You are on page 1of 19

International Journal of Naval Architecture and Ocean Engineering 16 (2024) 100578

Contents lists available at ScienceDirect

International Journal of Naval Architecture


and Ocean Engineering
journal homepage: www.journals.elsevier.com/international-journal-of-naval-
architecture-and-ocean-engineering/

A numerical study on the optimization of the slit shape of a jet


injection propeller
Jun-Hee Lee a, Ju-Han Lee a, Myeong-Min Kim a, Dohan Oh b, Kwang-Jun Paik a, *
a
Department of Naval Architecture and Ocean Engineering, Inha University, Incheon, Republic of Korea
b
Department of Naval Architecture and Ocean Engineering, Hongik University, Sejong, Republic of Korea

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

Keywords: The shape of the slit that injects the jet from the surface of the propeller was optimized through numerical
Coanda effect computations. The high-pressure drop on the propeller surface caused by jet injection could be improved by
Computational Fluid Dynamics (CFD) modifying the hydrofoil geometry of the propeller. As the cover length increased and the slit was located at the
Design optimization
center of the propeller, the volume of jets leading to the trailing edge increased, and the propulsion performance
Propeller performance
Propeller efficiency
was improved. As the height of the slit increased, the thrust increased due to the Coanda effect, and the torque
decreased because of the thrust of the jet. The jet injection pattern differed according to the area of the slit and
tunnel, which caused a difference in the propeller performance. The jet pattern changed according to the area of
the slit and tunnel, leading to a change in propeller performance. It was effective in improving the efficiency by
injecting from as wide an area as possible, and the efficiency was improved by approximately 2 % considering the
pump efficiency through optimization of the slit shape.

1. Introduction (Timoshevskiy et al., 2018) and hole shape (Kadivar et al., 2020)
through experiments. It was confirmed that this effect increased lift on
In 2013, International Maritime Organization (IMO) institutional­ the rudder (Seo et al., 2010, 2016). Lu et al. (2018) injected a jet at the
ized Energy Efficiency Design Index (EEDI) and tightened regulations to suction side of NACA0066 to derive cavitation suppression and optimal
reduce greenhouse gas (GHG) emissions from ships; vessels built after mass flow coefficient according to several cavitation numbers.
2025 will need to reduce the number of greenhouse gases by 30% In many studies, jets were injected to control cavitation. Wang et al.
compared to the standard value (Joung et al., 2020). Various studies (2020) carried out a cavitation generation model experiment by jet in­
have been conducted to improve the efficiency of ships and reduce the jection using a hydrofoil of NACA66 (MOD) and clarified that jet in­
amount of GHG emitted from ships. Mizzi et al. (2017) improved the jection is effective in reducing cavitation. Pant and Frankel (2021)
propulsion efficiency by installing Propeller Boss Cap Fins (PBCF) conducted a numerical analysis of the Large Eddy Simulation (LES)
through computations and the development of optimization technology. turbulence model. They showed that jet injection and suction effectively
Go et al. (2017) designed a duct for the propeller based on NACA0015, control cavitation generated on the hydrofoil suction side. The jet in­
which improved the efficiency of the ship. jection is effective in cavitation control and in increasing hydrofoil lift
In this study, an attempt was made to improve the propeller pro­ (Maltsev et al., 2020; Oh and Kim, 2010; Oh and Noh, 2015). Further­
pulsion efficiency by injecting a jet from the surface of a marine pro­ more, it was used in research on ventilated supercavitation, which
peller and applying the Coanda effect. The Coanda effect refers to a applied the injection method to underwater bodies (Paik et al., 2018; Lee
phenomenon in which a fluid with a high velocity in contact with a et al., 2019). Xu et al. (2021) constructed a simulated ventilated
curved surface flows along the curvature of the curved surface (Burnazzi supercavitating flow and observed the geometry and characteristics of
and Radespiel, 2014, 2015). Timoshevskiy et al. (2016) designed an the supercavity based on the flow rate. Pham et al. (2022) compared the
experimental hydrofoil device capable of jet injection. They compared geometry and pressure of a supercavity injected with hot gas with
the cavitation control performance according to the slit shape experiment. They reported that the structure inside the hot-gas

Peer review under responsibility of The Society of Naval Architects of Korea.


* Corresponding author.
E-mail address: kwangjun.paik@inha.ac.kr (K.-J. Paik).

https://doi.org/10.1016/j.ijnaoe.2023.100578
Received 21 September 2023; Received in revised form 29 November 2023; Accepted 14 December 2023
Available online 23 December 2023
2092-6782/© 2023 Published by Society of Naval Architects of Korea. This is an open access article under the CC BY-NC-ND license
(http://creativecommons.org/licenses/by-nc-nd/4.0/).
J.-H. Lee et al. International Journal of Naval Architecture and Ocean Engineering 16 (2024) 100578

All computations were compared using an advance ratio (JA ) of 0.4,


and the advance ratio can be defined using Eq. (1) as follows:
VA
JA = (1)
nD

VA was fixed to 0.8 m/s as the advance speed of the propeller, and n was
fixed at 8 RPS (revolution per second) at the rotation rate. In addition,
the performance of the propeller was compared with KT and 10KQ using
Eqs. (2) and (3).
T
KT = (2)
ρn2 D4

10Q
10KQ = (3)
ρn2 D5

T means the thrust; Q means the torque of the propeller; ρ is the density
Fig. 1. Comparison of the shapes of Coanda and Normal propeller. of water.
Fig. 2 presents the names of each part of the Coanda propeller. The
passage that moves water from the shaft to the injection port is called a
tunnel, and the water is injected through the slit. The part that connects
the slit and blade at the end of the round tunnel is called the tip, and the
area where the jet is injected is called the jet side. Subsequently, the
changes in KT and 10KQ depending on the injection volume, were
divided into the total, blade, and inner structure, and compared. The
blade included a suction side, pressure side, and jet side, and the inner
structure included by tunnel and round tunnel.

2.2. Numerical modeling

All computations in this study were performed using commercial


CFD software, STAR-CCM+ 15.06 ver. For computations, implicitly
unsteady, incompressible flow as the physical model was selected, and
the continuity equation and Reynolds averaged Navier-Stokes equation
(RANS) were used as the governing equations in Eqs. (4) and (5).
∫ ∫
d
ρdΩ + ρui ni dS = 0 (4)
dt Ω S
Fig. 2. Definition for each part of the Coanda propeller.
∫ ∫ ∫ ∫
d ( )
ρdΩ + ρui ni dS = τij nj − pni dS + ρbi dΩ (5)
ventilated supercavity can be classified into three regions, and the dt Ω S S Ω

pressure inside the supercavity remains unchanged in the region close to


the cavitator. ui and bi means the velocity and volume tensors, respectively. p is the
Based on previous studies, Lee et al. (2021) designed a jet injected pressure and τij is the effective stress due to viscosity and turbulence, as
propeller through Eom et al. (2020), conducted a model test and shown in Eq. (6).
computation, and confirmed the possibility of improving the efficiency [(
∂ui ∂uj
)
2 ∂uk
]
of the jet-injected propeller. On the other hand, it is necessary to τij = μe + − δij (6)
∂xj ∂xi 3 ∂xk
improve cavitation performance deterioration from high-pressure drops
on the jet injection surface caused by jet injection. In this study, an The SST (Shear Stress Transport) k − ω turbulence model (Menter,
improved hydrofoil section (Lee et al., 2021) was used to reduce the 1994) combines the stability advantages of the k − ω model and inde­
occurrence of cavitation. In addition, the performance was compared by pendence from the free-stream boundary of the k − ε model, leading to
changing the slit shape, and the possibility of increasing propeller effi­ significant improvements in predicting adverse pressure gradient flows.
ciency through optimization was evaluated. This model has a high accuracy in simulating the predicting of separa­
tion and the ventilated flow (Zhou et al., 2011; Xu et al., 2021).
2. Numerical method In theory, the turbulence model assumes completely turbulent fluid
flow, but in the model-scale propeller, both turbulent and laminar flow
2.1. Propeller modeling regions coexist on the propeller surface. Kuiper (1981) confirmed
laminar flow in a model propeller. The research was conducted that
In this study, the performance was compared using a propeller that utilized the transition model provided by the SST k − ω turbulence
injects a jet with a slit on the surface of the propeller and a general screw model to perform calculations involving transition regions. A perfor­
propeller without a slit. The propeller diameter (D) was 250 mm, the mance analysis of a model-scale propeller was conducted using both a
pitch (P) of Coanda propeller at 0.7 R was P/D = 0.554, and the pitch of transition model and a turbulence model, and the transition model re­
a normal propeller was P/D = 0.514. The chord length (c) of both sults shows good agreement with the experiment (Bhattacharyya et al.,
propellers was 50 mm, and the detailed specifications of the propellers 2016; Baltazar et al., 2018, 2021; Kim et al., 2021; Wang and Walters,
are the same as Lee et al. (2021). Fig. 1 compares the shapes of both 2012). I n the study, the turbulence model used for the computations
propellers. used the SST k − ω model, and the gamma transition model (Menter
et al., 2015) was set considering that the Reynolds number in the

2
J.-H. Lee et al. International Journal of Naval Architecture and Ocean Engineering 16 (2024) 100578

propeller and perimeter grid was added to measure the ambient flow
injected from the propeller slits more closely than in previous studies
(Lee et al., 2021). A prism layer was used to generate prismatic cells, in
which the distance from the first grid cell to the surface wall was
determined so that the dimensionless wall distance (y+ ) was set to be 1
or less.
This study introduced the equations for thrust per unit area (Thrust
PUA) and torque per unit area (Torque PUA) to analyze the regions on
the propeller where thrust and torque are concentrated. The thrust PUA
was determined by combining the x-axis component of the pressure
acting on the tangential vector of the propeller surface with the wall
shear stress. Torque was determined by the moment generated in the
grid, utilizing the sum of rotational pressure and wall shear stress.
Thrust PUA and torque PUA are shown in Fig. 5 and can be defined using
Eqs. (7) and (8) as follows:
Fig. 3. Computation domain and boundary conditions. →
TPUA = P • →
nx + Sx (7)
( ( →))
QPUA = →r × P•→
n + S (8)
x

where P means the pressure, S and → n are the wall shear stress and the
tangential vector of the propeller surface, respectively. r is the moment
arm of the torque from the propeller center.
The thrust and torque can be defined using Eqs. (9) and (10), which
means the summation of the pressure and wall shear stress generated in
the unit area of the propeller surface separately in the thrust direction
and torque direction of the propeller.

T= TPUA Δs (9)

Q= QPUA Δs (10)

Fig. 4. Grid system for numerical simulations. where T means the total thrust of the propeller, Q means the total torque,
Δs means the area of each surface mesh on the propeller.
computation conditions corresponds to the transition regions (Re =
2.509 × 105 ). 2.3. Validation and grid dependency test
Fig. 3 shows the domain and boundary conditions used in the
calculation, giving mass flow inlet boundary conditions to the end of the The reliability of numerical computations was evaluated by
shaft for jet injection so that water entering the shaft can be injected comparing the results of model experiments and performing a grid de­
from the propeller surface. The domain size was set to 10D on both sides pendency test. The results were compared with the model experiment
from the center of the propeller and the distance between the inlet and results (Lee et al., 2021), as shown in Fig. 6. No significant change was
outlet to 40Dso that the wake generated by the propeller rotation does observed because the grid in the jet injection region around the propeller
not affect the boundary conditions. was reinforced, and it can be confirmed that in the case of propeller
Fig. 4 shows the grid system in the calculation domain. The simu­ efficiency, the difference from the experiment was relatively reduced at
lation was the same, except that a finer volume mesh around the a high advance ratio.

Fig. 5. Schematic diagram of thrust PUA and torque PUA.

3
J.-H. Lee et al. International Journal of Naval Architecture and Ocean Engineering 16 (2024) 100578

Fig. 6. Comparison of the propeller performance between the experiment and computation result.

A grid dependency test was conducted to assess the convergence of


Table 1
the results based on the number of grids. The Grid Convergence Index
Grid numbers for grid dependency test.
(GCI) can be found in the study of Celik et al. (2008). To distinguish the
Grid type Grid number discrete error from other errors, a recommended value for the grid
Coarse 1,012,145 refinement ratio is 1.3 or more, as suggested by Celik et al. (2008). In
Medium 2,293,012 this study, as it was composed of an unstructured gird, the grid numbers
Fine 4,610,918 are changed by adjusting the base size, and the grid refinement ratio was
√̅̅̅
set to 2. The test was performed with three grid numbers, and Table 1
lists the number of grids for each condition. The conditions corre­
Table 2 sponded to JA = 0.4 among the conditions validated with the experi­
Numerical uncertainty analyses from grid convergence at J = 0.4. ment, and were performed under the condition of jet injection, and the
KT 10KQ convergence of KT and 10KQ was evaluated, as listed in Table 2. After
Coarse 0.0521 0.0824 that, the grid was used for all computations because high convergence
Medium 0.0501 0.0753 was confirmed in the fine grid.
Fine 0.0497 0.0738
GCIfine 0.39 % 0.98 %
3. Results

3.1. Performance variation by hydrofoil geometry

Three hydrofoils were considered to improve the high-pressure drop


on the suction surface shown in the previous research, and the geome­
tries are shown in Fig. 7. The shape of the slit and the trailing edge was
corrected where the Coanda effect generated the additional thrust. The
geometry was corrected so that the curvature decreased from cases 1 to
3. The shape of the tunnel changed compared to Lee et al. (2021)
because of the change in hydrofoil geometry, but the area was set to be
the same. The height of the slit was 1.0 mm, as in the previous study. The
geometry of all other parts except the shape of the tunnel and the jet side
were the same, and the change in performance due to the hydrofoil
geometry was compared.
By adjusting the jet injection volume of the Coanda propeller, the
propeller thrust was matched with the normal propeller within 1%, and
this jet injection volume is made dimensionless like Eq. (11).
ṁVjet
Cj = 1 (11)
2
ρSVR2

where ṁ is the mass flow rate; Vjet is the velocity of the jet injected from
the slit; ρ is the density; S is the expanded area of the propeller; VR is the
relative velocity of the propeller at r/R = 0.7.
The head of the pump was also considered to determine the pump
power needed to inject the jet from the propeller. The head of the pump
is given considering only the velocity head of the outlet, as in Lee et al.
(2021), which seems to be in Eq. (12). In addition, Eq. (13) expresses the
required power of the pump derived using the head considered in this
Fig. 7. Hydrofoil geometry for each case.
way:

4
J.-H. Lee et al. International Journal of Naval Architecture and Ocean Engineering 16 (2024) 100578

Table 3
Computation conditions and propeller performance comparison for each case.
Type KT 10KQ Cj ṁ(kg /s) Vjet (m /s) PPump (W)

Normal prop. 0.060 0.064 – - - –


Lee et al. 0.005 0.052 0.68 7.01 16.74
(2021)
Case 1 − 0.007 0.054 0.69 7.09 17.31
Case 2 − 0.026 0.064 0.75 7.72 22.33
Case 3 − 0.087 0.103 0.96 9.83 46.16

Fig. 9. Hydrofoil geometry for each cover length.

Table 4
Propeller performance comparison for each cover length.
Type KT 10KQ Cj ṁ (kg /s) Vjet (m /s) PPump (W)

Normal prop. 0.060 0.064 – - - –


lCO. = 0.00 c − 0.026 0.064 0.75 7.77 22.66
lCO. = 0.05 c − 0.026 0.062 0.74 7.66 21.67
lCO. = 0.10 c − 0.026 0.062 0.74 7.67 21.77

Fig. 8. -CP distribution according to each case.

2
Vjet
H= (12)
2g

1+α
PPump = ṁgH (13)
ηPump

where α is the margin rate, ηPump is the pump efficiency, and the margin
ratio was assumed to be zero in this study.
Table 3 and Fig. 8 compare the performance of three thrust-matched
propellers with Cj , required power, and pressure distribution on the
propeller surface. The pressure distribution on the propeller surface was
dimensionless with − CP , which was defined by Eq. (14).
p
− CP = − (14)
1
2
ρ(nD)2
Fig. 10. -CP distribution according to each cover length.
where p is the pressure on the surface of the propeller.
As the curvature of the jet side decreases, the induced downwash by 3.2. Performance variation by cover length(lCO. )
the jet decreases, and the thrust increased by the Coanda effect for the
same jet volume decreases. Therefore, Cj of Case 3 appears highest. Next, a cover was added to the slit, and the propeller performance
Therefore, the required pump power is also required to be the highest. In was compared. The propeller cover is a part that continues to the hy­
addition, as the jet injection volume increased, a moment was generated drofoil at the upper end of the injection port, and propellers with lengths
in the propeller rotation direction because of the propulsive force of the of 0.05 c and 0.1 c were designed. The height of the slit was 1.0 mm, as in
jet injected from the slit. The propeller torque tends to decrease as the the previous condition, and all shapes were the same except for the cover
curvature of the jet side decreases because the propulsive force of the jet length. Fig. 9 shows the shape of the hydrofoil according to each cover
increases as the injection volume increases. length.
The pressure distribution on the propeller surface also differed Table 4 lists the propeller performance when the propellers having
depending on the shape of the jet side. Fig. 8(c), (d), and (e) confirmed each cover length match the thrust of the normal propeller. When the
that the high-pressure drop that occurred in the previous study was thrusts are matched, the torque does not make a large difference
reduced significantly. Furthermore, the area where the intensive pres­ depending on the length of the cover. In the case of the injection volume,
sure drop occurs is reduced as the curvature of the jet side is reduced. A however, a slight difference appeared in the process of matching the
high-pressure drop appears locally due to the high injection volume thrusts, so that the maximum difference was approximately 3.0 % for Cj
(Fig. 8(e)). and 4.6 % for the required power of the pump.
Although the pressure drop is improved in all jet side shapes Fig. 10 compares the propeller surface pressure distribution for each
compared to the previous study, the hydrofoil thickness on the trailing cover length. The change in the pressure distribution appears only near
edge side is reduced to reduce the curvature of the hydrofoil, as shown in the jet side. As the cover length increases, the area where the pressure
Fig. 7, and the thickness reduction increases as the curvature is reduced. drop appears on the jet side tends to decrease, but a concentrated
Therefore, the jet side shape of the propeller was fixed with Case 2, a pressure drop occurs near the trailing edge.
hydrofoil geometry in which the intensive pressure drop was greatly The changes in these pressure distributions were analyzed. Fig. 11
improved, and the thickness of the propeller on the trailing edge side compares the ambient flow at r/R = 0.84, where the difference in
was reinforced. pressure changes near the trailing edge appears largest. The tangential

5
J.-H. Lee et al. International Journal of Naval Architecture and Ocean Engineering 16 (2024) 100578

Table 5
Computation conditions and power required for pump according to the slit
position.
pS Cj ṁ (kg/s) Vjet (m/s) PPump (W)

0.1c, 0.3c, 0,5c 0.025 0.47 4.86 5.53


0.050 0.66 6.87 15.64
0.075 0.81 8.41 28.74
0.100 0.94 9.72 44.25

round tunnel was reduced.


As the length of the lid increased, the volume of the jets guided to the
jet side increased, and the pressure difference on the propeller surface
and inner structure tended to decrease. Nevertheless, considering the
thin and long cover shape, an excessively long cover length can be ex­
pected to reduce the performance in terms of structure. Therefore, the
cover length was set to lCO. = 0.05 c.

3.3. Performance variation by slit position(pS )

A propeller with the slit positions positioned at 0.1c, 0.3c, and 0.5c
was used to analyze the influence of the slit position, as shown in Fig. 12.
The slit height of the propellers is all 1.0 mm, and the areas of the
tunnels were all the same. When pS = 0.3 c and pS = 0.5 c, there is no
difference in the tunnel shape, but when pS = 0.1 c, there was insuffi­
Fig. 11. Comparison of the flow distribution around the section according to
each cover length at r/R = 0.84 (Lines: CP). cient space in the hydrofoil geometry, and it is impossible to have a
tunnel of the same shape. Therefore, the shape of the tunnel at pS = 0.1 c
was changed so that there would be no difference in the tunnel area.
Computations were performed at Cj = 0.025, 0.05, 0.075, and 0.1 to
analyze the change in propeller performance depending on the injection
volume from the slit position, the computation conditions, and the
required power of the pump, as listed in Table 5. The mass flow rate and
the Vjet were equal for each Cj because the height of the slit and the
injection range were the same for all three propellers, so the power of the
pump was equal.
The changes in propeller performance due to jet injection were
compared in several parts to analyze the interaction between the jet and
the propeller and the effect of the jet separately. The propeller was
divided into three parts: Total, Blade, and Inner structure. Total means
the whole propeller; Blade means the suction side, pressure side, and jet
side; Inner structure means the tunnel, round tunnel, and tip. Fig. 13
compares KT and 10KQ in each part. Fig. 13(a) compares the thrust and
Fig. 12. Hydrofoil geometry for each slit position.
torque of the entire propeller. The propeller torque tends to decrease as
the injection volume increases, which is similar to Lee et al. (2021). The
velocity around the propeller was made dimensionless by the relative moment generated by the propulsive force of the jet injected from the
velocity at r/R = 0.7 in the rθ − x plane and represented by the contour. slit led to decreased propeller torque. On the other hand, in the case of
The surrounding -CP distribution was compared through the lines. The thrust, the increase in thrust decreases as the injection volume increases
jet injected from the slit flowed on the propeller due to the Coanda ef­ because the position of the slit becomes closer to the leading edge. When
fect, and the length of the jet guided in the trailing edge direction pS = 0.1c, the thrust tended to decrease as the injection volume
increased as the cover length increased. The pressure distribution near increased.
the jet side changed as the length of the jet increased, and the point In the case of torque on the surface of the propeller, there is no large
where the pressure drop appeared as the cover length increased was difference depending on the slit position. In contrast, in the case of
guided toward the trailing edge. thrust, there is a slight difference in the increasing tendency depending
The length of the jet injected to the upper part of the jet side differed on the slit position. The thrust generated in the internal structure shows
according to the length of the cover (Fig. 11). As the length of the cover a large difference depending on the slit position, which can be analyzed
increased, the low-pressure region at the upper end of the jet injection from the difference in the hydrofoil geometry depending on the slit
region decreased, and the tendency to concentrate near the trailing edge position and the jet injection direction. Looking at the hydrofoil geom­
appeared. This jet injection tendency caused a high-pressure drop near etry with pS = 0.1 c in Fig. 12, the slit is near the leading edge. There­
the trailing edge, as shown in Fig. 10. In addition, the pressure distri­ fore, the jet direction is injected in the direction opposite to the thrust
bution in the round tunnel was changed depending on the cover length, direction of the propeller because the propeller pitch is small. Hence, the
and the influence of the jet injection method in the round tunnel was thrust of the jet reduces the propeller thrust. On the other hand, when
analyzed. The injected jet through the round tunnel may be sprayed pS = 0.5 c, the slit position was in the center of the chord and consid­
upward in the slit if the cover does not exist, and this causes a large ering the pitch of the propeller and the thrust direction of the jet, it can
pressure difference in the round tunnel. On the other hand, when a cover be expected that there are components in the same direction as the
was attached to prevent the slit from spraying upward, the jet that hit thrust direction of the propeller. In the case of torque in the internal
the cover was injected backward, and the pressure difference in the structure, the jet was injected in the rotation direction of the propeller,

6
J.-H. Lee et al. International Journal of Naval Architecture and Ocean Engineering 16 (2024) 100578

Fig. 13. Comparison of KT and 10KQ for each part of the propeller according to each slit position.

distribution and the iso-surface means V/VR = 0.4. The contour on the
right side of Fig. 14 is the flow velocity around the section, and the lines
mean the -CP distribution. When pS = 0.1c and pS = 0.3c, the jet did not
flow on the jet side and a very high-pressure drop appeared. When pS =
0.1 c, the jet injected from the slit passed through the jet side and was
injected to r/R = 0.9 or higher. The pressure drop due to the jet appeared
near the propeller tip that did not inject the jet. The flow distribution
around the propeller section confirmed the cause of this pressure dis­
tribution. When pS = 0.1c and pS = 0.3c, the jet did not flow the jet side
to the trailing edge and separated. Hence, the jet ejected from the pro­
peller has no narrow and fast flow. The jet flows on the jet side because
of the Coanda effect, but when it encounters an area where it rides up
and down on the jet side, it cannot flow this high curvature until the
trailing edge is separated. On the other hand, when pS = 0.5c, the ge­
ometry of the jet side exists only in the form of descending to the trailing
edge, so even if it is not a narrow and fast-flow jet, the jet will reach the
trailing edge relatively easily because of the Coanda effect. As a result,
unlike other propellers, no separation appeared, and a relatively low-
pressure drop occurred.
The pressure difference near the trailing edge appears depending on
the position of the slit. When pS = 0.1c, the influence of the jet barely
appeared near the trailing edge, but when pS = 0.3c, the high-pressure
region near the trailing edge decreased because of the separated jet at
the middle of the jet side. On the other hand, when pS = 0.5c, a high
pressure near the trailing edge was formed because of the influence of
the jet flowing on the jet side to the trailing edge.
The negative torque at high Cj means that the moment generated by
the jet injected from the slit of the propeller was larger than the total
torque on the propeller surface, as shown in Fig. 13(a). Considering the
actual physical phenomenon, the RPS of the propeller will increase. On
the other hand, in the computations of this study, a sliding mesh was
used, in which the propeller and the grid around the propeller were
Fig. 14. Comparison of the flow distribution around the propeller and the rotated together by specifying the RPS. Therefore, the RPS increased due
section (r/R = 0.81) according to each slit position (Left: CP (contour), V/VR = to jet injection. Hence, the torque of the propeller appeared as a negative
− 0.04 (iso-surface)/Right: tangential velocity (contour), -CP (lines)). value. Accordingly, in this study, this negative torque was considered
additional power to reduce the rotation velocity of the propeller to
even when the position of the slit was different. Hence, the torque tends perform the comparison at the same advance ratio, and the propeller
to decrease as the injection volume increases. efficiency is defined as Eq. (15)
Fig. 14 compares the flow velocity and pressure distribution around TVA
the propeller and propeller section at each slit position. On the left side ηO = (15)
2πn|Q| + PPump
of Fig. 14, the contour on the surface of the propeller shows the pressure

7
J.-H. Lee et al. International Journal of Naval Architecture and Ocean Engineering 16 (2024) 100578

Table 6
Computation conditions and power required for pump according to the slit
height.
hS (mm) Cj ṁ (kg/s) Vjet (m/s) PPump (W)

0.25 0.025 0.23 9.72 11.06


0.050 0.33 13.74 31.29
0.075 0.41 16.83 57.48
0.100 0.47 19.43 88.49
0.50 0.025 0.33 6.87 7.82
0.050 0.47 9.72 22.12
0.075 0.57 11.90 40.65
0.100 0.66 13.74 62.58
0.75 0.025 0.41 5.61 6.39
0.050 0.57 7.93 18.06
0.075 0.70 9.72 33.19
0.100 0.81 11.22 51.09
1.00 0.025 0.47 4.86 5.53
0.050 0.66 6.87 15.64
0.075 0.81 8.41 28.74
0.100 0.94 9.72 44.25
Fig. 15. Propeller efficiency according to each slit position.

Fig. 17. Comparison of the tangential velocity distribution at the slit center
according to the slit height (Cj = 0.05).

3.4. Performance variation by slit height(hS )

A propeller with four slit heights was designed, and the hydrofoil
geometries were compared to analyze the effect of the slit height
following the slit position, as shown in Fig. 16. Equal to the previous
condition, the area of the tunnel was changed together with the slit
height to maintain the area of the slit according to the slit height and a
Fig. 16. Hydrofoil geometry for each slit height. constant ratio to the area of the tunnel. Furthermore, the geometry was
kept the same for the height of the four slit heights to eliminate the
VA means the advanced speed of the propeller. influence of the geometry of the jet side. Because the effect of the tip
Fig. 15 shows the efficiencies of the Coanda propellers at each slit diameter was not as great as the effect of the slit height (Eom et al.,
position and the normal propeller. In Cj = 0.025, 0.05, the propeller 2021), the hydrofoil geometry about the slit height was changed by
efficiency of the Coanda propellers is higher than a normal propeller. decreasing the tip diameter to maintain the geometry of the jet side. The
Nevertheless, the propeller torque was reduced greatly when the jet considered slit heights were hS = 0.25 mm, 0.5 mm, 0.75 mm,
injection volume was high, as confirmed in Fig. 13. The additional and 1.0 mm, confirming that the area of the tunnel decreases as the
power was increased, and the efficiency was reduced greatly at high Cj . height of the slits decreases. The heights of the tunnel, the starting po­
The propeller efficiency of pS = 0.5c, in which the separation of the jet sition of the round tunnel, and the z-axis position of the tunnel were all
did not appear on the jet side, was the highest in all Cj . matched to minimize the influence of the tunnel. Except for the tunnels,
round tunnels, and tip shapes, all other shapes were the same.
A comparison of the performance changes depending on the position
of the slit, at pS = 0.5c, revealed the highest efficiency of the propeller Table 6 lists the computation conditions for each slit height and the
and the smallest pressure drop on the propeller surface. Therefore, the required power of the pump for each slit height. As the height of the slit
position of the silt was fixed to pS = 0.5c. changed, the area of the slit changed, so the injection volume and the
theoretical flow velocity at the slit (Vjet ) change according to the height

8
J.-H. Lee et al. International Journal of Naval Architecture and Ocean Engineering 16 (2024) 100578

generated around the tip tended to decrease as the slit height increased.
Fig. 19 compares the change in performance for each part depending
on each slit height. The overall thrust generated by the propeller tended
to increase as the height increased, and the torque tended to decrease.
The thrust in the inner structure did not differ significantly depending on
the slit height (Fig. 19 (c)), and in the case of the torque, a higher slit
height indicated a greater decrease in torque. In the case of the propeller
blade, however, the thrust increased as the slit height increased, and the
torque decreased. Hence, a higher slit height means a thicker jet injected
from the slit. Currently, the thicker the jet, the more the flow above the
jet side is guided. Therefore, the increase in thrust has increased rela­
tively. In contrast, in the case of torque, it tends to increase as the height
Fig. 18. -CP distribution according to the slit height (Cj = 0.05).
of the slit is lower in all Cj . The thrust and torque that occurred on the
propeller surface were compared to analyze the reason for this tendency.
Figs. 20 and 21 compare the thrust PUA and torque PUA at the lowest
of each slit to maintain the same Cj . Hence, the required power of the
(Cj = 0.025), and highest (Cj = 0.1) to analyze the significant thrust and
pump changes. The injection volume at the same Cj tended to increase as
torque changes on the propeller surface because of the jet injection
the height of the slit increased. On the other hand, the Vjet tended to
volume. At low Cj , the flow injected from the slit was the fastest jet in
decrease as the area of the slit increased. Owing to this effect, the
hS = 0.25 mm propeller, among the four slit heights, and the guided
required power of the pump for injecting the jet concerning the height of
downwash from the tip generates the highest thrust around the tip. Even
each slit tends to decrease as the height of the slit increases at the same
in the case of torque, relatively higher torque was generated around the
Cj .
tip of the hS = 0.25 mm propeller, which has the fastest jet speed at a low
Fig. 17 compares the actual flow velocity at the slit at the center of Cj and thrust. When the jet injection volume increased, however, the jet
the slits according to the height of each slit at Cj = 0.05. As shown in
injected from the slit flowed on the jet side due to the Coanda effect and
Table 6, the jet velocities at the slit do not show a linear relationship
thrust was generated. Regarding the thrust, a higher slit height meant a
depending on the slit height. Therefore, the jet velocities at the slit do
thicker jet, and the more thrust generated on the jet side tended to in­
not show a linear relationship depending on the slit height, as shown in
crease. The torque on the jet side and around the tip, which was affected
Fig. 17. By matching the ratio of the slit area change to the slit height to
by the flow velocity of the high-speed jet, caused the highest torque
the tunnel area, it was possible to confirm the same jet injection form in
generated on the surface of the hS = 0.25 mm propeller at all Cj . Even in
which the jet was injected from r/R = 0.55 or more.
the case of thrust, a positive thrust was generated near the tip because of
Fig. 18 compares the surface pressure distribution of the propeller
the high jet velocity of a low slit height. At the same time, it was possible
according to the slit height. A higher slit height indicates a greater
to confirm the negative thrust distribution that appeared when the jet
pressure drop generated on the surface of the propeller, particularly on
applied pressure to the jet side in the opposite direction of the thrust.
the jet side. As the slit height increases, the jet becomes thicker at the
This thrust reduction region tended to decrease as the slit height
same Cj , which increases the downwash above the jet side caused by the
increased. Hence, the thrust and torque of the pressure side do not show
jet and increases the pressure drop on the jet side. On the other hand, the
a large difference due to the height of the slit.
pressure drop around the tip appeared most at hS = 0.25 mm, which is
Fig. 22 compares the flow distribution at r/R = 0.81. There was no
the lowest slit height. As shown in Fig. 17, the jet was injected fastest at
significant pressure difference between the tip where the downflow was
hS = 0.25 mm, showing that the variation of flow occurs due to the
induced by the jet and the suction side and pressure side except for the
Coanda effect because of the curvature of the tip. The pressure drop
jet injection region where the jet was injected. As the jet became thicker,

Fig. 19. Comparison of KT and 10KQ for each part of the propeller according to each slit height.

9
J.-H. Lee et al. International Journal of Naval Architecture and Ocean Engineering 16 (2024) 100578

Fig. 20. Comparison of the thrust PUA distribution on the suction side and Fig. 21. Comparison of the torque PUA distribution on the suction side and
pressure side according to slit height. pressure side according to the slit height.

the pressure tended to decrease at the upper part of the jet side, and the height hS = 0.75 mm that can realize AS /AT = 1.0. The slit area did not
thrust increased because of the pressure difference between the jet side change because there was no change in slit height. Therefore, the in­
and the pressure side due to such an effect. That is, the thrust at the jet jection volume did not change depending on the ratio of the slit area to
side was dominated by the thickness of the jet rather than the flow ve­ the tunnel area. As a result, there was no difference in the required
locity at the upper part of the jet side. power of the pump for each area ratio, as shown in Table 7.
Fig. 23 compares the propeller efficiency according to the slit height. Fig. 25 compares the jet velocity in the center of the slit based on the
In all Cj , the efficiency of hS = 1.0 mm propeller, in which the thrust ratio of slit area to tunnel ratio (AS /AT ). As the slit area becomes wider
increased due to the Coanda effect of the jet and the torque on the than the tunnel area, the jet tended to be concentrated and injected at a
propeller surface appeared small, was the highest, and a lower slit height higher r/R. When AS /AT = 1.0, there was almost no area where the jet
indicated lower efficiency. In addition, the remaining propellers, except was sucked into the slit, whereas when AS /AT = 2.2, the jet was sucked
hS = 1.0 mm propeller showed lower efficiency than a normal propeller into the slit until around r/R = 0.68, and the jet was injected at higher
on all Cj . Hence, the lower the slit height, the more Vjet increases at the r/R = 0.68 with a very fast velocity.
same Cj , as shown in Table 6. The required power of the pump increased When the slit area was smaller than the tunnel area (AS /AT = 1.0),
according to Eqs. (12) and (13) when Vjet increases, which causes the the area of the outlet was small so that the flow velocity inside the tunnel
efficiency of the propeller to decrease. decreased and the pressure increased. The difference between the jet
injection speed at high r/R and low r/R was not large because of the
relatively high pressure in the tunnel. On the other hand, the internal
3.5. Performance variation by slit–tunnel area ratio(AS /AT ) pressure of the tunnel decreased when the area of the slit was larger than
the area of the tunnel (AS /AT = 2.2), and the flow velocity increased
Fig. 24 and Table 7 show the hydrofoil geometry and computation because of the narrowed area of the tunnel. Therefore, the jet is
conditions of the propeller used to analyze the performance change due concentrated on the high r/R because of the centrifugal force generated
to the ratio of the slit area (AS ) to the tunnel area (AT ). The slit area by the propeller rotation. At low r/R, where the jet was not injected, the
means the area of the twisted rectangle whose height is the slit height pressure inside the slit became low, and rather a flow sucked into the slit
and length is r/R = dd0.5 to 0.9. All other geometries are the same from the outside was generated.
except for the geometry of the tunnel when observing only the effect of The pressure distribution on the propeller surface also differs
the ratio of slit area to tunnel ratio. The ratio of the slit area to the tunnel depending on the injection pattern depending on the ratio of slit area to
area considered was AS /AT = 1.0, 1.6, and 2.2, and a higher slit height tunnel area, which was compared in Fig. 26. The pressure drop is the
indicated better propeller performance. On the other hand, when hS = smallest, and the pressure drop occurs on the jet side in the widest area
1.0 mm, the tunnel area was too large under the condition AS / AT = when the jet velocity is relatively slow and widespread in AS /AT = 1.0,
1.0. Hence, it cannot be implemented inside the hydrofoil geometry. as shown in Fig. 25. On the other hand, in the case of AS /AT = 2.2,
Therefore, the performance was evaluated by applying the highest slit

10
J.-H. Lee et al. International Journal of Naval Architecture and Ocean Engineering 16 (2024) 100578

Fig. 24. Hydrofoil geometry for each slit-tunnel area ratio.

Table 7
Computation conditions and power required for a pump according to the slit-
tunnel area ratio.
AS /AT Cj ṁ (kg/s) Vjet (m/s) PPump (W)

1.0, 1.6, 2.2 0.025 0.41 5.61 6.39


0.050 0.57 7.93 18.06
0.075 0.70 9.72 33.19
0.100 0.81 11.22 51.09

Fig. 22. Comparison of the flow distribution around the section according to
each slit height at r/R = 0.81 (Cj = 0.1, Contour: tangential velocity, Lines: CP ).

Fig. 25. Velocity distribution at the center of the slit (Cj = 0.050).

Fig. 23. Propeller efficiency according to each slit height.

where the is at high r/R, a very high-pressure drop appears near the
trailing edge in the region where the jet is intensively injected. There­
fore, propeller damage can occur because of cavitation and the thin
shape of the trailing edge.
Fig. 27 compares the propeller performance for each injection
pattern. Fig. 27(c) compares the thrust generated in the internal struc­ Fig. 26. -CP distribution according to each slit-tunnel area ratio (Cj = 0.050,
ture with the torque. A large difference in the torque reduction of the Isosurface: V/Vjet = − 0.4).

11
J.-H. Lee et al. International Journal of Naval Architecture and Ocean Engineering 16 (2024) 100578

Fig. 27. Comparison of KT and 10KQ for each part of the propeller according to each slit-tunnel area ratio.

Fig. 28. Comparison of the thrust PUA distribution on the suction side and Fig. 29. Comparison of the torque PUA distribution on the suction and pressure
pressure side according to slit-tunnel area ratio. sides according to slit-tunnel area ratio.

propeller depending on the injection pattern was observed. Even if the judged that this difference becomes even larger as the injection volume
jet is injected at the same Cj , the method of injecting a narrowed and increases.
faster jet was more effective in reducing the torque of the propeller than The cause of the change in the thrust and torque of the propeller due
the method of slow and wide injection. Considering the shape of the to the jet injection pattern was analyzed. Figs. 28 and 29 show the thrust
round tunnel and the injection pattern, a smaller slit area to tunnel ratio PUA and the torque PUA according to the injection pattern. When AS /
will allow the jet to provide more force in the thrust vectoring direction AT = 1.0, which disperse the same Cj and injects a jet, the maximum
of the propeller. thrust appearing in the jet injection area is the smallest. As shown in
Fig. 27(b) compares the thrust and torque generated on the propeller Fig. 26, when the slit-tunnel area ratio is different, the thrust PUA and
blade surface. The torque generated on the propeller surface tends to Torque PUA on the suction side increase like the injection pattern. In the
increase as the AS /AT increases. In the case of AS /AT = 2.2, the thrust case of AS /AT = 2.2, where the jet was concentrated at a high r/R, the
appears to be highest at a low injection volume, but the thrust increase jet velocity was relatively high, so downwash was induced at the upper
generated on the propeller blade surfaced tends to decrease as the in­ slit. The thrust and torque increased in areas where high-pressure drops
jection volume increases. Considering the slope of the graph, it was

12
J.-H. Lee et al. International Journal of Naval Architecture and Ocean Engineering 16 (2024) 100578

Table 8
Computation conditions and required power for the pump for the injection
range.
Slit type Cj ṁ (kg/s) Vjet (m/s) PPump (W)

B, M, T 0.025 0.28 7.99 9.10


0.050 0.40 11.30 26.11
0.075 0.49 13.84 47.97
0.100 0.57 15.98 73.85
BM, MT 0.025 0.40 5.61 6.39
0.050 0.57 7.93 18.06
0.075 0.70 9.72 33.19
0.100 0.81 11.22 51.09
BMT 0.025 0.49 4.61 5.25
0.050 0.70 6.52 14.86
0.075 0.85 7.99 27.29
0.100 0.99 9.23 42.02

when the injection volume was increased, and a strong jet was injected.
In AS /AT = 1.0, where the jet was injected over the entire slit, as the
injection volume increased, the area generating positive thrust increased
near the trailing edge, and the thrust tended to increase. Nevertheless,
for AS /AT = 2.2, concentrating at high r/R, the torque decreased
because of the increase in the area generating negative torque. In
conclusion, the thrust and torque variables depend on the injection
volume and pattern, but the torque mainly varies by the injection
Fig. 30. Comparison of the flow distribution around the section according to
pattern.
the slit position at r/R = 0.82 (Cj = 0.100 , Lines: CP ).
The cause of thrust and torque changes on the pressure side was
examined. Fig. 30 presents the flow around the propeller. It shows the
flow at r/R = 0.82 where the difference in the injection pattern is clear.
Differences in the jet velocity and thickness occurred as the slit-tunnel
area ratio increased, which increased the pressure drop at the jet side.
The jet flowed on the pressure side, and the velocity also increased.
Owing to the velocity, the pressure on the propeller surface became
lower, which reduced the pressure difference with the suction side. As
the slit-tunnel area ratio increased, the thrust was reduced at high in­
jection rates, as shown in Fig. 27(b) and Fig. 28.
Fig. 31 compares the efficiency for the slit-tunnel area ratio. In
condition Cj = 0.025 from AS /AT = 2.2, an efficiency improvement of
approximately 38.1 % was confirmed compared to a normal propeller.
Under other conditions, except for Cj = 0.025 from AS /AT = 2.2, the
efficiency was lower than normal propellers because the negative torque
was considered as the additional power, as expressed in Eq. (15).

3.6. Performance variation according to the slit type

The following six types of jet injection propellers were devised to


Fig. 31. Propeller efficiency according to each slit-tunnel ratio.
understand the performance change depending on the injection position
and range. Depending on the injection range, r/R = 0.3 to 0.5 was set for
appeared. On the other hand, there was no difference in thrust PUA and
the bottom, r/R = 0.5 to 0.7 for the middle, and r/R = 0.7 to 0.9 for the
torque PUA at the pressure side depending on the slit-tunnel area ratio
top, and they were called B, M, and T propellers, respectively. Here, the
because the effect of the jet on the propeller was low at a low Cj .
bottom and middle areas, middle and top areas, and propellers spraying
The high thrust and torque were generated near the slit curvature all three areas were called BM, MT, and BMT propeller, respectively.

Fig. 32. Propeller geometry for the injection range.

13
J.-H. Lee et al. International Journal of Naval Architecture and Ocean Engineering 16 (2024) 100578

Fig. 33. -CP distribution for each injection range (Cj = 0.05).

Fig. 34. Comparison of KT and 10KQ for each part of the propeller for each injection range.

Fig. 32 presents the propeller geometry. Table 8 lists the calculation PUA. The torque decreased in all injection ranges, but the thrust showed
conditions for each injection range. A wider injection range meant a various patterns. The T Propeller injected at0 r/R = 0.7 or higher
larger slit area, so the mass flow rate was higher than the other pro­ showed a sharp increase in thrust as Cj increased. The propeller rotation
pellers with the same Cj , but the required power for the pump was and jet injection generated a high relative velocity at the upper end of
smaller because of the slower Vjet . the slit, which induced a large downwash. In addition, when comparing
The pressure distribution on the propeller surface for each slit type the PUA distribution appearing on the pressure side, at r/R = 0.7 as the
was compared, as shown in Fig. 33. A comparison of the B, M, and T injection port, smaller thrust and torque PUA areas were noted on the
propellers with different injection positions showed a pressure drop pressure side. This was observed on the MT and BMT propellers, which
because the jet increased as r/R increased. For the B propeller, which jets share the injection range with the T propeller.
at the lowest r/R, almost no pressure drop was observed because of the The thrust of the B propeller blade decreased, even when Cj
jet, but a high pressure occurred on the jet side. This trend was increased, as shown in Fig. 34(b), which can be analyzed through the jet
confirmed for the BM and MT propellers. An intensive pressure drop injection area and pressure side. As shown in Fig. 33, the region where
occurred near the end in the jet injection region for the T and MT pro­ high pressure was generated decreased the thrust in the injection region.
pellers, which injected the jet at high r/R, and the BMT propeller had a Furthermore, a jet was injected at low r/R to induce flow on the pressure
wider area of pressure drop, but the value was relatively small. This side, which caused a thrust reduction while increasing the velocity. On
result was caused by an increase in mass flow rate with the same Cj , but a the other hand, the thrust generated by the inner structure was highest
decrease in jet velocity occurred as the slit area increased, as listed in in the propeller, and due to the short tunnel length, the force exerted by
Table 8. the jet on the rounded tunnel in the thrust direction was considered to
The propeller performance was compared according to each slit type, have the greatest influence. The BMT propeller, injected jet at all r/R
as shown in Fig. 34. Figs. 35 and 36 show the thrust PUA and torque positions, was indicated a thrust advantage.

14
J.-H. Lee et al. International Journal of Naval Architecture and Ocean Engineering 16 (2024) 100578

Fig. 35. Comparison of thrust PUA distribution on the suction side and pressure side according to the injection ratio.

15
J.-H. Lee et al. International Journal of Naval Architecture and Ocean Engineering 16 (2024) 100578

Fig. 36. Comparison of the torque PUA distribution on the suction side and pressure side according to the injection ratio.

16
J.-H. Lee et al. International Journal of Naval Architecture and Ocean Engineering 16 (2024) 100578

Fig. 37 compares the propeller efficiencies according to the propeller


type, showing that the BMT propeller is higher. At high Cj , where
negative torque appears, the propeller efficiency decreases. Hence, the
selection of Cj is important for increasing efficiency.

3.7. Optimization of the slit

A propeller has a thrust that must be generated at the design speed,


and the thrust must be matched to compare the propeller performance.
Therefore, the propeller performance was compared after matching the
thrust with the normal propeller, which is a propeller that does not inject
a jet within 1% or so. Lee et al. (2021) also compared the performance
with the propeller showing the highest efficiency. Slit shape optimiza­
tion of a jet injection propeller with a NACA66 was performed based on
the results of the previous section.
The case 2 geometry of Fig. 7 and a cover length of 0.05C were
applied to improve the high-pressure drop developed on the jet side. For
the slit type, the BMT type with the highest propeller efficiency was
Fig. 37. Propeller efficiency for each injection range.
adopted, and the slit height was set as large as possible. The slit shape
was optimized while adjusting the slit-tunnel area ratio. Table 9 and
Table 9 Fig. 38 show the slit shape and specifications of the optimized
Optimized slit shape specifications. propellers.
The propulsion performance of the three propellers was compared,
Parameter Lee et al. (2021) Optimized propeller
as shown in Table 10. The injection volume was adjusted, and the thrust
lCo. 0.00c 0.05c
was matched to harmonize the thrust with the normal propeller within
pS 0.5c
hS (mm) 1.0 0.8 1%. For the optimized propeller, the injection volume was increased
AS / AT 1.60 1.58 over the propeller of a previous study (Lee et al., 2021) to match the
thrust, but the slit shape was the BMT type with a broader exit area; Vjet
was reduced, and the power requirements decreased. The torque
The tendency of the torque generated on the propeller surface generated by the propeller was almost non-existent, meaning that very
differed according to the slit type and relative velocity. The torque on little power was required on the shaft to rotate the propeller. In other
the propeller surface increased when the jet was injected in a region words, the propeller can be rotated by the thrust of the jet injected from
where the relative velocity flowing into the propeller was large, or when the slit, and the propeller efficiency is maximized. In the case of the
the injection area was wide. The widest jetting BMT propeller has the optimized propeller, assuming a pump efficiency of 1.0, the efficiency
widest area to generate thrust and torque, but the thrust generated was was approximately 0.811, an approximately 35 % efficiency improve­
low at Cj = 0.025. On the other hand, as Cj increased, the thrust increase ment over the normal propeller. Assuming a high-efficiency pump and
rate was high, so Cj = 0.100 generated the maximum thrust. considering a pump efficiency of 0.75, there was an approximately 2 %

Fig. 38. w/jet hydrofoil geometry with an optimized slit shape.

Table 10
Performance comparison of optimized propeller with the matched thrust of a normal and previous propeller (NACA66 (MOD, t/c = 0.12)).
Type Cj KT 10KQ ṁ (kg/s) Vjet (m/s) ηP PPump (W) ηO

Normal propeller – 0.060 0.064 – – – – 0.601


Lee et al. (2021) 0.052 0.060 0.005 0.68 7.01 1.00 16.74 0.653
Optimized propeller 0.053 0.067 − 0.001 0.71 6.22 1.00 14.29 0.811
0.75 18.22 0.665

17
J.-H. Lee et al. International Journal of Naval Architecture and Ocean Engineering 16 (2024) 100578

showed relatively high performance, even at low Cj . On the other hand,


at high Cj , the thrust and torque were reduced because the jet flow was
induced on the propeller surface. As a result, it was ineffective in
improving the propeller performance. In contrast, the high r/R type
effectively increased the thrust and had little effect on the pressure side.
In addition, a larger injection area resulted in a greater jet thrust and a
smaller torque.
These results were combined to optimize the injection slit shape, and
the efficiency was compared with the normal propeller and the propeller
reported by Lee et al. (2021). When the pump efficiency was not
considered, it showed approximately 35 % efficiency improvement over
the normal propeller and approximately 24 % compared to Lee et al.
(2021). Considering a pump efficiency of 75 %, it showed an approxi­
mately 2 % increase in efficiency compared to the normal propeller.
This study confirmed the propeller performance improvement using
the Coanda effect through an optimized slit shape. Future research will
Fig. 39. -CP distribution of the optimized slit shape with the matched thrust. analyze the propeller performance by varying the propeller geometry.

efficiency improvement. Declaration of competing interest


The pressure distributions of three propellers were compared, as
shown in Fig. 39. The propeller optimized in this study greatly improved The authors declare that they have no known competing financial
the pressure drop compared to the propeller reported by Lee et al. interests or personal relationships that could have appeared to influence
(2021). The pressure drop generated on the jet side was also not large the work reported in this paper.
compared to normal propellers. On the other hand, the downstream flow
was induced because the jet had a higher velocity than the propeller Acknowledgement
rotation velocity, and a high-pressure drop around the slit tip was
generated. Hence, the cavitation performance is expected to deteriorate This work was supported by the National Research Foundation of
around the slit tip. Korea (NRF) grant (No. 2022R1F1A1074865) funded by the Ministry of
Science and ICT, Republic of Korea.
4. Conclusions
References
In this study, the hydrofoil geometry modification and slit shape
optimization were performed to improve the high-pressure drop that Baltazar, J., Rijpkema, D., de Campos, J.F., 2018. On the use of the γ-Reθtransition model
occurs during jet injection and increase the propulsion efficiency of the for the prediction of the propeller performance at model-scale. Ocean Engineering
170, 6–19.
propeller.
Baltazar, J., Rijpkema, D., Falcão de Campos, J., 2021. Prediction of the propeller
First, this study improved the cross-sectional shape of the propeller performance at different Reynolds number regimes with RANS. J. Mar. Sci. Eng. 9
by reducing curvature. The high-pressure drop on the propeller surface (10), 1115.
Bhattacharyya, A., Krasilnikov, V., Steen, S., 2016. Scale effects on open water
was also improved despite the increased injection volume. In the case of
characteristics of a controllable pitch propeller working within different duct
the installed cover, the injection volume required to match the thrust designs. Ocean Engineering 112, 226–242.
was similar. In addition, as the length of the cover increased, the amount Burnazzi, M., Radespiel, R., 2014. Design and analysis of a droop nose for Coanda flap
of flow increased to the trailing edge because of the Coanda effect. On applications. J. Aircraft 51 (5), 1567–1579.
Burnazzi, M., Radespiel, R., 2015. Synergies between suction and blowing for active
the other hand, if the cover is too long, it could lead to structural high-line flaps. CEAS Aeronautical Journal 6 (2), 305–318.
instability, vibration, and noise-induced problems arising from the Celik, I.B., Ghia, U., Roache, P.J., Freitas, C.J., 2008. Procedure for estimation and
cover. reporting of uncertainty due to discretization in CFD applications. Journal of Fluids
Engineering Transactions of the ASME 130 (7).
Second, the performance was compared depending on the slit posi­ Eom, M.J., Paik, K.J., Lee, J.H., 2020. Numerical study on the lift-drag and cavitation
tion. When the slit was located near the leading edge, the injected jet performances of a two-dimensional hydrofoil using the Coanda effect. Journal of
was separated without flowing to the trailing edge. Therefore, a high- Advanced Marine Engineering and Technology 44 (6), 457–466.
Go, J.S., Yoon, H.S., Jung, J.H., 2017. Effects of a duct before a propeller on propulsion
pressure drop was expected, and the cavitation performance deterio­ performance. Ocean Engineering 136, 54–66.
rated. By contrast, when the slit was positioned at 0.5c, the jet that Joung, T.H., Kang, S.G., Lee, J.K., Ahn, J., 2020. The IMO initial strategy for reducing
sprayed from the slit flowed to the trailing edge of the propeller, causing Greenhouse Gas (GHG) emissions, and its follow-up actions towards 2050. Journal of
International Maritime Safety, Environmental Affairs, and Shipping 4 (1), 1–7.
the largest thrust increase due to the Coanda effect. Kadivar, E., Timoshevskiy, M.V., Pervunin, K.S., el Moctar, O., 2020. Cavitation control
Third, at the same Cj , when the slit height was increased, the thrust using cylindrical cavitating-bubble generators (CCGs): experiments on a benchmark
increased due to the Coanda effect, and the total torque decreased. This CAV2003 hydrofoil. Int. J. Multiphas. Flow 125, 103186.
Kim, D.H., Jeon, G.M., Park, J.C., Shin, M.S., 2021. CFD simulation on predicting POW
is because a higher slit height resulted in a greater pressure drop on the
performance adopting laminar-turbulent transient model. Journal of the Society of
jet side, a greater pressure difference between the pressure side, suction, Naval Architects of Korea 58 (1), 1–9.
and jet side, and greater thrust. On the other hand, the surface torque Kuiper, G., 1981. Cavitation inception on ship propeller model. PhD Thesis. Technical
tended to decrease when the velocity of the jet decreased. The jet in­ Univ., Delft, The Netherlands.
Lee, J.H., Paik, B.G., Kim, K.Y., Kim, M.J., Kim, S., Lee, S.J., 2019. Experimental study on
jection pattern at the slit, depending on the slit-tunnel area ratio, was supercavitated body with static angle-of-attack. Journal of the Society of Naval
changed. When a concentrated jet was injected, the pressure drop was Architects of Korea 56 (6), 541–549.
stronger on the jet side, and the propeller performance was higher at low Lee, J.H., Paik, K.J., Lee, S.H., Kim, G.H., Cho, J.H., 2021. Experimental and numerical
study on the performance change of a simple propeller shape using the Coanda
Cj . Nevertheless, when Cj increased, the flow was induced on the pres­ effect. Appl. Sci. 11 (9), 4112.
sure side; the pressure difference and thrust rise rate decreased. Lu, S.P., Wang, W., Wang, X.F., 2018. Experiment research on cavitation control by
Finally, the propeller performance was compared with the injection active injection. In: In the 10th International Symposium on Cavitation.
Maltsev, L.I., Dimitrov, V.D., Milanov, E.M., Zapryagaev, I.I., Timoshevskiy, M.V.,
range and position of the injection port. A low r/R type was injected, Pervunin, K.S., 2020. Jet control of flow separation on hydrofoils: performance
which did not affect the propeller performance significantly, but the jet evaluation based on force and torque measurements. J. Eng. Thermophys. 29 (3),
424–442.

18
J.-H. Lee et al. International Journal of Naval Architecture and Ocean Engineering 16 (2024) 100578

Menter, F.R., 1994. Two-equation eddy-viscosity turbulence models for engineering Seo, D.W., Oh, J.K., Lee, S.H., 2010. A study for improvement of lift performance of a
applications. AIAA J. 32 (8), 1598–1605. Horn-type rudder with the Coanda effect. Journal of the Society of Naval Architects
Menter, F.R., Smirnov, P.E., Liu, T., Avancha, R., 2015. A one-equation local correlation- of Korea 47 (4), 543–552.
based transition model. Flow, Turbul. Combust. 95, 583–619. Seo, D.W., Lee, S.J., Oh, J., 2016. Performance analysis of stabilizer fin applied Coanda
Mizzi, K., Demirel, Y.K., Banks, C., Turan, O., Kaklis, P., Atlar, M., 2017. Design system. Journal of Ocean Engineering and Technology 30 (1), 18–24.
optimisation of propeller boss cap fins for enhanced propeller performance. Appl. Timoshevskiy, M.V., Zapryagaev, I.I., Pervunin, K.S., Markovich, D.M., 2016. Cavitation
Ocean Res. 62, 210–222. control on a 2D hydrofoil through a continuous tangential injection of liquid:
Oh, J.K., Kim, H.C., 2010. A study on the high lifting device equipped with the trailing experimental study. In: AIP Conference Proceedings, vol. 1770. AIP Publishing LLC,
edge rotor for the enhancement of circulation control. Journal of the Society of 030026. No. 1.
Naval Architects of Korea 47 (4), 533–542. Timoshevskiy, M.V., Zapryagaev, I.I., Pervunin, K.S., Maltsev, L.I., Markovich, D.M.,
Oh, J.K., Noh, J.K., 2015. On the lift enhancement technique of the trailing edge rotor of Hanjalić, K., 2018. Manipulating cavitation by a wall jet: experiments on a 2D
two-dimensional hydrofoil. Journal of the Korean Society of Marine Environment hydrofoil. Int. J. Multiphas. Flow 99, 312–328.
and Safety 21 (2), 200–206. Wang, W., Tang, T., Zhang, Q., Wang, X., An, Z., Tong, T., Li, Z., 2020. Effect of water
Paik, B.G., Kim, M.J., Jung, Y.R., Lee, S.J., Kim, K.Y., Ahn, J.W., Seol, H.S., Kim, K.S., injection on the cavitation control: experiments on a NACA66 (MOD) hydrofoil. Acta
2018. Fundamental studies for ventilated supercavitation experiments in new high- Mech. Sin. 36, 999–1017.
speed cavitation tunnel. Journal of the Society of Naval Architects of Korea 55 (4), Wang, X., Walters, K., 2012. Computational analysis of marine-propeller performance
330–340. using transition-sensitive turbulence modeling. J. Fluid Eng. 134 (7), 071107.
Pant, C.S., Frankel, S.H., 2021. Interaction between surface blowing and re-entrant jet in Xu, H., Luo, K., Dang, J., Li, D., Ye, C., Huang, C., 2021. Comparison of supercavity
active control of hydrofoil cavitation. Ocean Engineering 242, 110087. geometry and gas leakage behavior between the double-cavity flow pattern and the
Pham, V.D., Hong, J.W., Hilo, A.K., Ahn, B.K., 2022. Numerical study of hot-gas wake-closure flow pattern. AIP Adv. 11 (5).
ventilated supercavitating flow. Int. J. Nav. Archit. Ocean Eng. 14, 100470. Zhou, J.J., Yu, K.P., Yang, M., 2011. Numerical simulation of gas leakage mechanism of
ventilated supercavity under the condition of low Froude number. Eng. Mech. 28 (1),
251–256.

19

You might also like