at four different chordwise locations downstream of the
conducted by Weber et al. (2011) at Reynolds number between leading-edge to elucidate the nature of the streamwise counter-
5.05  105 and 5.2  105. Câmara and Sousa (2013) used DES rotating vortex pairs.

Table 1
Configurations and experimental conditions for hydrofoils with leading-edge tubercles.

Hydrofoil label A (mm) λ (mm) A/λ U0 (m/s) Re α Measurement planes

Streamwise Cross-stream

NACA634-021 baseline – – – 0.19 1.4  104 01, 101 151, 201 Mid-span –
A9λ18.75 9 18.75 0.48 0.19 1.4  104 01, 101 151, 201 Trough (z/s¼ 0.50) Mid (z/s ¼ 0.48) Peak (z/s ¼ 0.46) x/c¼ 0.12, 0.25, 0.38, 0.52
A9λ37.5 9 37.5 0.24 0.19 1.4  104 01, 101 151, 201 Trough (z/s ¼ 0.50) Mid (z/s ¼ 0.46) Peak (z/s ¼0.44) –
A1.875λ18.75 1.875 18.75 0.1 0.19 1.4  104 01, 101 151, 201 Trough (z/s¼ 0.50) Mid (z/s ¼ 0.48) Peak (z/s ¼ 0.46) –
A1.875λ37.5 1.875 37.5 0.05 0.19 1.4  104 01, 101 151, 201 Trough (z/s ¼ 0.50) Mid (z/s ¼ 0.46) Peak (z/s ¼0.44) –
338 Z. Wei et al. / Ocean Engineering 108 (2015) 336–349

2. Experimental details 600 mm (H)  1100 mm (L) (see Fig. 2). The two sides and bottom
walls were made of tempered glass, which allowed excellent
2.1. Design of hydrofoils optical access. In addition, there was another 40 mm (W) 
250 mm (H) tempered glass window at the back of water tunnel
The baseline hydrofoil in the present study was based on a for cross-stream measurements/visualizations. The free-stream
NACA634-021 profile, which has been studied by other researchers velocity was regulated by an axial pump with a frequency-
(Johari et al., 2007; Zhang et al., 2014) at higher Reynolds numbers inverter speed controller. Before entering into the 4-to-1 ratio
due to its close resemblance to the humpback whale pectoral flipper. contraction section and test-section, flow was conditioned by
Two sets of aluminium hydrofoils with tubercles were considered several honeycomb structures and fine-mesh screens to ensure
(see Table 1). The mean chord and span for all hydrofoils are low free-stream turbulence intensity. In the present study, the
c¼75 mm and s¼ 300 mm, respectively. For the first set, the wave- free-stream turbulence intensity was determined to be approxi-
length is λ ¼0.25c, and amplitudes are A¼0.025c and 0.12c. For the mately 1.1% at U0 ¼0.19 m/s from time-averaged PIV measure-
second set, the wavelength is λ ¼0.5c, with amplitudes remaining the ments taken for the region-of-interest. In addition, the uncertainty
same with the first set. These amplitudes and wavelengths fall within in free-stream Reynolds number has also been ascertained to be
the range of values associated with humpback whale flippers (Fish approximately 71.2%. All experiments were conducted at
and Battle, 1995; Johari et al., 2007). The design methodology for the U0 ¼ 0.19 m/s and the corresponding Reynolds number was
hydrofoils was similar to that used by Lohry et al. (2012). A Cartesian Re¼ U0c/ν ¼1.4  104, where c is the mean hydrofoil chord and ν
coordinate system was first defined with its origin located at the is water kinematic viscosity at working conditions. Note that lift
lower end of the mean leading-edge (see Fig. 1). x is on the lower end and drag force measurements were not performed here, as the
face pointing towards the trailing-edge, z was on the leading-edge present focus would be on the flow separation behaviour.
and pointing upwards, and y is determined by the right-hand law. To mount the hydrofoils for testing, two flat transparent acrylic
The trailing-edges of hydrofoils are located at xTE ¼ cand the plates were installed beneath free-stream water surface and the
leading-edge with n sinusoidal waves can be described as hydrofoils were installed between these two flat plates. Note that
  the upper plate was covered by black adhesive papers to reduce
2nπ z
xLE ¼ A sin ; 0 rz r s: ð1Þ laser reflection and to provide a black background for greater
s
contrast against the illuminated particles. In order to control the
where A is the amplitude of the sinusoidally-shaped tubercles and angles-of-attack precisely, a computer-controlled high-torque
z is the coordinate along the hydrofoil span. stepper motor (MDrives34ac Hybrid) was mounted rigidly to
To create the hydrofoils with tubercles, cross sections of the the upper end-plate, with its shaft coupled to the test hydrofoil
baseline hydrofoil was modified by a non-linear shearing trans- (see Fig. 2(a)). The angle-of-attack, α, was positive in the clockwise
formation which maintained the leading-edge radius of the base- direction from the top-view.
line hydrofoil, position of maximum thickness, profile behind
maximum thickness point and continuity of the cross section 2.3. Flow visualization and particle image velocimetry procedures
profile over the maximum thickness point. The nonlinear shear
transformation equations used are To evaluate the effects of leading-edge tubercles on the hydro-
( dynamic characteristics of the hydrofoils, both particle-streak
x þ 0:5xLE ð1 þ cosðπ x=0:3cÞÞ; 0 r x o 0:3c photography and PIV techniques were applied. They were used
x1 ¼ ; ð2Þ
x; x Z 0:3c to show the development of the vortices and flow separation
regions in both x–y and y–z planes. Fig. 2 shows the schematic
where x is the abscissa of baseline profile and x1 is abscissa of the
diagram of particle-streak photography experimental setup in
modified cross section. The derivatives of Eq. (2) can be expressed
both cross-stream and streamwise experiments.
as
For particle-streak photography experiment, 20 μm polyamide
(
dx1 1  5xLE π sinðπ x=0:3cÞ=3c; 0 r x o 0:3c particles were uniformly seeded in water. A 2 W, 532 nm
¼ : ð3Þ continuous-wave diode-pumped solid-state laser in conjunction with
dx 1; x Z 0:3c
beam-steering and sheet-forming optics was used to create a 2 mm
Note that Eq. (3) maintains the lead-edge radius, as well as the thick laser sheet to illuminate these particles. Using relatively long
smoothing of the two sections over maximum thickness point and exposure timing, the streaks of these particles were captured by a
boundary conditions. The dimension definitions, as well as the Canon 550D Digital Single-Lens-Reflex (DSLR) camera with a Canon
Trough-plane and Peak-plane cross sections obtained along the EF 50 mm f/1.4 USM lens. For streamwise visualization (x–y plane),
peak and valley of the sinusoidally-shaped leading-edge are the laser sheet was aligned perpendicular to the hydrofoil span. The
shown in Fig. 1. camera was positioned beneath the test section (Fig. 2(b)). For cross-
stream visualizations, the laser sheet was aligned perpendicular to
2.2. Water tunnel facility free-stream direction, and the camera with a 100 mm f2.8 lens was
located behindthe small window at the back of the water tunnel. For
The experiments were conducted in a low-speed recirculating streamwise visualization, the laser sheet was aligned at three different
water tunnel, where the test section size measured 450 mm (W)  spanwise positions named as Trough-plane, Mid-plane and Peak-

Fig. 1. (a) 2D plan view and (b) 3D view of a hydrofoil with tubercles, (c) Trough- and Peak-plane cross sections obtained by the non-linear transformation method.
Z. Wei et al. / Ocean Engineering 108 (2015) 336–349 339

Fig. 2. Schematics of the experimental setup and apparatus for (a) cross-stream and (b) streamwise particle-streak photography/PIV experiments.

plane. Trough-plane is the spanwise location that coincides with the directions, resulting in 99  74 vectors in each velocity map. The
middle of all troughs present on the leading-edge of hydrofoils, which vorticity component ωz was evaluated by a central-differencing
also corresponds to the mid-spans of all hydrofoils. The Peak-plane is scheme on the PIV measured velocity maps. Based on the PIV
the spanwise location coinciding with the peak point closest to the measurement procedures, velocity map resolution was approxi-
Trough-plane in the lower half parts of the hydrofoils. The Mid-plane mately 1.25 mm/vector. In addition, moving-average validation
is at the inflection point of the same tubercle between the Trough- was applied to validate vectors by comparing each vector with
plane and Peak-plane. For cross-stream experiments, the laser sheet the average of other vectors in a defined neighbourhood. Erro-
positions were referenced from the quarter-chord location. neous vectors would be replaced with vectors interpolated from
For particle-streak photography experiments, the DSLR camera surrounding ones. According to the above post-process techniques,
was connected to a computer and controlled remotely by a the uncertainty levels of measured velocity components were
proprietary software, which provided full user control over the limited to within 7 1% (Keane and Adrian, 1992). Considering
camera shutter, aperture and ISO settings. Images were taken the flow conditions, experimental setup (camera, laser and stepper
using an exposure time of Δt¼ 1/25 s. When the images were motor) and image processing, the total experimental uncertainty
taken at this exposure time, they could indicate the overall flow in the velocity measurements and vorticity calculation were
separation bubbles and instantaneous flow-fields. It should be estimated to be approximately 3.6% and 4%, according to the
noted that 30 s videos were also taken alongside the images to principles described by Moffat (1988).
assess the unstable nature of flow-fields as well, which had proved For cross-stream PIV cases, the laser sheet was aligned normal
to be highly useful in shedding light upon the flow structures to the free-stream direction, and a 105 mm f2 lens was used
associated with different highly dynamic flow fields (New and Tsai, instead. Time interval between the two frames was chosen such
2007; New and Tsovolos, 2011, 2012; Shi and New, 2013). that most of particles would remain illuminated by the laser sheet
Particle image velocimetry was used to measure the velocity and within this time interval. As such, a time-interval of 10,000 μs was
vorticity fields. The setup for PIV experiment was almost the same used and a total of 1000 image pairs were captured for each case.
with particle-streak photography experiment in Fig. 2. The continuous The physical measurement window was approximately
wave laser was replaced by a Litron double-pulsed Nd:YAG laser with 130 mm  100 mm. The raw datasets were processed by using
a maximum energy of 200 mJ/pulse. The instantaneous particle the same procedure as in streamwise experiment, resulting in
images were captured by a Dantec Dynamics FlowSense 2M/E 99  74 vectors in each velocity map. The vorticity component, ωx,
double-frame CCD camera attached to a Nikon f/1.4D 50 mm lens, was also evaluated from the PIV velocity maps.
which was synchronized with the double-pulsed Nd:YAG laser by a
Dantec Dynamics PIV Timer Box connected to a workstation equipped
with timer and image-grabber boards. Both streamwise and cross- 3. Results and discussion
stream PIV experiments were carried out to interpret the flow
dynamics. Similar measurement procedures can be found in a wide Particle-streak flow visualization images highlight the charac-
variety of different flow scenarios previously (Soria et al., 2003; New teristic flow features for hydrofoils with leading-edge tubercles
and Tay, 2006; New and Tsovolos, 2009; Shi et al., 2014; New et al., and provide a first-hand appreciation of the flow fields. Since
2014a, 2014b, take for instance). For streamwise cases, the three these images only show the flow patterns in relatively short time
measurement planes corresponded to the Trough-plane, Mid-plane durations, it is difficult to shed light on the mean near-field flow
and Peak-plane in particle-streak photography experiments. A total of structures. Thus, mean flow field from PIV data is used to reveal
615 image-pairs (i.e., instantaneous velocity fields) were captured for the overall flow separation. The rationale behind these reported
each angle-of-attack to ensure satisfactory convergence in the mean results is to illustrate the important flow features and physics
flow field. The physical measurement window was approximately associated with tubercles, as well as to serve as comparisons
120 mm  90 mm, mapped out by the 1600  1200 pixel CCD between different tubercle configurations.
camera.
The time interval between two frames was 1450 ms, which was 3.1. Flow control behaviour for hydrofoil A9λ18.75
set smaller than the time required for the particles to translate 20%
of the interrogation window (i.e. 32  32 pixel). During post- In this section, to reveal the fundamental mechanism and flow
processing procedures, 2-step refinement multi-grid cross-corre- separation behaviour of hydrofoils with leading-edge tubercles,
lation technique with initial and final interrogation window sizes the A9λ18.75 hydrofoil with a relatively larger tubercle amplitude
of 128 pixel2 and 32 pixel2 were applied, where the square inter- and smaller wavelength is selected for a closer inspection. As its
rogation areas overlapped 50% in both horizontal and vertical tubercle amplitude-to-wavelength ratio of 0.48 is relatively large,
340 Z. Wei et al. / Ocean Engineering 108 (2015) 336–349

Fig. 3. Instantaneous streamwise particle-streak photography flow fields along the upper surface at α ¼01, 101, 151 and 201 for (a) baseline NACA634-021 hydrofoil along its
mid-span, (b) A9λ18.75 hydrofoil along its Trough-plane, (c) Mid-plane, and (d) Peak-plane.

very pronounced streamwise counter-rotating vortex pairs (CVPs) When the leading-edge tubercles are introduced, the relatively
would be produced in the presence of a free-stream. Particle- stable two-dimensional flows give way to unstable three-
streak photography images and mean vorticity field with stream- dimensional flow behaviour. As shown in Fig. 3(b), along the
lines from PIV results are presented in Figs. 3–6. These figures Trough-plane of the A9λ18.75 hydrofoil, the flow always separates
include the streamwise and cross-stream results at four angles-of- from the leading-edge for the four angles-of-attack. It has to be
attack of α ¼01, 101, 151 and 201. The results for the baseline mentioned that the separation regions are smaller than that of the
hydrofoil are also included as a benchmark for comparison baseline hydrofoil. In the particle-streak photography video, it can
purposes. be seen that the flow tends to reattach to the hydrofoil surface
For the baseline hydrofoil (Fig. 3(a)), adverse pressure gradient once separated flow is produced. In addition, no obvious recircu-
due to the relatively thick profile exerts a significant influence at lating flow can be observed from the video (i.e. not shown here for
the present low Reynolds number. Obvious flow separation can be the sake of brevity). In Fig. 4(b), strong separated shear layers
detected even at α ¼ 01, where the separation point is located at could be noticed at α ¼01 and 201. Closer inspection of the
approximately half-chord position. As the angle-of-attack streamlines shows that no flow separation bubble is generated.
increases to α ¼101, the flow has already separated from the As soon as the flow separates from the leading-edge, it reattaches
leading-edge. For angles-of-attack of α ¼ 151 and 201, very evident to the hydrofoil surface. In addition, there is a critical point
recirculating flows could be observed. It was described by Zhang observed in the mean streamlines in Fig. 4(b) for α ¼151. As
et al. (2014) that at Re¼2.0  105, the flow remains attached to described by Rostamzadeh et al. (2014), it is in fact a saddle point
most of the baseline NACA634-021 wing surface at α ¼191 (see showing flow separation in a delta-shaped region near the trailing
Fig. 3(b) in Zhang et al. (2014)). It could be discerned that at such a edge in three-dimensional cases.
low Reynolds number, such a hydrofoil will stall more easily than Along the Mid-plane of the A9λ18.75 hydrofoil shown in Fig. 3(c),
those at significantly higher Reynolds numbers (Johari et al., 2007; it is noticed that the mildly separated flow remains mostly attached to
Zhang et al., 2014). the hydrofoil surface at α ¼01, and the flow attaches to the entire
Note that the images in Fig. 3 only show the instantaneous flow hydrofoil surface at α ¼101. Even when the angle-of-attack increases
separation behaviour. To ascertain that these observations continue to to as high as α ¼201, the flow remains attached to the surface till
hold true in a more persistent manner, streamlines and vorticity fields about 1/3 c from the leading-edge. It can be observed from the video
derived from mean PIV data are presented in Fig. 4 to inspect the that no flow recirculation bubble is formed. It should be mentioned
resultant flow separation behaviour. For the baseline hydrofoil, it can that for this plane at α ¼151 and 201, the flow separation becomes
be observed that the PIV results agree well with particle-streak more unstable as well. Some images show the flow remaining
photography images (see Fig. 4(a)). At α ¼01, the flow separates from attached to most of the hydrofoil surface, while other images show
the hydrofoil surface at about half-chord location, and forms a small the flow detaching from the surface from about 1/3 c from the
flow recirculating bubble. At α ¼ 101, a significant separated shear leading-edge. Video playback also shows that the flow only reattaches
layer and flow recirculation bubble are generated. As the angle-of- to the hydrofoil surface intermittently. During the experiment, more
attack increases to α ¼ 151 and 201, large flow recirculation regions than 20 images were taken for each case and the one in Fig. 3(c) at
can be observed. These results are similar to those of other NACA α ¼15 and 201 is a typical representation of the observed flow
hydrofoils, such as the NACA0010 described by New et al. (2014a, separation behaviour. It can be seen from the mean streamlines and
2014b). vorticity results (Fig. 4(c)) that the level of flow separation control
Z. Wei et al. / Ocean Engineering 108 (2015) 336–349 341

Fig. 4. Mean streamlines and vorticity fields ðω z Þ of (a) baseline hydrofoil along its mid-span, as well as A9λ18.75 hydrofoil along its (b) Trough-plane, (c) Mid-plane and
(d) Peak-plane. For each laser sheet position, angles-of-attack of α¼ 01, 101, 151 and 201 are considered.

Fig. 5. Cross-stream particle-streak photography images along the upper surface of A9λ18.75 hydrofoil at α ¼01, 101, 151 and 201along (a) x/c¼ 0.12, (b) x/c¼ 0.25 and (c) x/
c ¼0.38 cross-stream locations.

appears to be more promising. At α ¼01, the flow separates from the remains mostly attached to the entire hydrofoil surface at α ¼01,
hydrofoil surface at about half-chord location and forms a very weak 101 and 151. As the angle-of-attack increases to α ¼201, the flow
shear layer and small flow recirculation bubble. Similarly for α ¼101, would still be attached to the hydrofoil surface till about 1/3c
151 and 201, only weak shear layers are generated and the flow location from the leading-edge. From the corresponding mean
reattaches to the whole hydrofoil surface. streamlines and vorticity field (Fig. 4(d)), no obvious separated
It is interesting to note that flow separation control along Peak- shear layer is detected and the flow attaches to the whole
plane appears to be better. The delay in flow separation along hydrofoil surface for all four angles-of-attack. This phenomenon
Peak-plane has already been reported by Johari et al. (2007) and is quite different from that observed by Zhang et al. (2014) along
Zhang et al. (2014). It can be noticed from Fig. 3(d) that the flow the Peak-plane at a higher Reynolds number of Re¼2.0  105,
342 Z. Wei et al. / Ocean Engineering 108 (2015) 336–349

Fig. 6. Mean cross-stream vorticity fields ðω x Þ of A9λ18.75 hydrofoil along (a) x/c ¼0.12, (b) x/c ¼0.25, (c) x/c¼ 0.38 and (d) x/c ¼ 0.52 cross-stream locations. Four angles-of-
attack of α ¼01, 101, 151 and 201 are considered.

where separation shear layers could be observed. Interestingly, will render them incoherent rapidly. In Fig. 5(c), as the laser sheet
there is a critical point lying along the hydrofoil surface at α ¼201, was moved to x/c ¼0.38 position, only weak CVPs can be observed
indicating the presence of strong three-dimensional flows. at α ¼ 01. In contrast, no CVPs can be observed at α ¼101, 151 and
As to how exactly leading-edge tubercles alter flow separation 201. However, evidence of spanwise flow still exists and the
characteristics, some researchers have attributed it to the genera- particle streaks show that they diverge and converge along
tion of streamwise CVPs (Miklosovic et al., 2004; Custodio, 2007; Trough-plane and Peak-plane, respectively.
Stanway, 2008; Hansen et al., 2011; Zhang et al., 2014 and To further ascertain these observations in Fig. 5, attention should
Rostamzadeh et al., 2014, take for instance). However, the forma- be paid towards the mean vorticity derived from PIV data. In Fig. 6,
tion mechanism of streamwise CVPs and their role in flow mean vorticity results agree well with the particle-streak photogra-
separation control remain to be fully revealed. As mentioned phy images. At x/c¼0.12 and 0.25 locations, these CVPs can be
previously, each tubercle behaves like a small delta-wing observed regularly, especially at α ¼01, 101 and 151. Compared to the
(Custodio, 2007 and Stanway, 2008) and thus a CVP is generated results in Fig. 5, it is found that these CVPs do not show up very well
over each tubercle in the presence of a free-stream. This is also the through the particle-streak photography technique at α ¼ 01. In Fig. 6
purpose of presenting cross-stream particle-streak photography (c), at α ¼01 and 101, regular CVPs can still be clearly seen. However,
images and mean PIV vorticity field in Figs. 5 and 6. To zoom into at α ¼151 and 201, some of the CVPs have already dissipated. To
these vortices, only the results for a few tubercles are shown. confirm how far these CVPs could travel along the hydrofoil surface,
However, the readers should keep in mind that the flow behaviour the laser sheet was also positioned at x/c¼0.52 location (see Fig. 6
for other tubercles is generally similar. (d)). Results show that the CVPs could not be discerned clearly at
In Fig. 5(a), the CVPs can be clearly discerned at α ¼ 101, 151 and α ¼01 and 101. For larger angles-of-attack of α ¼151 and 201, no
201 when the laser sheet was located at x/c ¼0.12 position. As the coherent vortex structures can be found.
laser sheet moved downstream to the quarter-chord position (i.e. As mentioned earlier, a streamwise CVP is created over each
x/c¼ 0.25), the vortex patterns reveal that no CVPs can be observed tubercle. It is also found that this CVP share similar characteristics
at α ¼01. However, it can be detected that the flow streaks with vortices formed by delta-wings and conventional vortex
converge along the Peak-plane but diverge along the Trough- generators (VGs). They grow in size but decrease in intensity,
plane. As the angle-of-attack increases to α ¼101, very regular which coincides with the results in Figs. 5 and 6. Fig. 7 depicts a
CVPs can be observed. Compared to the CVPs for α ¼ 101 at x/ sketch for the formation and evolution of these streamwise CVPs.
c¼ 0.12 (see Fig. 5(a)), the size of these CVPs increases and they As a result of spanwise flow arising from the variation in the
move away from each other, which agree well with the expected tubercle leading-edge sweep, streamwise CVPs are generated
behaviour of CVPs in general. along the leading-edge close to the peak for each tubercle (see
At higher angles-of-attack of α ¼151 and 201, some vortex cores Fig. 7(a)). From the end-view sketch (see Fig. 7(b)), the vortex on
of the CVPs cannot be discerned clearlyand is believed to be a the left side rotates clockwise, while the one on the right side
result of vortex interactions between adjacent CVPs. As the CVPs rotates counter-clock wise. Due to the tubercle geometry and
evolve, they tend to move away from each other. When two characteristics of conventional CVPs, each vortex rotates in a way
adjacent vortices near the Trough-plane move closer to each other that causes them to migrate towards the Trough-planes (Fig. 5(b)).
such that they interact intensely, enhanced viscous dissipations These CVPs drive flows to converge along the Peak- plane while
Z. Wei et al. / Ocean Engineering 108 (2015) 336–349 343

Fig. 7. Sketches of vortex structures produced by leading-edge tubercles. (a) Three-dimensional view, (b) end-view.

Fig. 8. Unstable flow separation caused by interactions between counter-rotating vortex pairs along the Mid-plane of A9λ18.75 hydrofoil at α ¼151. Both (a) and (b) flow
instances are captured using a short exposure time of 1/25 s at two different time points.

causing them to diverge along the Trough-plane, which agrees these CVPs interact with one another and dissipate, unstable flow
well with the behaviour described in Fig. 5. The two adjacent separation will be produced. Despite the mean vorticity results
vortices close to the Trough-plane will combine to create low depicting the CVPs as stable vortical entities, it is expected that
pressure and high pressure gradient at the troughs of tubercles. their instantaneous behaviour are not as stable as these in the
Furthermore, as the cross section along Trough-plane is relatively mean data. The size and relative position of each vortex core in the
blunt, the flow always separates from the leading-edge at the CVPs also vary over time. Figs. 11 and 12 show the distribution of
troughs (Custodio, 2007). these instantaneous CVPs over the tubercles at α ¼101 and 151. As
It is also clear from the results presented above that the CVPs the laser sheet was moved to a position farther behind at x/
interact with each other as soon as they approach closer to one c¼ 0.25, the vortices become far more complex and some of the
another. When they interact, unstable flow separation behaviour CVPs become far too weak or incoherent to be observed. Hence,
can be observed. This phenomenon is more evident at larger only the results for the two laser sheet positions at x/c¼ 0.12 and x/
angles-of-attack. For the present A9λ18.75 hydrofoil, the flow c¼ 0.25 locations are presented. In Fig. 11, the laser sheet position
separation becomes unstable at α ¼151 and 201. Fig. 8 shows is closer to the leading-edge, thus the CVPs can be clearly observed
two particle-streak photography images taken at two different over each tubercle. The vortices vary both in size and position, and
time points for the A9λ18.75 hydrofoil along the Mid-plane at overlapping of vortices can be clearly seen. As indicated in Figs. 11
α ¼151, where the unstable flow separation behaviour can be seen. and 12, typical vortex pairs from two adjacent CVPs along the
In Fig. 8(a), the flow attaches to most of the hydrofoil surface. In Trough-plane are circled and marked with numbers. For Pair 1, the
contrast, the flow only attaches to a small part of the hydrofoil positive vortex is below the negative vortex, while the positive
surface as shown in Fig. 8(b) and leads to a sudden flow vortex is over the negative vortex for Pair 2. For Pairs 3 and 4,
separation. interaction of the two adjacent negative and positive vortices
To further ascertain this unstable behaviour, instantaneous PIV becomes very pronounced. For the laser sheet position at x/c ¼0.25
vorticity fields for the baseline and A9λ18.75 hydrofoils along the (Fig. 12), the size and relative position of the vortices change more
Mid-plane at α ¼ 151 are also presented in Figs. 9 and 10, significantly. In Pairs 1 and 6, the positive vortex is above the
respectively. For the baseline hydrofoil, the flow separation is negative vortex, while the negative vortex is just above the
relatively stable with a series of strong separation vortices being positive vortex for Pairs 2 and 3. For Pairs 4 and 5, interaction of
produced. From the image-sequence, they also lead to relatively the positive and negative vortices along the Trough-plane can also
consistent flow separation region size (Fig. 9). In contrast, highly be clearly observed.
unstable flow separations are observed for the A9λ18.75 hydrofoil
along the Mid-plane, as shown in Fig. 10, where the separation 3.2. Effect of amplitude and wavelength for tubercles
region varies over time. In Fig. 10(a and b), strong separated
vortices can be noticed with significant separation region sizes. The effects of amplitude and wavelength for the leading-edge
On the other hand, Fig. 10(e and f) show that as soon as the flow tubercles on flow separation will be presented and discussed in
separates from the leading-edge, it reattaches back to the hydrofoil this section. These results can also be applied to understand the
surface at approximately half-chord location. The entire vorticity flow separation and vortex structures for high Reynolds number
field sequence also shows that the separation region becomes cases, such as the study by Johari et al. (2007). The Reynolds
smaller as the separated vortices become weaker. number used by Johari et al. (2007) was Re¼ 1.83  105 and
So far, it is known that the formation of CVPs accounts for flow substantially larger than Re¼1.4  104 in this work. However, the
separation control in hydrofoil with leading-edge tubercles. As large-scale vortical structures are fundamentally similar. This
344 Z. Wei et al. / Ocean Engineering 108 (2015) 336–349

Fig. 9. Time-sequenced vorticity fields (ωz) obtained for the baseline hydrofoil at α ¼151, where the first image was arbitrarily set to t¼ 0.0 s and subsequent images are at
t¼ 3.3 s apart. (a) t ¼ 0s (b) t ¼3.3s (c) t ¼ 6.7s (d) t ¼ 10s (e) t ¼ 13.3s (f) t ¼ 16.7s (g) t ¼20s (h) t ¼ 23.3s.

Fig. 10. Time-sequenced vorticity fields (ωz) obtained for the A9λ18.75 hydrofoil at α ¼151 along the Mid-plane, where the first image was arbitrarily set to t¼ 0.0 s and
subsequent images are at t¼ 3.3 s apart. (a) t ¼ 0s (b) t ¼3.3s (c) t ¼ 6.7s (d) t ¼ 10s (e) t ¼ 13.3s (f) t ¼ 16.7s (g) t ¼20s (h) t ¼ 23.3s.

Fig. 11. Instantaneous cross-stream vorticity fields (ωx) of A9λ18.75 hydrofoil along x/c ¼ 0.12 cross-stream location at angles-of-attack of (a) α ¼101 and (b) α¼ 151. (I), (II) and
(III) represent three different arbitrary time points.

phenomenon was also described by Erm (2003), where it was hydrodynamic performance improvement that can be gained
found that both flow patterns are very similar for Reynolds through wavelength variations. This means that the amplitude-
numbers differing by over an order of magnitude. to-wavelength ratio, A/λ, is also an important parameter (Hansen
Johari et al. (2007) suggested that as the tubercle amplitude et al., 2011). In fact, as the ratio of A/λ varies, the leading-edge
varies, maximum lift coefficient, stall characteristics and drag sweep angle is altered as well, which in turn modifies the
coefficient will vary correspondingly as well. In contrast, varying strength of CVPs. The streamwise PIV mean streamlines with
the wavelength has little effects. In recent research, it was vorticity field for the A1.875λ18.75 hydrofoil is shown in Fig. 13.
noticed that for a given amplitude, there is a limit to how much This hydrofoil has the same wavelength as the previous
Z. Wei et al. / Ocean Engineering 108 (2015) 336–349 345

Fig. 12. Instantaneous cross-stream vorticity fields (ωx) of A9λ18.75 hydrofoil along x/c ¼0.25 cross-stream location at angles-of-attack of (a) α ¼ 101 and (b) α ¼ 151. (I), (II)
and (III) represent three different arbitrary time points.

Fig. 13. Mean flow streamlines and vorticity fields ðω z Þ of A1.875λ18.75 hydrofoil along its (a) Trough-plane, (b) Mid-plane and (c) Peak-plane. For each laser sheet position,
angles-of-attack of α¼ 01, 101, 151 and 201 are considered.

A18.75λ18.75 hydrofoil, while the amplitude becomes much Along the Peak-plane, no flow recirculation bubble is detected
smaller. Similarly, three laser sheet positions and four angles- at α ¼101 and 151, and a moderately-sized recirculation bubble
of-attack are considered. In Fig. 13, although the flow separation close to the trailing-edge can be observed at α ¼ 201. Similar
control is worse than A9λ18.75 hydrofoil, it remains better than phenomenon can also be seen along the Trough-plane at α ¼101
the baseline hydrofoil. Along the Mid-plane at α ¼01, the flow and 151. Along the Mid-plane, the flow separation behaviour
separation point is at about half-chord location, similar to the becomes rather irregular. At α ¼101, the separation point is at
baseline case. Along the Trough-plane, the separation point about 1/3c from the leading-edge with a resulting large recircula-
moves slightly upstream at α ¼ 01, while it moves downstream tion region. As the angle-of-attack increases to α ¼ 151, the flow
for the Peak-plane at α ¼01. The flow separation region size surprisingly remains attached to most of the hydrofoil surface,
decreases as the laser sheet location moves from the Trough- with only a very small separation bubble produced. Interestingly, a
plane to the Peak-plane. large flow recirculation bubble can be observed for α ¼201, formed
346 Z. Wei et al. / Ocean Engineering 108 (2015) 336–349

Fig. 14. Mean flow streamlines and vorticity fields ðω z Þ of A9λ37.5 hydrofoil along its (a) Trough-plane, (b) Mid-plane and (c) Peak-plane. For each laser sheet position,
angles-of-attack of α¼ 01, 101, 151 and 201 are considered.

after the flow attaching to the hydrofoil surface till approximately small flow recirculation region is formed. As for α ¼151 and 201,
1/3c location. the regular flow separation region could not be observed, the
Fig. 14 shows the mean PIV streamlines with vorticity field for modified flow topology (critical points) reveals very unstable and
A9λ37.5 hydrofoil at three laser sheet positions and four angles-of- strong turbulent three-dimensional flow behaviours.
attack. The A9λ37.5 hydrofoil has the same tubercle amplitude Fig. 15 shows the mean PIV streamlines with vorticity field for
with its A9λ18.75 counterpart, even though the wavelength is the A1.875λ37.5 hydrofoil. Undoubtedly, the tubercles have lost
doubled. Compared to the results in Figs. 4 and 13, flow separation their effectiveness in terms of flow separation control. For all the
control for A9λ37.5 is better than A1.875λ18.75, but worse off than four angles-of-attack and three laser sheet positions, the flow
A9λ18.75. Similar to the results of A9λ18.75 in Fig. 4, very strong separation behaviour is almost the same as that of baseline
separated shear layers are generated along Trough-plane at the hydrofoil (see Fig. 4(a)). This once again confirms that
four angles-of-attack (see Fig. 14(a)). Close inspection of the amplitude-to-wavelength ratio, A/λ, plays an important role in
streamlines shows that the flow in fact remains attached to the flow separation control. With reference to the results presented,
hydrofoil surface with no recirculation bubble. In addition, a node flow separation improvement has a strong dependence on A/λ,
of detachment (ND) can be identified, which is particularly evident showing an order of A9λ18.75, A9λ37.5, A1.875λ18.75 and
at α ¼201. This ND marks the location where limiting streamlines A1.875λ37.5, in terms of worsening flow separation control
converge in three-dimensional flow field, which agree well with effectiveness. Note that this also coincides with decreasing order
the phenomenon revealed by Rostamzadeh et al. (2014). The flows of A/λ.
along the Peak-plane converge at ND, and they seem to be It was mentioned by Simons (2002) that low Reynolds number
emanating from the hydrofoil surface and then diverging in the boundary layer flows and separation bubbles are surprisingly
two-dimensional streamline figures along the Trough-plane. The complicated and difficult to predict. Hence, in the present work
flow even moves upstream, as shown in Fig. 14(a) at α ¼ 201. Along where a significantly lower Reynolds number of Re¼ 1.4  104 is
the Mid-plane, there is only a small flow recirculation region at considered, the boundary layer flows and separation bubbles will
α ¼201. The flow attaches to the whole hydrofoil surface at α ¼ 01, be more complicated. The results show that the flow separation
101 and 151. Compared with the results for A9λ18.75 hydrofoil (see could be significantly suppressed in both pre-stall and post-stall
Fig. 4(d)), the flow behaviour is more unstable and irregular along regimes for A9λ18.75, A9λ37.5 and A1.875λ18.75 hydrofoils, espe-
the Peak-plane. At α ¼01, there is no flow separation. As the angle- cially along the Mid-plane and Peak-plane. However, whether the
of-attack increase to α ¼ 101, the flow starts to separate from the lift and drag characteristics could be improved needs to be
hydrofoil surface at about 1/3c from the trailing-edge, where a confirmed in future experiments.
Z. Wei et al. / Ocean Engineering 108 (2015) 336–349 347

Fig. 15. Mean flow streamlines and vorticity fields ðω z Þ of A1.875λ37.5 hydrofoil along its (a) Trough-plane, (b) Mid-plane and (c) Peak-plane. For each laser sheet position,
angles-of-attack of α¼ 01, 101, 151 and 201 are considered.

This work highlights the formation of CVPs and effect of characteristics can possibly be achieved when this is coupled with
tubercles on the near-field hydrodynamic characteristics for a the best amplitude. In this study, based on the performance of
given hydrofoil profile (NACA634-021) at very low Reynolds flow separation control, the best tubercle configuration can be
number. From earlier research (Johari et al., 2007; Zhang et al., found in A9λ18.75 hydrofoil.
2014), it is clear that the tubercles generate CVPs. It has also been It can be seen from the above results that the vortex generating
argued by Van Nierop et al. (2008) that it is implausible for these mechanism for tubercles would become more prominent at larger
tubercles to act as vortex generators, since their wavelength and angles-of-attack. Hansen et al. (2011) suggested that the effective
amplitude are much larger than the boundary layer thickness, but device height to boundary layer ratio, heff/δ, is an important
instead they alter the pressure distribution on the hydrofoil sur- parameter to describe the tubercles. The value of heff increases
face, increasing the extent of flow attachment. for larger angle-of-attack, and the tubercles will perform more like
Despite the above view point, results in this work clearly show vortex generators. This parameter could also be used as a reference
that leading-edge tubercles create very regular CVPs and evolution to determine the optimum tubercle amplitude for a given hydro-
of these CVPs greatly affects the flow separation behaviour. As foil profile. For each angle-of-attack, heff represents the height of
shown in Fig. 7 that just like a small delta-wing, a CVP is created tubercles seen by free-stream, it is 2Asinα or (cPeak-plane - cTrough-
over each tubercle, and there are also several analogies between plane) sinα. δ is the boundary layer thickness obtained at a distance
these tubercles and vortex generators (Lin, 2002; Hansen et al., of A from the leading-edge of baseline hydrofoil in the mean PIV
2011). The CVPs produced by VGs were found to increase in size velocity field.
but decrease in intensity along the streamwise direction as shown The procedures for obtaining the boundary layer thickness
Figs. 5 and 6 (Godard and Stanislas, 2006). In fact, both VGs and will be briefly explained here. The cross section profile of baseline
performance enhancement with tubercles can be understood hydrofoil was firstly fitted in the mean velocity field. Then, a
through detailed flow patterns in the boundary layer (Zverkov point at the chordwise location of x ¼ A on the baseline hydrofoil
and Zanin, 2008). As described by Lin (2002), the device angle has surface was marked. A line passing this point and normal to the
been found to be an important performance parameter for a vortex hydrofoil surface was further drawn in the mean velocity field.
generator, and this parameter has also been suggested to describe Finally, the entire velocity values along this line were exported,
the leading-edge tubercles by Hansen et al. (2011). In this case, the and the boundary layer thickness corresponds to the distance
device angle is also the tubercle sweep angle, governed by the from the hydrofoil surface to a location on the line where velocity
ratio of A/λ. It is expected that there exists an optimal amplitude- is 99% of the free-stream velocity. The boundary layer thickness
to-wavelength ratio, and the best hydrofoil performance for hydrofoils with A ¼ 9 mm, was determined as 4.17 mm,
348 Z. Wei et al. / Ocean Engineering 108 (2015) 336–349

Acknowledgments

The authors gratefully acknowledge the support of the present


study under Singapore MINDEF Defense Innovative Research
Programme (Project Agreement number: 9012101770), as well as
G. C. Koh for his assistance in the experiments and data analysis.

References

Bolzon, M.D., Kelso, R.M., Arjomandi, M., 2015. Tubercles and their applications.
Journal of Aerospace Engineering, 04015013.
Custodio, D., 2007. The Effect of Humpback Whale-like Leading-edge Protuberances
on Hydrofoil Performance. Worcester Polytechnic Inst, Worcester, MA, M.S.
Thesis.
Câmara, J.F.D., Sousa, J.M.M., 2013. Numerical study on the use of a sinusoidal
Fig. 16. Dependence of heff/δ on angle-of-attack α for hydrofoils with leading-edge leading edge for passive stall control at low Reynolds number. In: 51st AIAA
tubercles. Results for two different wave amplitude are included. Aerospace Sciences Meeting, AIAA Paper 2013-0062.
Custodio, D., Henoch, C.W., Johari, H., 2015. Aerodynamic characteristics of finite
span wings with leading-edge protuberances. AIAA J. 53 (7), 1878–1893.
Dropkin, A., Custodio, D., Henoch, C.W., Johari, H., 2012. Computation of flowfield
around an airfoil with leading-edge protuberances. J. Aircr. 49 (5), 1345–1355.
5.96 mm and 7.2 mm for α ¼101, 151 and 201, respectively. Note Delgado, H.E.C., Esmaeili, A., Sousa, J.M.M., 2014. Stereo PIV measurements of low-
that these values are smaller than the tubercle amplitude. For aspect-ratio Low-Reynolds-number wings with sinusoidal leading edges for
hydrofoils with amplitude of A ¼ 1.875 mm, the thickness of improved computational modeling. In: Proceedings of the 52nd Aerospace
Sciences Meeting, AIAA, National Harbor, MD.
boundary layer was estimated to be 1.976 mm, 1.973 mm and Erm, L.P., 2003. Measurement of flow-induced pressures on the surface of a model
2.52 mm, for α ¼ 101, 151 and 201, respectively. Obviously, these in a flow visualization water tunnel. Exp. Fluids 35, 533–540.
values are close to the amplitude. Fig. 16 shows the dependence Fish, F.E., Battle, J.M., 1995. Hydrodynamic design of the humpback whale flipper. J.
of parameter heff/δ upon the angle-of-attack α for the hydrofoils Morphol. 225, 51–60.
Fish, F.E., Lauder, G.V., 2006. Passive and active flow control by swimming fishes
with tubercles considered in this work. It is evidently seen that and mammals. Annu. Rev. Fluid Mech. 38, 193–224.
for the two wave amplitudes, the parameter of heff/δ increases Fish, F.E., Weber, P.W., Murray, M.M., Howle, L.E., 2011. Marine applications
biomimetic humpback whale flipper. Mar. Technol. Soc. J. 45 (4), 198–207.
first and then decreases after reaching the maximum value at
Godard, G., Stanislas, M., 2006. Control of a decelerating boundary layer. Part 1:
approximately α ¼151. For A ¼ 0.12c, the value is around 0.7. As to Optimisation of passive vortex generators. Aerosp. Sci. Technol. 10, 181–191.
A ¼0.025c, heff/δ has a value around 0.5, which means that Goruney, T., Rockwell, D., 2008. Flow past a delta wing with a sinusoidal leading-
compared with large amplitude tubercles, small amplitude edge: near-surface topology and flow structure. Exp. Fluids 47 (2), 321–331.
Hansen, K.L., Kelso, R.M., Dally, B.B., 2011. Performance variations of leading-edge
tubercles works more like vortex generator with normal heff/δ tubercles for distinct aerofoil profiles. AIAA J. 49 (1), 185–194.
range of 0.1–0.5 (Lin, 2002). Johari, H., Henoch, C., Custodio, D., Levshin, A., 2007. Effects of leading-edge
protuberances on aerofoil performance. AIAA J. 45 (11), 2634–2642.
Keane, R.D., Adrian, R.J., 1992. Theory of Cross-correlation Analysis of PIV Images.
49. Applied Scientific Research, pp. 191–215.
Lohry, M., Clifton, D., Martinelli, L., 2012. Characterization and design of tubercle
leading-edge. In: Seventh International Conference on Computational Fluid
Dynamics, ICCFD. , Princeton, NJ.
4. Conclusions
Lin, J.C., 2002. Review of research on low-profile vortex generators to control
boundary-layer separation. Prog. Aerosp. Sci. 38, 389–420.
In this paper, flow separation control for hydrofoils with Moffat, R.J., 1988. Describing the uncertainties in experimental results. Exp. Therm.
Fluid Sci. 1 (1), 3–17.
leading-edge tubercles was experimentally conducted under a
Miklosovic, D.S., Murray, M.M., Howle, L.E., Fish, F.E., 2004. Leading-edge tubercles
very low Reynolds number of Re ¼1.4  104. Two tubercle ampli- delay stall on humpback whale (megaptera novaeangliae) flippers. Phys. Fluids
tudes and two wavelengths are considered to investigate their 16 (5), 39–42.
effects on the flow separation behaviour. On the other hand, Murray, M.M., Fish, F.E., Howle, L.E., Miklosovic, D.S., 2005. Effects of leading edge
tubercles on a representative whale flipper model at various sweep angles. In:
results are also presented to shed more light on the flow separa- Proceedings of the Fourteenth International Symposium on Unmanned Unteth-
tion control behaviour associated with the application of tubercles ered Submersible Technology (UUST), Autonomous Undersea Systems Inst., Lee,
in hydrofoils. The main findings obtained in the present research NH, Aug.
Miklosovic, D.S., Murray, M.M., 2007. Experimental evaluation of sinusoidal
are summarized as follows: leading-edges. J. Aircr. 44 (4), 1404–1407.
Under very low Reynolds number, the flow can separate rather New, T.H., Chan, Y.X., Koh, G.C., Chung, H.M., Shi, S.X., 2014a. Effects of corrugated
easily from the baseline hydrofoil surface. The tubercles behave aerofoil surface features on flow separation control. AIAA J. 52 (1), 206–211.
New, T.H., Tay, W.L., 2006. Effects of cross-stream radial injections on a round jet. J.
like small delta-wings, where a CVP is formed over each tubercle. Turbul. 7 (57), 1–20.
The CVPs significantly modify the boundary layer flow through New, T.H., Tsai, H.M., 2007. Experimental investigations on indeterminate-origin V-
chordwise and spanwise interactions. For the four hydrofoils with and A-notched jets, AIAA J. 4 (45), 828–839.
New, T.H., Tsovolos, D., 2009. Influence of nozzle sharpness on the flow fields of V-
tubercles, performance of overall flow separation control depends notched nozzle jets. Phys. Fluids 21, 1–8.
on the parameter of A/λ. For larger A/λ, flow separation control is New, T.H., Tsovolos, D., 2011. On the vortical structures and behaviour of inclined
better, while the tubercles almost lose their flow separation elliptic jets. Eur. J. Mech. B. Fluids 30 (4), 437–450.
control capabilities for the smallest A/λ. For the A9λ18.75 hydro- New, T.H., Tsovolos, D., 2012. On the characteristics of minor-plane inclined elliptic
jets. Exp. Therm. Fluid Sci. 38, 94–106.
foil, the CVPs are significantly unstable with strong interactions New, T.H., Wei, Z.Y., Goh, G.C., Cui, Y.D., 2014b. On the role and effectiveness of
with their adjacent neighbours along the Trough-plane. When streamwise vortices in leading-edge modified wings at low Reynolds number.
they dissipate, very unstable flow separation can be observed. In: Proceedings of the 16th International Symposium on Flow Visualization.
Okinawa, Japan.
Furthermore, the flow separation region varies with time, which is Ozen, C.A., Rockwell, D., 2010. Control of vortical structures on a flapping wing via a
particularly noticeable at angles-of-attack of α ¼151 and 201. The sinusoidal leading-edge. Phys. Fluids 22 (2), 021701.
CVPs are observed to be responsible for flow separation mitiga- Pedro, H.T.C., Kobayashi, M.H., 2008. Numerical Study of Stall Delay on Humpback
Whale Flippers. AIAA Paper No 2008-0584.
tions and the tubercles are deduced to take on the role of vortex Rostamzadeh, N., Hansen, K.L., Kelso, R.M., Dally, B.B., 2014. The formation
generators. mechanism and impact of streamwise vortices on NACA 0021 airfoil’ s
Z. Wei et al. / Ocean Engineering 108 (2015) 336–349 349

performance with undulating leading edge modification. Phys. Fluids 26, Van Nierop, E.A., Alben, S., Brenner, M.P., 2008. How bumps on whale flippers delay
107101. stall: an aerodynamic model. Phys. Rev. Lett. 100 (5), 054502.
Shi, Shengxian, New, T.H., 2013. Some observations in the vortex-turning behaviour Weber, P.W., Howle, L.E., Murray, M.M., 2010. Lift, drag and cavitation onset on
of noncircular inclined jets. Exp. Fluids 54 (11), 1614. rudders with leading-edge tubercles. Mar. Technol. 47 (1), 27–36.
Shi, Shengxian, New, T.H., Liu, Yingzheng, 2014. Flapping dynamics of a low aspect- Weber, P.W., Howle, L.E., Murray, M.M., Miklosovic, D.S., 2011. Computational
ratio energy-harvesting membrane immersed in a square cylinder wake. Exp. evaluation of the performance of lifting surfaces with leading-edge protuber-
Therm. Fluid Sci. 46, 151–161. ances. J. Aircr. 48 (2), 591–600.
Simons, M., 2002. Model Aircraft Aerodynamics. Special Interest Model Books, Zhang, M.M., Wang, G.F., Xu, J.Z., 2014. Experimental study of flow separation
Dorset, England, UK p. 112. control on a low Re aerofoil using leading-edge protuberance method. Exp.
Skillen, A., Revell, A., Pinelli, A., Piomelli, U., Favier, J., 2015. Flow over a wing with Fluids 55, 1710.
leading-edge undulations. AIAA J. 53 (2), 464–472. Zverkov, I., Zanin, B., 2008. Disturbances growth in boundary layers on classical and
Soria, J., New, T.H., Lim, T.T., Parker, K., 2003. Multigrid CCDPIV measurements of wavy surface wings. AIAA J. 46 (12), 3149–3158.
accelerated flow past an airfoil at angle of attack of 301. Exp. Therm. Fluid Sci.
27, 667–676.
Stanway, M.J., 2008. Hydrodynamic Effects of Leading-edge Tubercles on Control
Surfaces and in Flapping Foil Propulsion. Massachusetts Inst. of Technology,
Cambridge, MA, M.S. Thesis.

You might also like