You are on page 1of 12

Ocean Engineering 253 (2022) 111174

Contents lists available at ScienceDirect

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

Experimental investigation of free-surface effects on flow characteristics of


a torpedo-like geometry having a cambered nose
F. Sarigiguzel a, A. Kilavuz a, M. Ozgoren b, T. Durhasan c, B. Sahin d, *, L.A. Kavurmacioglu e,
H. Akilli a, E. Sekeroglu f, B. Yaniktepe f
a
Cukurova University, Engineering Faculty, Department of Mechanical Engineering, Adana 01330, Turkey
b
Necmettin Erbakan University, Engineering Faculty, Department of Mechanical Engineering, Konya 42090, Turkey
c
Adana Alparslan Turkes Science and Technology University, Faculty of Aeronautics and Astronautics, Department of Aerospace Engineering, Adana 01250, Turkey
d
Istanbul Aydin University, Engineering Faculty, Department of Mechanical Engineering, Istanbul 34295, Turkey
e
Istanbul Technical University, Facultyt of Mechanical Engineering, Istanbul, Turkey
f
Osmaniye Korkut Ata University, Engineering Faculty, Department of Energy Systems Engineering, 80000, Osmaniye, Turkey

A R T I C L E I N F O A B S T R A C T

Keywords: In this study, the free-surface effect in the wake region of a torpedo-like geometry having a cambered nose was
Free-surface investigated experimentally via Particle Image Velocimetry (PIV) and dye flow visualization methods. The
PIV Reynolds number was taken as Re = 2 × 104 and 4 × 104 while keeping the angle of attack (α =0 ). The torpedo-

Torpedo-like geometry
like geometry is submerged at various positions from the free-surface in the range of 0.50 ≤ h/D ≤3.50 in which
Turbulent flow
Vorticity
(h) is the distance from the water free-surface to the central plane of the geometry, and D is the diameter of the
midsection of the torpedo-like geometry. The effect of the free-surface of water on the flow properties is also
measured and presented comparatively in terms of normalized contours of instantaneous vorticity, time-
averaged streamwise velocity, vorticity contours, Reynolds stress correlation, turbulent kinetic energy, veloc­
ity fluctuations in streamwise and cross-streamwise, vortex shedding frequency, spectral density distribution as
well as pointwise variation. It is observed that the flow structures become asymmetric when the body is located
near the free-surface for the submersion ratio of h/D ≤1.0. With the largest submersion ratio of h/D = 3.50, the
flow characteristics are almost identical and symmetrical on both sides of the geometry.

1. Introduction the optimal nose profile for deep-water operations using velocity and
pressure distributions along with overall drags. Alvarez et al. (2009)
In the last few decades, many researchers have been working on the studied an optimized hull shape with conical nose and stern sections for
designing and analyzing of autonomous underwater vehicles (AUVs) Froude numbers Fr = 0.26 and 0.30, including the wave resistance and
because of advances in technology and computer sciences and suitable surface deformation by a first-order Rankine panel method and found
for use in hazardous areas. The importance of AUVs to the marine the reduction up to 25% in the estimated total resistance. Gao et al.
environment and their promising future has been recently reviewed by (2016) conducted a study of comparison between experiment and
Wynn et al. (2014). An early design of an optimized AUV was proposed computational fluid dynamics (CFD) in the range of 3.6 × 105 ≤ Re ≤
by Johnson (1980), who showed that the vehicle could be used effec­ 4.2 × 106 for AUV models consisting of different noses and tails with a
tively for underwater tasks. Huggins and Packwood (1994) proposed the cylindrical hull for optimization of AUV hull shape. They recorded that
optimum dimensions for a deep-diving long-range autonomous vehicle the optimal geometry could be basically a body with a long nose, a
having a streamlined hull shape with a length-to-diameter ratio of 4.5. minimal middle section, and an appropriate tail. While the presence and
Nouri et al. (2016) provided an optimization method based on the locations of propellers often overshadow the geometrical study of the
pressure distribution along with the geometry and reported that stern sections for high-speed operations, there are a few noticeable
increasing the volume of the nose section only increases the pressure variants such as the well-known submarine-like SUBOFF type Givler
drag component. Ignacio et al. (2019) provided an extensive study on et al. (1991), modified spheroid type Sarkar et al. (1997), elliptical type

* Corresponding author.
E-mail address: besirsahin@aydin.edu.tr (B. Sahin).

https://doi.org/10.1016/j.oceaneng.2022.111174
Received 13 April 2021; Received in revised form 3 March 2022; Accepted 23 March 2022
Available online 15 April 2022
0029-8018/© 2022 Elsevier Ltd. All rights reserved.
F. Sarigiguzel et al. Ocean Engineering 253 (2022) 111174

(AFTERBODY 1 and 2) Jagadeesh and Murali (2005); Sarkar et al. the wall surface friction and the pressure difference between the up­
(1997) and appendage type propellers mounted on the sides. Recently, stream and downstream of the vehicle during its movement. The drag
Kilavuz et al. (2022) had investigated the flow characteristics of an resistance is unsteady due to the instability of the wake flow region,
optimized torpedo-like geometry having an elliptical nose according to which consequently produces vibrations on the vehicle affecting its
Myring expression in the range of angles of attack of 0◦ ≤α ≤ 12◦ , sub­ maneuverability. Therefore, in order to achieve high precision control of
mersion ratios of 0.5≤h/D ≤ 3.50 for Re = 2 × 104 and 4 × 104. They AUVs or create a new design under the water surface, a detailed study
noted that the influence of the jet-like flow was more outstanding for α characterizing the flow structures in the wake region must be performed
= 4◦ and 8◦ having lower values of velocity fluctuations. For h/D = 0.5, to understand flow behavior in detail. Consequently, this study in­
0.75, and 1.0 in the range of 4◦ ≤α ≤ 12◦ , the free-surface was deformed vestigates the hydrodynamic characteristics of unsteady flow structures
with the partial extension of the geometry nose to outside of the water around a model of underwater vehicle, called torpedo-like geometry,
surface where the shed vortices in the wake area were prohibited to placed at various submersion ratios under the free-surface. It is worth
reattach the free-surface and progressed in the stern section. Another noting that this work was the first to execute an experiment and
preliminary CFD study of Kilavuz et al. (2021) for an optimized extensive spectrum analysis of the flow field utilizing PIV data for free-
torpedo-like geometry having an elliptical nose, three-dimensional surface flow conditions. The free-surface effect on the wake region of a
control volume was solved using LES (Large Eddy Simulation) turbu­ torpedo-like geometry with a cambered nose is investigated experi­
lence and VOF (volume of fluid) multiphase models for different sub­ mentally via the PIV technique and the dye flow visualization method
mersion ratios and angles of attack. They recorded that the flow for various submersion ratios (0.50≤h/D ≤ 3.50) and Reynolds numbers
characteristics such as streamlines and velocity contours due to the (Re = 2 × 10⁴ and Re = 4 × 10⁴), comparatively.
hydrodynamic deformation effect were asymmetrical structure in close
location to the free-surface in the range of 0.75≤h/D ≤ 1.50 at Re = 40 2. Experiments
× 103 and stated that the drag coefficient magnitudes increased as the
model was submerged closer to the free-surface position. 2.1. The model of torpedo-like geometry
For studies of the wake region, it was reported from the experimental
results of a submarine model that flow structures and the hydrodynamic The hull of the torpedo-like geometry is based on an axisymmetric
coefficients were negligible variations between Reynolds numbers 42 × body, which comprises three sections: The nose, mid-section, and stern.
106≤Re ≤ 131 × 106 Bridges et al. (2003). For slightly lower Reynolds In general, the mid-section has a cylindrical form, whereas the nose and
number range, for example, 1.05 × 105≤Re ≤ 3.67 × 105, the drag and stern have a profile of ellipse, sphere, or parabola. The torpedo-like
lift coefficient variation was noted to decrease around 25% with geometry designed with the help of Myring Equations (Myring, 1976).
increasing Reynolds numbers Jagadeesh et al. (2009) for the elliptical Myring profile is commonly used for the nose and stern sections for the
stern of AFTERBODY 1. With the introduction of the free-surface to the design of operational torpedoes and AUVs.
design environment, the effect of submersion ratio h/D is investigated The dimensions of the geometry have been designed, taking the
for various flow conditions, for example, to understand the flow inter­ features of available experimental facilities into account, such as water
action with the free-surface for complex geometries such as torpedo-like channel dimensions, camera field of view, free-stream velocity and
geometries, simplified geometries such as cylinders or spheres Derakh­ holding apparatus. Profiles of the nose, mid-section, and stern of the
shandeh and Alam (2019); Ozgoren et al. (2011); Oshkai and Rockwell torpedo-like geometry, are defined by the following equations, respec­
(1999); Sheridan et al. (1997); Benusiglio et al. (2015). The wave tively:
propagation and jet flow due to the placement at small submersion ratios
Df √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅2
direct the lower shear layer toward the free-surface Lee and Daichin yf = ± 2Dx − x 0 ≤ x ≤ Lf (1)
(2004). The jet flow also enhances the clockwise vortical structures 2D
causing elongated recirculating regions Kahraman et al. (2012). Drag D
and lift coefficients meanwhile displayed the same trend in all yc = ± Lf < x < (L − Ls ) (2)
2
free-surface studies of Tian et al. (2019) and Mitra et al. (2019), showing
⎛ √̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅̅ ⎞
a decrease in drag and lift with an increase of submersion with the [ ]2
D (x − L + L )
subsiding jet flow and smaller wave formation. Free-surface and wave (3)
s
ys = ± ⎝ − + D2 − ⎠ (L − Ls ) ≤ x ≤ L
2 2.35
drag effect on the spheres that were submerged in the proximity of the
free surface as a function of their depth and velocity was examined
As shown in Fig. 1a, the model of the torpedo-like geometry has a
experimentally by Benusiglio et al. (2015), and they recorded that the
total length (L) of 0.2 m and consists of a cambered nose having a pri­
free-surface of the wave drag was of the order of the hydrodynamic drag.
mary diameter (Df) of 0.045 m and a length (Lf) of 0.05832 m, a cylin­
Free-surface effects were negligible for higher submersions that were
drical mid-section having a diameter (D) of 0.04 m and a length Lc = L –
reported experimentally for a sphere such as h/D = 2.0 Dogan et al.
Ls – Lf, and a stern length (Ls) of 0.08 m. Here, the "±" sign in the
(2018) while for an AUV with SUBOFF geometry, free-surface effects
equations (1)–(3) should be used respectively "+" and "–" signs for
were this time negligible for h/D > 3.0 Salari and Rava (2017), and
obtaining y ≥ 0 and y < 0 values.
finally in the case of h/D = 3.3 for an AUV with a given angle of drift up
The examined torpedo-like geometry having a cambered nose is
to 18◦ free-surface effects were also negligible Amiri et al. (2019). Such
similar to the missile shroud separation structure and its nose creates a
studies are often conducted between small and intermediate Froude
boundary layer separation from the nose and reattachment to the fuse­
numbers, for example, Fr ≤ 0.5, since the AUVs operate around these
lage surface (Celi˙ker, 2015). Later, the retarded flow separation takes
Froude numbers resulting in the smaller wave. The total resistance and
place again so that the negative impact on flow characteristics of drag
flow characteristics particularly depend on the hydrodynamic design of
increment diminishes.
the AUV Javadi et al. (2015); Shariati and Mousavizadegan (2017). Flow
characteristic examination of a body having a cambered nose and
tampered stern using the PIV method as in the present study has not 2.2. Experimental setup
been encountered in the literature.
The experiments of dye flow visualization were conducted in a large-
1.1. Motivation of study scale water channel located in the Department of Mechanical Engi­
neering Laboratory of Cukurova University. The water channel test
AUVs consume energy to overcome a frictional resistance caused by section is constructed of transparent Plexiglas with a wall thickness of

2
F. Sarigiguzel et al. Ocean Engineering 253 (2022) 111174

laser sheets of 145 mJ pulse energy were sent perpendicular to the flow
at 15 Hz with a time delay predetermined for each free-stream flow
velocity value. The laser source was located underneath the water
channel and adjusted to illuminate the middle section of the torpedo-like
geometry as shown in Fig. 2b. A 12-bit CCD camera captured images of
the illuminated region shown by the field of view (FoV) with 1600 ×
1200 pixels, corresponding velocity field of 188mm × 140 mm
(4.7Dx3.5D) and total velocity vectors of 7326 (99 × 74). The captured
images were then processed using the Dantec Dynamic Studio software
to determine the instantaneous velocity field by using the Adaptive PIV
method and the grid size of 1.92mm × 1.92 mm (0.048D × 0.048D). The
estimated uncertainty for the PIV measurements in a flow field was re­
ported to be fewer than 2% (Ozgoren, 2006). Throughout the mea­
surement of 1000 double-frame instantaneous images were taken in
66.6 s. The Reynolds number, according to the length of the geometry, is
defined as Re = UL/υ where υ is the kinematic viscosity of water. The
experiments were done at Reynolds numbers (Re = U∞ L/υ) of Re = 2 ×
10⁴ and 4 × 10⁴ with corresponding free-stream velocities (U∞ ) of 0.1
m/s and 0.2 m/s, respectively. The Froude number is defined concerning
the channel water depth (hw ), and the distance between the free-surface
and center of the torpedo-like geometry (h) in Fig. 2 that are commonly
used in the literature.
U∞
Frhw = √̅̅̅̅̅̅̅̅ (4)
ghw

U∞
Frh = √̅̅̅̅̅ (5)
Fig. 1. (a) The geometrical definition of the torpedo-like geometry, and (b) a gh
photograph of the manufactured model.
According to Equations (4) and (5), the Froude numbers for all cases
are in the range of 0.0085 ≤ Fr ≤ 0.452, which is smaller than 1.0 and
0.015 m with upstream and downstream fiberglass reservoirs, and the
causes a subcritical flow regime indicating negligible magnitudes of
channel has the following dimension: a length of 8 m, a width of 1m, and
free-surface deformation. For all cases, the blockage ratio was calculated
a height of 0.75 m.. The hydraulic conditions throughout the experiment
to be less than 0.3%, indicating negligible surface deformation due to
are given in Table 1.
the placement of the model for h/D > 3.50.
The fluorescent dye produced by mixing the powder of Rhodamine B
The effect of the submersion ratio on the vortex shedding frequency
with water was injected into the test region utilizing the height of from
was examined and for this purpose, the variation of Strouhal number
the dye reservoir as providing identical velocity discharge through the
(St) was calculated using PIV data as follows;
holes on the body. Two small holes exist at the model surface, less than 1
mm in diameter, located at the upper and lower sides of the nose section St = fL/U∞ (6)
and two more holes at the stern section. These holes overlapped the laser
Here, f is the dominant vortex shedding frequency.
sheet emanating from the continuous laser beam generated by the
From the images obtained by the PIV system, the time-averaged flow
Spectra-Physics unit during the tests to eliminate yawing. The model
data can be determined using the following equations:
was fixed behind the front view with an annular rod of 0.005 m in
Time-averaged velocity components in the flow direction (x-direc­
diameter and 0.1 m length. It was seen that the fixing set influences on
tion);
the flow structure were insignificant. The laser sheet was set to pass
through the longitudinal vertical plane of the model as shown in Fig. 2a. 1 ∑N

Images of the hydrodynamics of the flow structures were then recorded < u(i, j)〉 = un (i, j) (7)
N n=1
with a SONY HD-SRI video camera. In order to better understand the
flow structure characteristics in the wake region of the torpedo-like Time-averaged velocity components perpendicular to the flow di­
geometry model, the PIV technique was used along with the dye visu­ rection (y-direction):
alization technique. The PIV experiments were carried out at the
1 ∑N
Advanced Fluid Mechanics PIV laboratory of Osmaniye Korkut Ata < v(i, j)〉 = vn (i, j) (8)
N n=1
University. The PIV technique is used to measure the instantaneous
velocity field of the desired region in the flow field. The obtained data Time-averaged vorticity;
could then be further analyzed and processed to determine the turbu­
lence behaviors. For the PIV measurement, the laser sheet was generated 1 ∑N
< ω(i, j)〉 = ωn (i, j) (9)
by a double-pulse YAG laser source created two laser beams at 1064 nm N n=1
wavelength and separated to 532 nm wavelength for safety. The isolated
Time-averaged fluctuating velocity values of the streamwise velocity
component, urms (the root-mean-square (RMS) of velocity fluctuations)
Table 1
and cross-streamwise velocity component vrms;
Hydraulic conditions throughout the experiments.
[ ]12
Water temperature 20 ◦ C 1 ∑N
< u(i, j)rms >= [un (i, j) − u(i, j) ]2 (10)
Free-stream velocities 0.1 m/s and 0.2 m/s N n=1
Water density 998.2 kg/m3
Dynamic viscosity of water 0.001002 kg/m.s
Turbulence intensity less than 1%

3
F. Sarigiguzel et al. Ocean Engineering 253 (2022) 111174

Fig. 2. Schematic representation of experiment setup for (a) Dye and (b) PIV flow visualization methods.

[ ]12 geometry. Fig. 3 shows qualitatively how the free-surface affects the
1 ∑N
< v(i, j)rms >= [vn (i, j) − v(i, j) ]2 (11) wake flow of the torpedo-like geometry model for Re = 2 × 10⁴. Critical
N n=1
appearance of the vortical flow is depicted with “A, B, C, D, E, F, G".
Time-averaged Reynolds stress correlations; During the experiment and animation of dye visualization, it is observed
that the flow separates from the cambered nose section at a point where
1 ∑N
the curve slope becomes negative as seen by “A" for all submersion ra­
(12)
′ ′
< u v (i, j)〉 = [[un (i, j) − u(i, j) ] ][[vn (i, j) − v(i, j) ] ]
N n=1 tios. The separated flow from the cambered nose section re-attaches to
the fuselage of the torpedo-like geometry, and then the boundary layer
Time-averaged normal stresses of velocity components; begins to develop again on the surface, and developing boundary layer
thickness becomes smaller than a geometry without a cambered nose. At
1 ∑N
(13) the beginning of the stern section, the flow separates creating the wake
′ ′
< u u (i, j)〉 = [[un (i, j) − u(i, j) ] ][[un (i, j) − u(i, j) ] ]
N n=1
region around the geometry. The rotational flow having lower pressure
creates the reversed flow in the wake. The velocity difference between
1 ∑N
the wake and free-surface flow occurs from the restricted flow zone due
(14)
′ ′
< v v (i, j)〉 = [[vn (i, j) − v(i, j) ] ][[vn (i, j) − v(i, j) ] ]
N n=1 to the free-surface effect to cause many small-scale vorticity concen­
trations presented by the symbols of C, D, and E. On the other hand, the
Time-averaged turbulent kinetic energy
( ) separated shear layer along the lower side of the wake flow region leads
1 ∑
N to the development of the Von Karman vortex streets for all submersion
(15)
′ ′ ′ ′
< TKE(i, j)〉 = 0.75
N n=1
[u u (i, j) + v v (i, j) ] ratios as seen from vorticity clusters F and G with almost identical dis­
tance for all h/D cases. It is observed that Kelvin-Helmholtz vortices,
Where N is the total number of the PIV images taken as a total of 1000 dominantly, form around the periphery of the reversed flow circle sur­
images and (i,j) is the location in the flow. rounding the torpedo-like model after h/D = 1.50. The flow structures
obtained for h/D ≥ 1.50 are similar and indicating that the free-surface
3. Results and discussion effect rate decreases with increasing submersion ratio, h/D, and it be­
comes negligible when h/D ≥ 2.0. As seen from the wake structures for
3.1. Observations for instantaneous flow patterns the submersion ratios h/D = 2.0 and h/D = 3.50, the snapshot images
exhibit virtually symmetrical flow patterns in the free shear layers with
Dye observations were performed at the different submersion ratios, “C, D, F, and G". For the submersion ratio range of 1.50≤h/D ≤ 3.50,
h/D, to visualize the flow structure downstream of the torpedo-like rotational flow region in the wake shows a combination from the
separated flow around the stern and then it travels further downstream.

4
F. Sarigiguzel et al. Ocean Engineering 253 (2022) 111174

Fig. 3. Dye visualization around the torpedo-like geometry in the range of 0.75≤h/D ≤ 3.50 at Re = 2 × 10⁴.

Meanwhile, the submersion ratio of h/D = 0.50, the dye visualization these submersion ratios beneath the free-surface owing to the 3-D effect
test was not performed due to the cambered nose and free-surface and the wave generation.
reflection, but a quantitative flow observation by the PIV method was Especially for h/D = 0.75 and 1.0, the negative vortices designated
conducted. with a dashed line from the top side of the geometry reach below the
Presentative instantaneous vorticity patterns (ωL/U∞ ) for Re = 2 × centerline of the geometry with the downward moving an augmented
10⁴ (left column) and Re = 4 × 10⁴ (right column) in the range of flow region and crumble the positive vortices displayed with a solid line
0.75≤h/D ≤ 3.50 are given in Fig. 4. For all submersion ratios h/D, the from the bottom layer, creating small-scale vorticity concentrations and
flow separation from the cambered nose section and its re-attachment to increasing the turbulence rate. This behavior is observed to be more
the hull is seen to be similar for both Reynolds numbers. For the sub­ dominant at Re = 4 × 10⁴ while persisting up to the submersion ratio of
mersion ratio of h/D = 0.50, the reattached flow forming in the bottom h/D = 1.0 for both cases. The shed vortices travel downstream following
side covers the periphery of the hull up to the end of the stern. For the the Von Karman vortex streets. It is demonstrated that the developing of
submersion ratio of h/D = 0.75, the separated flow from the nose section the Kelvin Helmholtz vortices are more prevalent in the flow region
re-attaches to the fuselage periphery further downstream. Small-scale between the free-surface and the wake for the submersion ratios of h/D
vorticity concentrations with positive magnitudes are also observed at = 0.75 and 1.0 in a more scattered level compared to the other

5
F. Sarigiguzel et al. Ocean Engineering 253 (2022) 111174

Fig. 4. Presentative instantaneous vorticity patterns (ωL/U∞ ) for Re = 2 × 10⁴ (left column) and Re = 4 × 10⁴ (right column) in the range of 0.50 ≤h/D ≤ 3.50.

submersion ratios of h/D = 1.5, 2.0 and 3.50. For the higher submersion around the nose and stern sections, and interaction of separated flows
ratios of h/D = 2.0 and 3.50, the separated flow from the periphery of from the bottom and upper stern sides, as well as reattaching the free-
the geometry progresses identically toward the wake region and leading surface, all contribute to the emergence of a three-dimensional wake
to develop a reversed flow around the stern. The separated flow en­ region. Additionally, it is observed from dye and PIV animations that the
velops the geometry and decreases the free-surface flow effect, which flow structures in the wake region underneath the free surface inherit­
becomes greater due to the momentum transfer at Re = 4 × 10⁴ than Re ably change three-dimensionally in a chaotic manner.
= 2 × 10⁴. The incoming flow toward the stern causes a flow separation
anywhere on the nose. It re-attaches to the hull, wrapping the whole
body for all submersion ratios and both Reynolds numbers. The devel­ 3.2. Time-averaged flow characteristics
oping boundary layer on the cylindrical portion begins to separate again
further downstream on the fuselage, and then the more apparent vortex- Fig. 5 presents distributions of the velocity vectors with time-
shedding happens from the bottom portion of the geometry. At Re = 4 × averaged normalized streamwise velocity component < u/U∞ >,
2
10⁴, instantaneous vorticity clusters (ωL/U∞ ) in the wake area according vorticity <ω L/U∞ >, Reynolds stress correlation <u v /U∞ > and tur­
′ ′

to the submersion ratios of 0.75≤h/D ≤ 3.50 interact at closer places. 2


bulent kinetic energy <TKE/U∞ > at submersion ratios, h/D of 0.75,
For the sake of 3-D flow structure interpretation, the torpedo-like ge­ 1.0, 1.50, 2.0, and 3.50 for Re = 2 × 10⁴. The results display that for the
ometry model was positioned longitudinally in the horizontal plane submersion ratios of h/D = 0.75 and 1.0, a higher level of the time-
during the experiment. Therefore, in the side view plane, the rotation of averaged streamwise velocity components < u/U∞ > is observed in
the vorticity takes place around the lateral axis. On the other hand, the the shear layer zone developed above the torpedo-like geometry due to
rotation of the vorticity in the plan view plane happens around the the restricted area formed by the upper surface of the model boundary
vertical coordinate system. In this sense, the vortical flow periphery of and the free-surface. Consequently, an augmented flow is formed in the
the torpedo is three-dimensional. Furthermore, the flow measurement gap between the free-surface and upper boundary of the model. The
was performed only x-y plane to minimize the free-surface disturbing augmented flow formed by the presence of the free-surface is observed
effect during the experiment. However, the restricted flow stream be­ to influence the wake region and turbulent flow statistics. Due to the
tween the torpedo-like geometry and free-surface, free-surface defor­ accelerated flow formation in the upper boundary of the model for h/D
mation, hydrodynamic buoyancy force effect, separated flow structures < 1.0, the wake region loses its symmetricity about the centerline of the

6
F. Sarigiguzel et al. Ocean Engineering 253 (2022) 111174

Fig. 5. Comparison of the time-averaged dimensionless streamwise components of the velocity vectors < u/U∞ >, vorticity contours <ωL/U∞ >, and Reynolds stress
2 2
correlation <u v /U∞ > and turbulent kinetic energy < TKE/U∞ for submersion ratios, h/D of 0.75, 1.0, 1.5, 2.0, and 3.50 at Re = 2 × 10⁴.
′ ′

torpedo-like model. On the other hand, for h/D ≥ 1.50, the wake region prominent as illustrated by the lower shear layer and occurs in the
restores its symmetrical flow structure, indicating that the free-surface dominant manner as given by instantaneous vorticity structures indi­
influence decreases with an increase in the submersion ratio, h/D. cated in the top third of Fig. 4. These contours demonstrate asymmet­
This variation trend is also confirmed by the time-averaged normalized rical flow patterns due to the free-surface effect. It is observed that the
vorticity < ωL/U∞ > patterns given in Fig. 6. The presented vorticity maximum value of each flow pattern migrates to the closer the end due
patterns for the considered submersion ratios (h/D) support that the to having a greater Reynolds number and increasing momentum trans­
free-surface effect gradually decreases as the submersion depth ratio (h/ port for Re = 4 × 10⁴. As demonstrated in the bottom row the optimized
D) is increased. For h/D = 1.50, 2.0, and 3.50, the shear layer extends stern geometry, the time-averaged streamwise velocity < u/U∞ > comes
beyond the wake region on both the upper and lower sides of the ge­ closer to the end of the stern with increasing magnitudes and the area of
ometry. It is observed that the peak magnitudes of < u v / U∞
′ ′
2
> and lower velocity contours at Re = 4 × 10⁴ than Re = 2 × 10⁴ shrinks in size.
<TKE/U2∞ > are located on the lower side of geometry for h/D ≤ 1.0 Fig. 7 displays the distributions of the time-averaged normalized
since vortex shedding occurs on the lower side, and vortices are directed Reynolds stress correlation < u v /U2∞ > in the range of 0.75≤h/D ≤
′ ′

toward the free-surface. The absolute difference in the maximum and 3.50 along vertical lines in the wake region (4.0≤x/D ≤ 7.0) depicted at
minimum values of the Reynolds stress correlation < u v / U∞
′ ′
2
> de­ the bottom right image shown in Fig. 6 for Re = 2 × 10⁴ and Re = 4 ×
creases by around 90% for h/D ≥ 1.50, confirming that the free-surface 10⁴. In the region where h/D is close to the free-surface, the highest
impact becomes insignificant in cases of those submersions. Conform­ value of the Reynolds stress correlation < u v /U2∞ > appears in the
′ ′

ably, as the submersion ratio (h/D) is raised, the turbulent kinetic en­ lower part of the wake region. Also, due to the free-surface effect, the
ergy < TKE/U2∞ > values fall by up to 20%. Reynolds stress correlation <u v /U2∞ > is not evenly distributed with
′ ′

The time-averaged flow patterns at h/D = 1.0 for Re = 2 × 10⁴ and respect to the symmetry axis of the torpedo-like model. The smallest gap
Re = 4 × 10⁴ are compared in Fig. 6. The maximum values of each between the top surface of the geometry and free-surface for h/D = 0.75
contour are provided for the sake of comparison and numerical valida­ and at x/D = 4.0, larger negative values of the <u v /U2∞ > occur owing
′ ′

tion purposes. As it is seen from the contours of the time-averaged ve­ to the accelerated and reversed flow over the stern. Near-symmetrical
locity fluctuations in streamwise < urms /U∞ > and cross-streamwise flow structures obtained for h/D ≥ 1.50 indicate that the free-surface
<vrms /U∞ > directions, Reynolds stress correlation <u v / U2∞ >, tur­ effect decreases by increasing the submersion ratio. At the largest sub­
′ ′

2
bulent kinetic energy <TKE/U∞ > and streamwise velocity components mersion ratio of h/D = 3.50, variation of the Reynolds stress correlation
<u/U∞ > in Fig. 6, their peak values at Re = 4 × 10⁴ move immediately < u v /U2∞ > having double peaks as negative and positive magnitudes
′ ′

to the bottom of the stern section where the vortex-shedding is reflects an identical aspect that supports the diminishing free-surface

7
F. Sarigiguzel et al. Ocean Engineering 253 (2022) 111174

Fig. 6. Presentation of the Reynolds numbers of Re = 2 × 10⁴ and Re = 4 × 10⁴ influences on the time-averaged velocity fluctuations in streamwise < urms / U∞ >,
2 2
and cross-streamwise <vrms /U∞ > directions, Reynolds stress correlation <u v /U∞ >, turbulent kinetic energy <TKE/U∞ > and streamwise velocity components <u/
′ ′

U∞ > at h/D = 1.0.

effect. Vortex formation recovery is faster at Re = 4 × 10⁴ than Re = 2 × and for both Reynolds numbers, but they are not in sequence concerning
10⁴ hence the Reynolds stress correlation < u v /U2∞ > at Re = 2 × 10⁴ is
′ ′
the location of the torpedo-like geometry due to the free-surface dis­
greater magnitudes at stations x/D = 6 and 7 than that of Re = 4 × 10⁴. turbing and unstable flow behaviors. When the distance of the central
The Reynolds stress correlation < u v /U2∞ > becomes zero for both
′ ′
axis of the torpedo-like geometry to the free-surface is h/D = 3.50, the
Reynolds numbers beyond the lower shear layer at h/D = 3.50 due to the time-averaged velocity component < u/U∞ > takes a negative value
nearly uniform flow condition. between locations x/D = 5.0 and 5.7. After that, it proceeds a zero value
Fig. 8 presents the variation of the time-averaged streamwise ve­ at x/D = 5.7 for the case of Re = 2 × 10⁴. The positions where the time-
locity component < u/U∞ >, root mean square of velocity fluctuations in averaged velocity component < u/U∞ > is zero along the x/D axis
streamwise velocity component <urms /U∞ > and cross-streamwise ve­ indicate the combination point of the free-shear layers separating from
locity component <vrms /U∞ >, according to submerging depth ratios of the surrounding of the stern known as a free stagnation point. On the
0.75 ≤ h/D ≤ 3.50 in the wake region along center lines starting from x/ other hand, submersion ratios of 0.75≤h/D ≤ 3.50 at Re = 4 × 10⁴, the
D = 5 at y/D = 0 for Re = 2 × 10⁴ and Re = 4 × 10⁴. The time-averaged stagnation points and zero magnitudes of the time-averaged velocity
streamwise velocity component increases, in sequence, quadratically component < u/U∞ > also occur very close to the end of the stern
around < u/U∞ ≥0.8 and 0.6 for Re = 2 × 10⁴ and Re = 4 × 10⁴, and because of the greater inertia force of the incoming flow. For both
then they continue to rise slowly until reaching free-stream value far Reynolds numbers, the time-averaged velocity fluctuations <urms /U∞ >
wake region. The reason for the enhancement is the momentum transfer and <vrms /U∞ > rise deeply after taking the largest values of each sub­
from the core flow region to the wake region. Variations of these fluc­ mersion ratio becomes nearly asymptotic for the < urms /U∞ > while
tuating parameters have similar trends in the range of 0.75 ≤ h/D ≤ 3.50 decreasing for the <vrms /U∞ >. Under the intensive free-surface

8
F. Sarigiguzel et al. Ocean Engineering 253 (2022) 111174

2
Fig. 7. Distributions of the time-averaged normalized Reynolds stress correlation < u v /U∞ > depend on the submersion ratios of 0.75≤h/D ≤ 3.50 along vertical
′ ′

lines (4.0≤x/D ≤ 7.0) for Re = 2 × 10⁴ and Re = 4 × 10⁴.

influence, the time-averaged velocity fluctuations in cross-streamwise the bottom portion of the free-shear layer, but the accelerated flow from
<vrms /U∞ > for h/D = 0.75 and 1.0 are larger than the case of h/D > the upper side region between the geometry and the free-surface arises
1.50. The distance h/D of the central axis of the torpedo-like geometry to the vortex-shedding toward the wake region as seen in the instantaneous
the free-surface affects the cross-streamwise velocity fluctuations <vrms / vorticity contours in Fig. 4, which yields the St = 2.22. The interaction of
U∞ > more than the streamwise velocity fluctuations <urms / U∞ >. the vortex-shedding around the circumference of the torpedo-like ge­
ometry at h/D = 1.0 augments due to the lessening effect of free-surface,
the dominant vortex-shedding frequency (f) corresponding to the
3.3. Spectral analysis Strouhal number ascends to the value of St = 2.58. The peak magnitude
of the spectral density (Su ), attenuates by approximately 41% as opposed
Distributions of spectral density (Su ) and the time history of to the case of h/D = 0.75. The incoming flow through the cambered nose
streamwise velocity (u) are given in Fig. 9 of 0.50 ≤h/D ≤ 3.50 at Re = 2 for the free-surface touching case at h/D = 0.5, the flow passing the body
× 10⁴. The power spectrum analysis was performed in the wake region at is directed mainly to the bottom side due to the blockage effect and
4015 points, and the dominant vortex-shedding frequency was deter­ lower hydrodynamic resistance; thus, the spectral density magnitude
mined for each submersion ratio. The contours of the spectral density resulted in Su = 18.62. When the submersion ratio is set to h/D = 0.75,
(Su ) were drawn at the dominant vortex-shedding frequencies (f). They the accumulated flow in front of the camber with scattered and absorbed
were determined, in sequence, f = 1.36, 1.11, 1.29, 1.37, 0.90 and 0.96 energy distributions travels downstream around the model, the accel­
Hz for h/D = 0.50, 0.75, 1.0, 1.50, 2.0, and 3.50 at Re = 2 × 10⁴ from the erated flow tracing over the model comes to the stern section, and moves
result of Fast Fourier Transform (FFT) of the time history of 1000 into the wake region downward by virtue of the prevailing deformation
instantaneous velocity components (u). The instantaneous streamwise and free-surface effect. Additionally, the more powerful stream from the
velocity component (u) is used to determine the spectral magnitude (Su). lower portion of the body transfers more momentum to the wake region,
Distributions of spectral density (Su ) in the flow field give information and the merging of all separated flow around the stern in the wake re­
about the coherence and strength of the dominant vortex-shedding gion augments the turbulence kinetic energy, yielding the highest
(Ekmekci and Rockwell (2010) and Durhasan (2020)). The wake flow spectral density magnitude of Su = 34.26 in the bottom portion of the
structures are affected by the free-surface for h/D ≤ 1.0 and the domi­ wake region. On the contrary, the submersion ratio is increased to h/D
nant vortex-shedding occurs because of the rolling up of vortices from = 1.0, the separated flow arrives at the wake region equivalently from
the lower shear layer. The highest value of the Strouhal number (St) the stern perimeter with more interaction, and the spectral density value
exists at the smallest submersion ratio of h/D = 0.50 as St = 2.72 drops to Su = 20.36. As the submersion ratio rises, the spectral density
because other submersion ratios are weaker than the vortex-shedding on magnitudes remain relatively constant around Su = 16, despite the fact
the lower shear layer of the model. The peak magnitude of the spectral that the dominating vortex shedding frequencies and related St numbers
density (Su ) increases when the submersion ratio (h/D) shifts to the change only little due to the reduced free-surface impact. The contours
value of h/D = 0.75. The more dominant vortex-shedding persists from

9
F. Sarigiguzel et al. Ocean Engineering 253 (2022) 111174

Fig. 8. Variation of the time-averaged normalized uniform velocity < u/U∞ >, time-averaged velocity fluctuations in streamwise <urms /U∞ > and cross-streamwise
<vrms /U∞ > velocity components in the range of 0.75≤h/D ≤ 3.50 along the centerline (y/D = 0) for Re = 2 × 10⁴ and Re = 4 × 10⁴.

of the spectral density for h/D ≥ 1.50 depict the vortices roll up alter­ cambered nose are experimentally investigated using PIV and fluores­
nately into the wake region from both sides of the torpedo-like geome­ cent dye methods under the effect of the free-surface in the range of
try. The influence of the free-surface on vortical flow instability is 0.50≤ h/D ≤ 3.50 at Re = 2 × 10⁴ and Re = 4 × 10⁴. The following
insignificant in the range of 1.50≤ h/D ≤ 3.50, and the peak magnitude remarks are made.
of the spectral density (Su ), has nearly the same difference. However, the The dye visualization experiments reveal that the flow separated
free-surface effect on the submersion ratio can transition depending on from the cambered nose reattaches to the hull and then separated again
the incoming flow velocity, geometry shape and size, angle of attack, at the stern section. The free-surface influences the structures of the
geometry appendages, blockage ratio and turbulence intensity level. For Kelvin Helmholtz vortices above the centerline of the model for h/D =
higher submersion ratios such as h/D = 1.50, 2.0, and 3.50, Strouhal 0.75 and 1.0.
numbers decline to the values of St = 1.74, 1.80 and 1.92, in sequence. The accelerated flow stream over the torpedo-like geometry beneath
The frequency shift that occurs when the body is close to the free-surface the free-surface is flourished at smaller submersion ratios of h/D < 1.50
may be managed with active or passive flow control methods, which can and thus the wake region exhibits asymmetry. At the higher submersion
be very useful for AUV design to minimize erratic force loading and ratios of h/D = 2.0 and 3.50, the free-surface effects are negligible as the
maneuvering constrictions. Depending on the submersion ratio, the wake region becomes identical on both sides of the centerline, similar to
free-surface effect dramatically changes the vortex-shedding frequency the wake structures observed under a uniform flow condition.
corresponding to St numbers and spectral density variations due to A double increase in the Reynolds number causes enhanced wake
surface wave deformation phenomena as similarly interpreted by recovery, leading to an overall smaller recirculating flow region close to
Benusiglio et al. (2015). the end of the stern due to the greater inertia force of the incoming flow.
The Reynolds number changes the critical magnitudes considerably
4. Conclusion while slightly affecting the turbulence intensity distributions in the
wake region.
In this study, flow behaviors of a torpedo-like geometry with a The powerful vortex-shedding in the bottom free-shear layer causes

10
F. Sarigiguzel et al. Ocean Engineering 253 (2022) 111174

Fig. 9. Distributions of spectra density (Su ) and time history of streamwise velocity (u) in the range of 0.50 ≤h/D ≤ 3.50 at Re = 2 × 10⁴.

the largest Strouhal number occurrence at h/D = 0.50 for Re = 2 × 10⁴, surface.
whereas the accelerated flow region for h/D = 0.75 decreases the
Strouhal number by approximately 18%. At h/D = 1.0, lowering the CRediT authorship contribution statement
free-surface effect yields a more significant Strouhal number. Further
submerged cases such as h/D = 1.50, 2.0, and 3.50, Strouhal numbers F. Sarigiguzel: Investigation, Visualization, Writing – original draft.
decrease remarkably compared to submerged cases of h/D ≤ 1.0. A. Kilavuz: Investigation, Visualization. M. Ozgoren: Conceptualiza­
Spectral density variations show the wake region deformation along tion, Investigation, Writing – original draft, Project administration. T.
with the dissipation of three-dimensional large, small and eddy vortices Durhasan: Investigation, Writing – original draft. B. Sahin: Writing –
under the free-surface flow effect depending on submersion ratios of the review & editing, Supervision. L.A. Kavurmacioglu: Investigation. H.
torpedo-like geometry. The obtained vortex shedding frequency values Akilli: Investigation. E. Sekeroglu: Investigation. B. Yaniktepe:
and spectral density distributions can be used to design the flow control Investigation.
method for the fuselage and stern section of the geometry such as adding
fixed, adjustable or flexible wings. Declaration of competing interest
In the present study, experimentally obtained data and detailed flow
physics may assist in designing adjustable or flexible wing mechanisms The authors declare that they have no known competing financial
to eliminate the non-uniform load distribution around the torpedo-like interests or personal relationships that could have appeared to influence
geometry when it travels close to the water free-surface and changes the work reported in this paper.
direction with a certain sideslip angle. It is also expected that, the ob­
tained data can be a reference to validate numerical studies in order to
design and optimize underwater vehicles to operate near the free-

11
F. Sarigiguzel et al. Ocean Engineering 253 (2022) 111174

Acknowledgments Javadi, M., Manshadi, M.D., Kheradmand, S., Moonesun, M., 2015. Experimental
investigation of the effect of bow profiles on resistance of an underwater vehicle in
free surface motion. J. Mar. Sci. Appl. 14, 53–60. https://doi.org/10.1007/s11804-
The authors wish to acknowledge the financial supports of the Sci­ 015-1283-0.
entific and Technological Research Council of Turkey (TUBITAK) under Johnson, H., 1980. Design features and test results of an unmanned free swimming
Contract No. 214M318 and Cukurova University Scientific Research submersible. In: OCEANS’80. IEEE, pp. 304–309. https://doi.org/10.1109/
OCEANS.1980.1151400.
Project Coordinators (BAP) under Contract No. FYL-2019-11552. The Kahraman, A., Ozgoren, M., Sahin, B., 2012. Flow structure from a horizontal cylinder
authors thank the Mechanical Engineering Department of Cukurova coincident with a free-surface in shallow water flow. Therm. Sci. 16, 93–107.
University and Advanced Fluid Mechanics PIV laboratory of Osmaniye https://doi.org/10.2298/TSCI110719087K.
Kilavuz, A., Ozgoren, M., Durhasan, T., Sahin, B., Kavurmacıoglu, L., Akıllı, H.,
Korkut Ata University, Turkey for using the water channel and mea­ Sarigiguzel, F., 2021. Analysis of attack angle effect on flow characteristics around
surement systems. torpedo-Like geometry placed near the free-surface via CFD. J. Polytech. 24 (4),
1579–1592.
Kilavuz, A., Sarigiguzel, F., Ozgoren, M., Durhasan, T., Sahin, B., Kavurmacioglu, L.A.,
References Sekeroglu, E., Yaniktepe, B., 2022. The impacts of the free-surface and angle of
attack on the flow structures around a torpedo-like geometry. Eur. J. Mech. B Fluid
Alvarez, A., Bertram, V., Gualdesi, L., 2009. Hull hydrodynamic optimization of 92, 226–243.
autonomous underwater vehicles operating at snorkeling depth. Ocean Eng. 36, Lee, S.J., Daichin, A., 2004. Flow past a circular cylinder over a free-surface: interaction
105–112. https://doi.org/10.1016/j.oceaneng.2008.08.006. between the near wake and the free-surface deformation. J. Fluid Struct. 19,
Amiri, M.M., Sphaier, S.H., Vitola, M.A., Esperança, P.T., 2019. URANS investigation of 1049–1059. https://doi.org/10.1016/j.jfluidstructs.2004.06.006.
the interaction between the free-surface and a shallowly submerged underwater Mitra, A., Panda, J.P., Warrior, H.V., 2019. The effects of free stream turbulence on the
vehicle at steady drift. Appl. Ocean Res. 84, 192–205. https://doi.org/10.1016/j. hydrodynamic characteristics of an AUV hull form. Ocean Eng. 174, 148–158.
apor.2019.01.012. https://doi.org/10.1016/j.oceaneng.2019.01.039.
Benusiglio, A., Chevy, F., Raphaël, É., Clanet, C., 2015. Wave drag on a submerged Myring, D.F., 1976. A theoretical study of body drag in subcritical axisymmetric flow.
sphere. Phys. Fluids 27, 072101, 2015. Aeronaut. Q. 27, 186–194. https://doi.org/10.1017/S000192590000768X.
Bridges, D.H., Blanton, J.N., Brewer, W.H., Park, J.T., 2003. Experimental investigation Nouri, N.M., Zeinali, M., Jahangardy, Y., 2016. AUV hull shape design based on desired
of the flow past a submarine at angle of drift. AIAA J. 41, 71–81. https://doi.org/ pressure distribution. J. Mar. Sci. Technol. 21, 203–215. https://doi.org/10.1007/
10.2514/2.1915. s00773-015-0343-0.
Celi˙ker, H.E., 2015. CFD Analysis of Missile Shroud Separation. Middle East Technical Oshkai, P., Rockwell, D., 1999. Free surface wave interaction with a horizontal cylinder.
University. J. Fluid Struct. 13, 935–954. https://doi.org/10.1006/jfls.1999.0237.
Derakhshandeh, J.F., Alam, M.M., 2019. A review of bluff body wakes. Ocean Eng. 182, Ozgoren, M., 2006. Flow structure in the downstream of square and circular cylinders.
475–488. https://doi.org/10.1016/j.oceaneng.2019.04.093. Flow Meas. Instrum. 17, 225–235. https://doi.org/10.1016/j.
Dogan, S., Ozgoren, M., Okbaz, A., Sahin, B., Akilli, H., 2018. Investigation of flowmeasinst.2005.11.005.
interactions between a sphere wake and free surface. J. Fac. Eng. Archit. Gazi. Univ. Ozgoren, M., Pinar, E., Sahin, B., Akilli, H., 2011. Comparison of flow structures in the
33, 1123–1133. https://doi.org/10.17341/gazimmfd.453552. downstream region of a cylinder and sphere. Int. J. Heat Fluid Flow 32, 1138–1146.
Durhasan, T., 2020. Flow topology downstream of the hollow square cylinder with slots. https://doi.org/10.1016/j.ijheatfluidflow.2011.08.003.
Ocean Eng. 209, 107518. https://doi.org/10.1016/j.oceaneng.2020.107518. Salari, M., Rava, A., 2017. Numerical investigation of hydrodynamic flow over an AUV
Ekmekci, A., Rockwell, D., 2010. Effects of a geometrical surface disturbance on flow moving in the water-surface vicinity considering the laminar-turbulent transition.
past a circular cylinder: a large-scale spanwise wire. J. Fluid Mech. 665, 120–157. J. Mar. Sci. Appl. 16, 298–304. https://doi.org/10.1007/s11804-017-1422-x.
https://doi.org/10.1017/S0022112010003848. Sarkar, T., Sayer, P.G., Fraser, S.M., 1997. A study of autonomous underwater vehicle
Gao, T., Wang, Y., Pang, Y., Cao, J., 2016. Hull shape optimization for autonomous hull forms using computational fluid dynamics. Int. J. Numer. Methods Fluid. 25,
underwater vehicles using CFD. Eng. Appl. Comput. Fluid Mech. 10, 599–607. 1301–1313. https://doi.org/10.1002/(SICI)1097-0363(19971215)25:11<1301::
https://doi.org/10.1080/19942060.2016.1224735. AID-FLD612>3.3.CO;2-7.
Givler, R.C., Gartling, D.K., Engelman, M.S., Haroutunian, V., 1991. Navier-Stokes Shariati, S.K., Mousavizadegan, S.H., 2017. The effect of appendages on the
simulations of flow past three-dimensional submarine models. Comput. Methods hydrodynamic characteristics of an underwater vehicle near the free surface. Appl.
Appl. Mech. Eng. 87, 175–200. https://doi.org/10.1016/0045-7825(91)90005-Q. Ocean Res. 67, 31–43. https://doi.org/10.1016/j.apor.2017.07.001.
Huggins, A., Packwood, A.R., 1994. The optimum dimensions for a long-range, Sheridan, J., Lin, J.-C., Rockwell, D., 1997. Flow past a cylinder close to a free surface.
autonomous, deep-diving, underwater vehicle for oceanographic research. Ocean J. Fluid Mech. 330, 1–30. https://doi.org/10.1017/S002211209600328X.
Eng. 21, 45–56. https://doi.org/10.1016/0029-8018(94)90028-0. Tian, W., Song, B., Ding, H., 2019. Numerical research on the influence of surface waves
Ignacio, L.C., Victor, R.R., Francisco, D.R.R., Pascoal, A., 2019. Optimized design of an on the hydrodynamic performance of an AUV. Ocean Eng. 183, 40–56. https://doi.
autonomous underwater vehicle, for exploration in the Caribbean Sea. Ocean Eng. org/10.1016/j.oceaneng.2019.04.007.
187, 106184. https://doi.org/10.1016/j.oceaneng.2019.106184. Wynn, R.B., Huvenne, V.A.I., Le Bas, T.P., Murton, B.J., Connelly, D.P., Bett, B.J.,
Jagadeesh, P., Murali, K., 2005. Application of low-Re turbulence models for flow Ruhl, H.A., Morris, K.J., Peakall, J., Parsons, D.R., Sumner, E.J., Darby, S.E.,
simulations past underwater vehicle hull forms. J. Nav. Architect. Mar. Eng. 2, Dorrell, R.M., Hunt, J.E., 2014. Autonomous Underwater Vehicles (AUVs): their
41–54. https://doi.org/10.3329/jname.v2i1.2029. past, present and future contributions to the advancement of marine geoscience.
Jagadeesh, P., Murali, K., Idichandy, V.G., 2009. Experimental investigation of Mar. Geol. 352, 451–468. https://doi.org/10.1016/j.margeo.2014.03.012.
hydrodynamic force coefficients over AUV hull form. Ocean Eng. 36, 113–118.
https://doi.org/10.1016/j.oceaneng.2008.11.008.

12

You might also like