CFDLV16 N5 P59 77
CFDLV16 N5 P59 77
CFD Letters
Journal homepage:
https://semarakilmu.com.my/journals/index.php/CFD_Letters/index
ISSN: 2180-1363
1 Department of Naval Architecture, Faculty of Engineering, Diponegoro University, Jl. Prof. Soedarto, S.H, Tembalang, Semarang 50275,
Indonesia
2 Program of Maritime Technology, School of Ocean Engineering, Universiti Malaysia Terengganu, Malaysia
3 Department of Mechanical Engineering, Universitas Negeri Semarang, Semarang, Indonesia
Article history: The deep-v planing hull is designed to operate at high speeds because most of the hull’s
Received 20 July 2023 weight is supported by the hydrodynamic lift acting on the hull base. Planing hull form
Received in revised form 23 August 2023 characteristics such as deadrise angle, chines, and extended stern significantly affect
Accepted 19 September 2023 the ship’s hydrodynamic performance. The addition of the interceptor is an innovation
Available online 3 January 2024
to reduce the total resistance of the ship by controlling the trim angle. However, the
form of the ship’s stern is not always the same; thus, it needs to be studied based on
the form of the ship’s stern. The extended stern form refers to modifying the hull
geometry at the rear, particularly the stern extension beyond its conventional length.
This research aimed to analyze the hydrodynamic performance of the interceptor at
the extended stern angle. Furthermore, Computational Fluid Dynamics (CFD)
simulations were performed to analyze the effect of the extended stern form. A
numerical model of the deep-V planing hull with variations of the stern extension was
developed, and the flow behavior around the hull was analyzed using CFD techniques.
Simulations were conducted under various operating conditions, including different
speeds and interceptor strokes. The results indicated that the extended stern's
different forms could affect the ship's resistance, trim, and heave. The reduction in
resistance was seen at moderate speeds, thereby reducing steep trim angles. The
greater the extended stern angle, the more significant the reduction in ship resistance
at Fr 0.58 by 26%. Likewise, the combination of interceptor and extended stern
experienced a decrease in resistance in the semi-displacement phase with a
percentage of 33% resistance, 66% trim, and 47% heave. The interceptor stroke (d)
depended on the boundary layer (h). The extended stern with angles of 10°, 20°, and
Keywords: 30° were found to have d/h ratios of 0.38, 0.37, and 0.34. However, it should be noted
Drag; Interceptor; Extended Stern; that extending the stern without interceptors and with interceptors at high speeds
Planing Hull; CFD; Boundary Layer could result in a dangerous increase in resistance on high-speed vessel.
*
Corresponding author.
E-mail address: samuel@ft.undip.ac.id (Samuel)
https://doi.org/10.37934/cfdl.16.5.5977
59
CFD Letters
Volume 16, Issue 5 (2024) 59-77
1. Introduction
Boats with planing hulls continue to be developed with various innovations such as fast patrol
boats, pilot ships, and Coast Guard patrol vessels [1]. In planing hull ships, most of the hull weight is
supported by hydrodynamic lift acting on the hull base at high speeds; thus, it has excessive trim [2].
Reducing ship resistance is a ship architect’s goal and a way to save energy. Some research had been
carried out by modifying the hull and the propulsion system but it required much money.
Stern appendages such as the interceptor [3], trim tab [4], and Hydrofoil [5] serve to reduce
resistance. The addition of an interceptor is an innovation on the ship’s stern that has been studied
extensively in recent years. The interceptor is a thin plate mounted perpendicular to the ship's stern.
It increases pressure under the stern by blocking flow through the hull and reduces the trim angle by
generating negative moments [3]. Tsai and Hwang conducted a model test to determine the effect
of the stern flap, interceptor, and a combination of both on the planing hull of the ship. The
integrated interceptor with a stern flap indicated more effective results in reducing resistance than
using each appendage [4]. Furthermore, Day and Cooper studied the effects of using interceptors on
sailing yachts. The interceptor showed a 10-18% reduction in resistance over a speed range of 8-20
knots [6].
Karimi et al., [1] conducted trials on a catamaran planing hull model in calm and wavy water
conditions. Applying an interceptor could reduce heave, pitch motion, and vertical acceleration.
Researchers were increasingly exploring the interceptor because of its extraordinary advantages.
Ghassemi et al., [7] conducted tests on a planing hull with an interceptor. The result was high
pressure generated between the interceptor and bottom, wet surface, and reduced resistance by the
interceptor. Seok et al., [3] discovered that the pressure at the bottom of the hull increased in
proportion to the interceptor's speed and height for the interceptor's blocking effect. Suneela et al.,
[8] considered that interceptor height played an essential role in ship performance. The installation
location of the interceptor was also an essential factor, as illustrated by the test model by Karimi et
al., [10] the ship’s resistance was clearly reduced by the interceptor located at the stern and midship,
and the maximum resistance reduction for mono hull and catamaran ships were 15% and 12%,
respectively. Sahin et al., [9] revealed the best interceptor position between chine and the centerline,
and the effect of interceptor height was directly related to ship speed. The magnitude of the lift
created by the interceptor usually depended on the height and flow velocity. However, other
parameters might affect the effectiveness of the interceptor [10].
Not all ships have a perpendicular stern form. Kim et al., developed three hull models with the
stern bottom of the VPS and VWC ships extended aft at 2.9 % LWL, while the VWS ships were at 4.3
% LWL. However, after being modified, there were no significant changes, and a porpoising
phenomenon occurred at a speed of 30 knots. Adding an auxiliary device at the stern prevented the
porpoising phenomenon [11]. As Mansoori et al., [12] pointed out, the interceptor applied intense
pressure aft of the ship. It also reduced the resistance and trim of the ship and increased the lift force
coefficient, which directly affected the porpoising instability. According to the findings, the
interceptor could entirely manage the porpoising issue.
Park et al., [2] conducted research using the Aragon 2 hull form as a development of the three
hulls using wave-controlled interceptors. As a result, pitch and heave could be reduced by 41.3% and
33.4%. According to studies on interceptors on regular waves, heave could be reduced by 16% to
18%. [13]. Interceptor research using the Aragon 2 hull form has also been carried out by Samuel et
al., with a 57% drag reduction in the close-to-chine position [14]. In other research with different
interceptor forms, including the v-form interceptor, the result was a 21% drag reduced at a Froude
number of 0.87 [15]. The choice of the optimal interceptor height greatly affected its efficiency [16].
60
CFD Letters
Volume 16, Issue 5 (2024) 59-77
The different forms of the stern certainly affected the effectiveness of the interceptor, such as the
deadrise angle and variation of the stern’s bottom, which was lengthened to reduce trim and
resistance at high speeds.
Yousefi et al., [17] revealed three CFD methods related to a planing hull, i.e., the Finite Element
Method (FEM), Finite Volume Method (FVM), and Boundary Element Method (BEM). Due to its
greater accuracy than other methods, FVM is the most popular method for predicting the planing
hull characteristics. Sukas et al., [18] conducted research using CFD, experimental, and empirical
methods. They discovered that the CFD method with overset mesh could overcome the problem of
flow at high speed around planing ships. Numerical simulation improved accuracy in resolving flows
around the hull. Research using the overset mesh method had also been carried out by Fathuddin et
al., [19] which concluded that the overset mesh method had good accuracy in predicting resistance,
trim, and heave of planing-type ships. Current advances in computer technology can create complex
hydrodynamic force models and analyses. FVM-based numerical simulation methods are widely
employed in interceptor analysis research. The RANS equation is used by the FVM approach to
describe turbulent flow in both water and air.
Based on the background above, the interceptor is a tool to control trim; thus, the ship’s
resistance can be reduced at a certain speed. However, applying an interceptor with an extended
stern form cannot be equated with a conventional ship. In this research, the interceptor would be
simulated to obtain the ship’s resistance at each speed. The approach to numerical simulations
employed FVM based on RANS. The interceptor would generate a moment and change the pressure
distribution at the ship’s stern. CFD results would be verified with the experiment of Park et al., [2]
in calm water conditions. Section 2 provided an overview of the research procedures, and Section 3
provided various hydrodynamic characteristics such as resistance, trim, heave, lift force, and wave
patterns as previously discussed.
2. Methodology
2.1 Research Object
This research used a Deep-V planing hull tested experimentally in calm water conditions with ITTC
standards, Aragon 2 is the sixth generation hull development from Kim et al., [20]. The ship, designed
without passengers, was built on a full scale by the Korea Research Institute of Ships and Ocean
Engineering (KRISO) for military purposes. Turning characteristic testing was carried out by Kim et al.,
[21]. Park et al., [2] conducted subsequent research by integrating the interceptor. The Aragon 2 CAD
model is illustrated in Figure 1, and the main dimension can be seen in Table 1.
61
CFD Letters
Volume 16, Issue 5 (2024) 59-77
Table 1
Main dimension
Dimension Full scale Model scale Unit
Scale 1 5.33
LOA 8.00 1.50 Meter
LWL 7.53 1.41 Meter
B 2.20 0.41 Meter
T 0.42 0.08 Meter
∆ 3000 19.78 Kg
This research aims to analyze the ship's drag, trim, and heave due to the interceptor on the
different forms of the extended stern. The interceptor used was the Humphree X300 interceptor with
a span(s) width of 300 mm and a stroke height (d) of 50 mm [22]. The extended stern’s bottom was
a modification to the ship stern to reduce the total resistance of the ship, trim angle, and the power
needed at high speeds; hence, it could improve ship performance [11]. The variations of the research
model are shown in Table 2. Furthermore, the visualization of the extended stern and the installation
of the interceptors are shown in Figures 2 and 3, as well as the variations in the height of the
interceptor in Figure 4.
Table 2
Model variation
Parameter Variable
Extended stern 10°, 20°, 30°
Interceptor height 20%, 60%, 100%
Froude number 0.29; 0.58; 0.87; 1.16; 1.45
62
CFD Letters
Volume 16, Issue 5 (2024) 59-77
In this study, a method called computational fluid dynamics (CFD) is used to simulate a ship
model. The simulation relies on the unsteady Reynolds-Averaged Navier-Stokes equations, a
problem-solving approach based on the conservation of mass and momentum principles, to conduct
the hydrodynamic simulation [23]. Here is the URANS equation.
𝜕𝑈𝑖
=0
𝜕𝑥𝑖 (1)
Where 𝑈𝑖 and 𝑢𝑖′ express the mean and fluctuation velocity component in the direction of the
Cartesian coordinate 𝑥𝑖 , 𝑃 is the mean pressure, 𝜌 is the density, 𝜈 is the molecular kinematic
viscosity and 𝑆𝑖𝑗 is the mean strain-rate tensor. The strain-rate tensor is defined as
1 ∂𝑈𝑖 ∂𝑈𝑗
𝑆𝑖𝑗 = ( + ) (3)
2 ∂𝑥𝑗 ∂𝑥𝑖
The last term on the right-hand side of Eq. (2) is denoted as the Reynolds stress tensor which is
given by
______
∂𝑈𝑖 ∂𝑈𝑗 1 ∂𝑈 2
𝜏𝑖𝑗 = 𝑢𝑖′ 𝑢𝑗′ =𝜇𝑡 ( + − 3 ∂𝑥 𝑘 𝜕𝑖𝑗 ) − 3 𝜌𝑘𝜕𝑖𝑗 (4)
∂𝑥𝑗 ∂𝑥𝑖 𝑘
1 𝜌𝜏𝑖𝑗
𝜇𝑡 =
2 𝑆𝑖𝑗 (5)
63
CFD Letters
Volume 16, Issue 5 (2024) 59-77
The k-ε turbulence model specifies that the turbulent eddy viscosity is calculated by
𝑘2
𝜇𝑡 = 𝑐𝜇 𝜌 (6)
𝜀
The turbulent kinetic energy 𝑘 and the rate of dissipation of the turbulent energy 𝜀 are calculated
below
∂𝜌𝑘 ∂𝜌𝑈𝑗 𝑘 ∂ 𝜇𝑡 ∂𝑘
+ = + [(𝜇 + ) ] + 𝑃𝑘 − 𝜌𝜀 (7)
∂𝑡 ∂𝑥𝑗 ∂𝑥𝑗 𝜎𝑡 ∂𝑥𝑗
∂𝜌𝜀 ∂𝜌𝑈𝑗 𝜀 ∂ 𝜇𝑡 ∂𝜀 𝜀
+ = + [(𝜇 + ) ] + (𝑐𝜀1 𝑃𝑘 − 𝑐𝜀2 𝜌𝜀) (8)
∂𝑡 ∂𝑥𝑗 ∂𝑥𝑗 𝜎𝜀 ∂𝑥𝑗 𝑘
When the energy dissipation rate 𝜀 and the kinetic energy 𝑘 are combined, the turbulent viscosity
𝜇𝑡 may be determined. A near-wall function employs a realistic two-layer turbulence technique (k-ԑ
model) to depict the velocity profile close to the wall. In this calculation, the time-step is determined
based on the ITTC guideline [24]. At the interface between the phases of water and air, the Volume
of Fluid (VOF) approach was used to simulate changes in the free surface. [25]. The time step (Δt)
used in the unsteady simulation should be small enough to complete the motion on a free surface.
The time step is the period interval for each iteration calculation. The time step used for this
simulation was a function of the ship’s speed (V) and waterline length (L) according to Eq. (9), as
recommended by the ITTC. Determining the time step in this research depended on the Courant-
Friedrichs-Lewy (CFL) number. The fluid particles' total number of points traveled throughout the
interval was denoted by the CFL number. The time step used is smaller for faster ships. This research
took the average time step value at 0.008, shown in Figure 5.
𝐿
𝛥𝑡 𝐼𝑇𝑇𝐶 = 0.005~0.01 (9)
𝑉
Prism layers were made close to the hull to capture the boundary layer's flow accurately, and six
prism layers were employed in this research. Y+ is defined as the non-dimensional distance of the
first grid node from the wall surface, normalized by the local viscous length scale. ITTC recommends
that the Y+ value is 30 < Y+ < 100 [24]. It is calculated using Eq. (10). Figure 6 shows y+ values on
ships, where the average y+ values are 50-100 at Fr 1.45. At the interface between the phases of
water and air, the Volume of Fluid (VOF) approach was used to simulate changes in the free surface
64
CFD Letters
Volume 16, Issue 5 (2024) 59-77
[25]. However, it should be noted that the Y+ value needs to be studied to capture the turbulence
phenomenon well.
(𝜌 . 𝑈 . 𝑦) (10)
𝑌+ =
𝜇
Where U the friction velocity at the wall is, y is the distance from the wall to the first grid node,
and 𝜇 is the dynamic viscosity of the fluid.
Meshing in this research used the overset mesh method. The overset mesh is a technique that
involves the use of donor-acceptor cells. The study considered two geometries: the background
served as the donor, while the overset served as the donor recipient. Overset grid is more accurate
than other discrete methods such as morphing grid and moving grid. Research by De Luca et al., in
2016 compared the overset grid and moving grid methods. The result reported that the overset grid
indicated better results than the moving grid [26]. De Marco et al., compared the overset grid method
with the morphing grid. Both revealed good results, but the overset grid was slightly better in
predicting the accuracy of the results [27]. Nevertheless, the overset mesh took a long time because
the two geometries interacted. The domain size and boundary conditions refer to the ITTC
recommendation. The details of the domain size are shown in Table 3, while the boundary conditions
are shown in Table 4. Figure 7 illustrates the size of the computational domain and boundary
conditions. Based on the length between perpendiculars (L), the domain length was measured from
−2.5 L to 1 L with the coordinates of the zero point at the stern and the vessel's draft. The width was
set to 1.5 L and the water depth was 2 L. The inflow, bottom, side, and top flow limits were described
with inlet velocity; the outlet limits were pressure outlets placed far enough away to ensure no flow
reflection occurred, and the fluid could fully expand. The body surface employed a no-slip limit
condition. The simulation was carried out on half the hull, it was modeled to save computational
time. Normal velocity and gradient variables were zero in the symmetry plane condition.
Table 3
Computational domain
Parameter Background Overset
Length 1L from FP 0.25 from FP
2.5L from AP 0.25 from AP
Height 1L from deck 0.75H from deck
2L from keel 0.75 from keel
Breadth 1.5L from symmetry 0.5B from symmetry
65
CFD Letters
Volume 16, Issue 5 (2024) 59-77
Table 4
Boundary conditions
Surface Boundary conditions
Top Velocity inlet
Side Velocity inlet
Bottom Velocity inlet
Inlet Velocity inlet
Outlet Pressure outlet
Symmetry Symmetry plane
Body Wall (no slip)
Mesh concentration was measured using isotropic or anisotropic methods based on x, y, and z
coordinates. The denser the mesh concentration would increase the computation time; hence, the
mesh density was only carried out in certain parts. Figure 8 demonstrates the concentration of the
mesh in specific parts, and the mesh density significantly affects the simulation results.
66
CFD Letters
Volume 16, Issue 5 (2024) 59-77
In essence, a model is translated into computer code, enabling the execution of a CFD simulation
that generates data utilized in engineering analysis. The process of verification and validation involves
scrutinizing both the code and simulation results to identify any errors. Five mesh quantity grids
totalling 0.40; 0.69; 1.02; 1.55; and 2.01 in million cells with the Froude number 1.45 shown in Table
5. Resistance is indicated by the non-dimensional R/Δ, trim by degree, and heave by the non-
dimensional rise of CG/draft units. Ahmed et al., compare experimental and computational results,
the results show that the resistance values at various speeds for the fine mesh show a very good and
improved agreement with the experimental data, with an error of less than 6% [28]. This indicates
that CFD model is capable of simulating the steady flow around a ship hull with an acceptable
accuracy and thus can be used as a complementary tool to laboratory model tests for ship design and
ship hydrodynamic research [29]. Figure 9 demonstrates that the data obtained was increasingly
convergent as the number of meshes increased. The results from grids 4 and 5 indicated good value
stability. However, grid 5 required a longer computation time. Therefore, the simulation in this
research used grid 4 with a total mesh of 1.55M and a large percentage of error for resistances of
8.89%, 5.97% trim, and 12.97% heave, by still considering the shorter computational time with
convergent results.
Table 5
Total number of mesh
Grid no Mesh quality Number of cells
1 Very coarse 404474
2 Coarse 696084
3 Medium 1016192
4 Fine 1550408
5 Very fine 2010230
67
CFD Letters
Volume 16, Issue 5 (2024) 59-77
The convergence of model data for resistance, trim, and heave values was evaluated by time,
which showed that the data converged after 4 seconds. The convergence of data to time is illustrated
in Figure 10.
68
CFD Letters
Volume 16, Issue 5 (2024) 59-77
Figure 11 shows the resistance, trim, and heave results between the experiment and CFD. The
pattern's alignment suggested that the numerical simulation findings and the experiment were both
reliable. However, calculation gaps/errors were different from the experimental results due to the
limitations of numerical simulations in modeling according to actual conditions. Similar cases also
occurred in the research by Song et al., with an average error for resistance of 2.65% and trim and
sinkage of 9.45% and 7.96% [30]. In this research, there was an average error for resistance of 8.92%
and trim and heave, respectively, 3.80% and 12.3% for all speed variations.
A ship moving on the water will produce a phenomenon of water flow from the front to the stern.
The waves generated as a result of the flow of water hitting the ship’s front will pass through the
middle of the ship and then go to the ship’s rear, which creates a phenomenon of water flow at the
ship’s rear due to changes in the flow velocity generated from the ship’s front. Therefore, the stern
hull form's choice also significantly affects the water flow behind the ship. Besides, the determination
of the form of the stern hull will significantly affect the magnitude of the resistance value of the ship.
Figure 12 shows the reduced resistance, trim, and heave values with different extended stern forms
with angles of 10°, 20°, and 30°. The diagram illustrates that the greater the extended stern angle,
the greater the effect on the ship’s resistance, trim, and heave. The effective reduction occurred at
the extended stern angle of 30° Froude number 0.58 with a reduction percentage of 26% resistance,
69
CFD Letters
Volume 16, Issue 5 (2024) 59-77
45% trim, and 20% heave. Extended stern with angles more than 30° is not recommended, as shown
on the graph will only add drag at high speeds.
The interceptor is a small vertical plate that is usually found close to the stern and runs
perpendicular to the hull. The working principle of the interceptor is shown in Figure 13. The main
role of the interceptor is to provide a balance between the moment caused by the interceptor and
the trim. The increased pressure distribution near the interceptor increased the coefficient of friction
and lift. The resulting lift force is greater than the frictional force produced by the interceptor if the
interceptor height is properly chosen. The optimal height of the interceptor will increase its efficiency
when properly selected [16]. The interceptor's efficiency would be weak, optimal, or equal if the
moment it produces is less than, greater than, or equal to the trim moment. The ship's trim will not
be properly controlled in a weak interceptor while unfit. The ship's resistance will then increase,
causing a negative trim and a safety issue [31]. Therefore, determining the thickness of the ship's
transom's boundary layer was the initial stage of designing the interceptor. Figure 14 shows the value
of the ship’s boundary layer at each speed variation. The boundary layer thickness at transom
estimation is based on the water length and speed. An option is CFD, as done in the previous part.
However, using CFD to estimate the boundary layer thickness for boats is difficult and time-
consuming. Another option is to use an analytical solution as suggested in Ref. [32]. The bottom of
the planing boats is almost flat. Thus, as an initial approximation, the boundary layer near the
transom (for aft interceptor location) can be considered in the same way as the one over the flat
plate having the same length as the hull. The thickness of the boundary layer in turbulent flow can
70
CFD Letters
Volume 16, Issue 5 (2024) 59-77
be calculated [32] via Eq. (11). Where, LWL is the water line length and ReLWL is the corresponding
Reynolds number.
Fig. 13. Interceptor working principle Fig. 14. Boundary layer thickness
The interceptor caused considerable changes in pressure around the bottom of the ship, especially
on the transoms. Pressure variations affected resistance, draft height, and lift force. Based on recent
research, the height of the interceptor should not be higher than 60% (d/h ≤ 0.6) of the transom
boundary layer. For optimal efficiency, when the height of the interceptor is 60% of the boundary
layer, the span length of the interceptor must be seven times the height of the interceptor [31].
Considering the variation angles of the interceptor installations, as illustrated in Figure 15, the height
of the interceptor (d) could be calculated using Eq. (12).
𝑑 = 𝑑𝑖 cos 𝜃 (12)
Where (di) was the interceptor’s stroke height, and (θ) was the interceptor’s inclination angle.
The height comparison of the interceptor and boundary layer (d/h) was presented in Figure 16 for
each speed variation on resistance and trim. As previously mentioned, when the d/h ratio was much
smaller than 0.6, the efficiency of the interceptor became weak to control trim and reduce resistance.
The interceptor would lose effectiveness when d/h was more significant than 0.6. The interceptor
would generate a powerful moment which could create a negative trim. Even worse, it could capsize
the ship. To overcome this problem, the height of the interceptor must be reduced. The height of the
drop would reduce the lift force and combine the interceptor with a trim tab [33].
In this research, the d/h value in the range of 0.31 - 0.43 was still satisfactory concerning previous
research, which was not more than 0.6. The ratio of the height of the interceptor to the boundary
layer (d/h) suggested for ships with extended stern 10° was 0.38, while ships with extended stern 20°
and 30° were 0.37 and 0.34 respectively, referring to the improvement in the best performance value
of the ship on Froude number 0.87. If the d/h value were too large or exceeded 0.6, it would result
in excessive trim and increased resistance, which could cause safety problems.
The effectiveness of the interceptor could be increased by paying attention to the height of the
interceptor at each speed. The interceptor, which could be controlled in height when the ship was
operating, could increase effectiveness when adjusting to the ship's speed.
71
CFD Letters
Volume 16, Issue 5 (2024) 59-77
The research results are shown in Figure 17 with variations in the extended stern angles of 10°,
20°, and 30° at interceptor height conditions of 100%, 60%, and 20%. Generally, high-speed planing
vessels are divided into three phases: the displacement phase, the semi-displacement/ semi-planing
phase, and the planing phase. Figure 17 (a) demonstrates the improvement in resistance values in
the displacement, semi-displacement, and planing phases, namely at the Froude number 0.29 to
1.45. However, the best resistance improvement occurred in the semi-displacement phase at Froude
number 0.87, with a reduction percentage of 24% extended at 10°, 29% extended at 20°, and 33%
extended at 30° with 100% interceptor. Likewise, with the conditions of 60% and 20% interceptor, it
had the same pattern as the 100% interceptor; i.e., there was an improvement in the resistance value
in the semi-displacement phase. The best resistance value improvement occurred at the Froude
number 0.87, with a percentage reduction in resistance values in the 13% to 29% range. However, at
Froude number > 1.45, the interceptor was not needed because it would only increase the ship’s
resistance.
The trim analysis results can be seen in Figure 17 (b). The trim decrease had the same pattern in
each variation. The trim value was improved on the Froude number 0.87 with a reduction percentage
of 54% extended 10°, 60% extended 20°, and 66% extended 30° at the height of 100% interceptor.
As for the height of 60% and 20% improvement, the best trim value was the Froude number 0.87,
72
CFD Letters
Volume 16, Issue 5 (2024) 59-77
with a reduction percentage of 24% to 49%. Trim visualization can be seen in Figure 18 with a
significant reduction in trim angle from the bare hull and extended stern variations and the addition
of an interceptor at Froude number 0.87. The interceptor was not recommended at high speeds
because it would cause excessive trim.
Figure 17 (c) illustrates the heave graph; there were improvements in the heave value with a
consistent pattern for each variation and the heave value with a range of 27% to 47%. The best
reduction in heave value was 47% at the Froude number 0.87 extended 30° with a height of 100%
interceptor. It is similar to the interceptor height of 20% and 60%, with a reduction in the heave value
of 38% and 45% at the Froude number 0.87.
The speed and height of the interceptor were essential factors in the effectiveness of the
interceptor. Figure 17 (d) shows the lift force at the extended stern angle and interceptor height
variation. This research discovered that the lift force due to the interceptor got more significant as
the ship's speed increased, but the change was not too significant. The highest increase occurred at
the Froude number 1.45 with a percentage of 2.73% at the extended stern 30°, and the interceptor
height was 100% with an average lift force of 1.72% at each speed variation.
73
CFD Letters
Volume 16, Issue 5 (2024) 59-77
The pressure distribution at the bottom of the ship in the bare hull and extended stern conditions
using the interceptor is illustrated in Figure 19. The pressure in the transom for the hull without the
interceptor was observed to be less than employing the interceptor. The pressure distribution
decreased from the high-pressure area in the bow to the stern where the interceptor had been
installed. It created an interceptor moment, which would affect trim. The moment triggered was due
to the presence of a flowing fluid that was restrained by the interceptor and caused pressure. This
pressure creates an upward lift and reduces trim. The optimal interceptor was obtained when the
interceptor moment was equal to the trim moment.
Fig. 19. Pressure distribution of the bare hull and the interceptor at the bottom
Figure 20 shows the wave pattern on the free surface behind the ship's stern at Froude number
0.87. It revealed that the pressure value near the ship’s stern increased with the presence of an
interceptor, which caused the wave pool formed behind the ship’s stern to divide into two. It caused
a reduction in the ship's resistance value and trim value. A similar case was observed for other fraud
numbers where an interceptor was useful.
74
CFD Letters
Volume 16, Issue 5 (2024) 59-77
4. Conclusions
In this research, the effect of the extended stern form with different angles and the variation of
the height of the interceptor at various speeds was investigated. The conclusions obtained are as
follows:
i. The extended stern form without an interceptor could affect the ship's resistance, trim, and
heave values. This research revealed that the more significant the stern extend angle, the
greater the reduction in resistance in the semi-displacement phase. The effective reduction
occurred at the Froude number 0.58 to 0.87 with a percentage reduction in the resistance
value of 26% extended stern 30°. However, the resistance increased at high speeds as the
extended stern angle increased.
ii. Adding an interceptor on the extended stern significantly reduced resistance, trim, and
heave values. The effective reduction occurred in the semi-displacement phase, namely at
the Froude number 0.58 to 0.87, with a percentage reduction in the resistance value of 33%,
and trim and heave were 66% and 47% for the extended stern 30° with an interceptor height
of 100%.
iii. Using an interceptor at high speed was not recommended, or the Froude number was more
than 1.45 because it would increase the ship’s resistance and cause excessive trim that could
endanger safety.
iv. The recommended ratio of boundary layer and interceptor height (d/h) on ships with
extended stern 10°, 20°, and 30° were 0.38, 0.37, and 0.34 at Froude number 0.87.
Acknowledgement
The authors gratefully acknowledge financial support from Universitas Diponegoro for this work
under the project scheme of Riset Publikasi Internasional (RPI) No. 259/UN7.A/HK/VIII/2023.
75
CFD Letters
Volume 16, Issue 5 (2024) 59-77
References
[1] Karimi, Mohammad Hossein, Mohammad Saeed Seif, and Majid Abbaspoor. "A study on vertical motions of high-
speed planing boats with automatically controlled stern interceptors in calm water and head waves." Ships and
Offshore Structures 10, no. 3 (2015): 335-348. https://doi.org/10.1080/17445302.2013.867647
[2] Park, Jong-Yong, Hujae Choi, Jooho Lee, Hwiyong Choi, Joohyun Woo, Seonhong Kim, Dong Jin Kim, Sun Young Kim,
and Nakwan Kim. "An experimental study on vertical motion control of a high-speed planing vessel using a
controllable interceptor in waves." Ocean Engineering 173 (2019): 841-850.
https://doi.org/10.1016/j.oceaneng.2019.01.019
[3] Seok, Woochan, Sae Yong Park, and Shin Hyung Rhee. "An experimental study on the stern bottom pressure
distribution of a high-speed planing vessel with and without interceptors." International Journal of Naval
Architecture and Ocean Engineering 12 (2020): 691-698. https://doi.org/10.1016/j.ijnaoe.2020.08.003
[4] Tsai, J. F., and J. L. Hwang. "Study on the compound effects of interceptor with stern flap for two fast monohulls."
In Oceans' 04 MTS/IEEE Techno-Ocean'04 (IEEE Cat. No. 04CH37600), vol. 2, pp. 1023-1028. IEEE, 2004.
[5] Fitriadhy, Ahmad, Intan Nur Nabila, Christina Bangi Grosnin, Faisal Mahmuddin, and Suandar Baso. "Computational
Investigation into Prediction of Lift Force and Resistance of a Hydrofoil Ship." CFD Letters 14, no. 4 (2022): 51-66.
https://doi.org/10.37934/cfdl.14.4.5166
[6] Day, Alexander H., and Christopher Cooper. "An experimental study of interceptors for drag reduction on high-
performance sailing yachts." Ocean Engineering 38, no. 8-9 (2011): 983-994.
https://doi.org/10.1016/j.oceaneng.2011.03.006
[7] Ghassemi, H., M. Mansouri, and S. Zaferanlouei. "Interceptor hydrodynamic analysis for handling trim control
problems in the high-speed crafts." Proceedings of the Institution of Mechanical Engineers, Part C: Journal of
Mechanical Engineering Science 225, no. 11 (2011): 2597-2618. https://doi.org/10.1177/0954406211406650
[8] Suneela, J., P. Krishnankutty, and V. Anantha Subramanian. "Numerical investigation on the hydrodynamic
performance of high-speed planing hull with transom interceptor." Ships and Offshore Structures 15, no. sup1
(2020): S134-S142. https://doi.org/10.1080/17445302.2020.1738134
[9] Sahin, Omer Sinan, Emre Kahramanoglu, and Ferdi Cakici. "Numerical evaluation on the effects of interceptor
layout and blade heights for a prismatic planing hull." Applied Ocean Research 127 (2022): 103302.
https://doi.org/10.1016/j.apor.2022.103302
[10] Karimi, Mohammad Hosein, Mohammad Saeed Seif, and Madjid Abbaspoor. "An experimental study of
interceptor’s effectiveness on hydrodynamic performance of high-speed planing crafts." Polish maritime
research 20, no. 2 (2013): 21-29. https://doi.org/10.2478/pomr-2013-0013
[11] Kim, Dong Jin, Sun Young Kim, Young Jun You, Key Pyo Rhee, Seong Hwan Kim, and Yeon Gyu Kim. "Design of high-
speed planing hulls for the improvement of resistance and seakeeping performance." International Journal of
Naval Architecture and Ocean Engineering 5, no. 1 (2013): 161-177. https://doi.org/10.2478/IJNAOE-2013-0124
[12] Mansoori, Mehran, and Antonio Carlos Fernandes. "The interceptor hydrodynamic analysis for controlling the
porpoising instability in high speed crafts." Applied Ocean Research 57 (2016): 40-51.
https://doi.org/10.1016/j.apor.2016.02.006
[13] Suneela, Jangam, and Prasanta Sahoo. "Numerical investigation of interceptor effect on seakeeping behaviour of
planing hull advancing in regular head waves." Brodogradnja: Teorija i praksa brodogradnje i pomorske tehnike 72,
no. 2 (2021): 73-92. https://doi.org/10.21278/brod72205
[14] Samuel, S., Ocid Mursid, Serliana Yulianti, Kiryanto Kiryanto, and Muhammad Iqbal. "Evaluation of interceptor
design to reduce drag on planing hull." Brodogradnja: Teorija i praksa brodogradnje i pomorske tehnike 73, no. 3
(2022): 93-110. https://doi.org/10.21278/brod73306
[15] Samuel, S., Serliana Yulianti, Parlindungan Manik, and Abubakar Fathuddiin. "A study of interceptor performance
for deep-v planing hull." In IOP Conference Series: Earth and Environmental Science, vol. 1081, no. 1, p. 012004.
IOP Publishing, 2022. https://doi.org/10.1088/1755-1315/1081/1/012004
[16] Mansoori, M., and A. C. Fernandes. "Hydrodynamics of the interceptor on a 2-D flat plate by CFD and
experiments." Journal of Hydrodynamics, Ser. B 27, no. 6 (2015): 919-933. https://doi.org/10.1016/S1001-
6058(15)60555-8
[17] Yousefi, Reza, Rouzbeh Shafaghat, and Mostafa Shakeri. "Hydrodynamic analysis techniques for high-speed planing
hulls." Applied ocean research 42 (2013): 105-113. https://doi.org/10.1016/j.apor.2013.05.004
[18] Sukas, Omer Faruk, Omer Kemal Kinaci, Ferdi Cakici, and Metin Kemal Gokce. "Hydrodynamic assessment of
planing hulls using overset grids." Applied Ocean Research 65 (2017): 35-46.
https://doi.org/10.1016/j.apor.2017.03.015
76
CFD Letters
Volume 16, Issue 5 (2024) 59-77
[19] Fathuddiin, Abubakar, Samuel Samuel, Kiryanto Kiryanto, and Aulia Widyandari. "Prediksi hambatan kapal dengan
menggunakan metode overset mesh pada kapal planing hull." Rekayasa Hijau: Jurnal Teknologi Ramah
Lingkungan 4, no. 1 (2020): 24-34. https://doi.org/10.26760/jrh.v4i1.24-34
[20] D. J. Kim, and S. Y. Kim. "Comparative study on manoeuvring performance of model and full-scale waterjet
propelled planing boats." Society of Naval Architects and Marine Engineers (SNAME), 2017.
[21] Kim, Dong Jin, and Sun Young Kim. "Turning characteristics of waterjet propelled planing boat at semi-planing
speeds." Ocean Engineering 143 (2017): 24-33. https://doi.org/10.1016/j.oceaneng.2017.07.034
[22] Humphree, “Interceptors Series,” 2022. [Online]. Available: https://humphree.com/. [Accessed: 02-Aug-2021].
[23] Kinaci, Omer K., Omer F. Sukas, and Sakir Bal. "Prediction of wave resistance by a Reynolds-averaged Navier–Stokes
equation–based computational fluid dynamics approach." Proceedings of the Institution of Mechanical Engineers,
Part M: Journal of Engineering for the Maritime Environment 230, no. 3 (2016): 531-
548.https://doi.org/10.1177/1475090215599
[24] ITTC Specialist Committee. "Recommended procedures and guidelines-Practical Guidelines for Ship CFD
Applications." In 26th International Towing Tank Conference, Rio de Janeiro, Brasil. 2011.
[25] Suneela, J., P. Krishnankutty, and V. Anantha Subramanian. "Hydrodynamic performance of planing craft with
interceptor-flap hybrid combination." Journal of Ocean Engineering and Marine Energy 7 (2021): 421-438.
https://doi.org/10.1007/s40722-021-00211-0
[26] De Luca, Fabio, Simone Mancini, Salvatore Miranda, and Claudio Pensa. "An extended verification and validation
study of CFD simulations for planing hulls." Journal of Ship Research 60, no. 02 (2016): 101-118.
https://doi.org/10.5957/jsr.2016.60.2.101
[27] De Marco, Agostino, Simone Mancini, Salvatore Miranda, Raffaele Scognamiglio, and Luigi Vitiello. "Experimental
and numerical hydrodynamic analysis of a stepped planing hull." Applied Ocean Research 64 (2017): 135-154.
https://doi.org/10.1016/j.apor.2017.02.004
[28] Elhadad, Alaaeldeen M., and Abo El-Ela. "Experimental and Cfd Resistance Validation of Naval Combatant Dtmb
5415 Model." Experimental and Cfd Resistance Validation of Naval Combatant Dtmb 5415.
https://doi.org/10.37934/arfmts.107.2.84102
[29] Ahmed, Alaaeldeen Mohamed Elhadad. "Comparative investigation of resistance prediction for surface combatant
ship model using CFD modeling." Journal of Advanced Research in Fluid Mechanics and Thermal Sciences 107, no.
2 (2023): 225-235. https://doi.org/10.37934/arfmts.107.2.225235
[30] Song, Ke-wei, Chun-yu Guo, Jie Gong, Ping Li, and Lian-zhou Wang. "Influence of interceptors, stern flaps, and their
combinations on the hydrodynamic performance of a deep-vee ship." Ocean Engineering 170 (2018): 306-320.
https://doi.org/10.1016/j.oceaneng.2018.10.048
[31] Mansoori, M., A. C. Fernandes, and H. Ghassemi. "Interceptor design for optimum trim control and minimum
resistance of planing boats." Applied Ocean Research 69 (2017): 100-115.
https://doi.org/10.1016/j.apor.2017.10.006
[32] Schlichting, Hermann, and Joseph Kestin. Boundary layer theory. Vol. 121. New York: McGraw-Hill, 1961.
[33] Mansoori, M., and A. C. Fernandes. "Interceptor and trim tab combination to prevent interceptor's unfit
effects." Ocean engineering 134 (2017): 140-156. https://doi.org/10.1016/j.oceaneng.2017.02.024
77