Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Ocean Engineering 175 (2019) 149–162

Contents lists available at ScienceDirect

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

Accurate assessment of ship-propulsion characteristics using CFD T


a a a a,b,∗ a
Andro Bakica , Inno Gatin , Vuko Vukčević , Hrvoje Jasak , Nikola Vladimir
a
University of Zagreb, Faculty of Mechanical Engineering and Naval Architecture, Ivana Lučića 5, Zagreb, Croatia
b
Wikki Ltd, 459 Southbank House, SE1 7SJ, London, United Kingdom

A R T I C LE I N FO A B S T R A C T

Keywords: In this paper RANS based CFD calm-water resistance and propulsion simulations of two benchmark cases
Ship hydrodynamics (container ship KCS and bulk carrier JBC) are presented. Two-phase flow is handled using the Volume-of-Fluid
NavalHydro method, while the discontinuities at the free surface are resolved using the Ghost Fluid Method. Propeller is
Foam–extend modeled using an idealized pressure-jump actuator disk model, allowing large time step size suitable for naval
Wake field
hydrodynamics with the preservation of accuracy and CPU time efficiency. Computations are performed with the
Actuator-disk
Self-propulsion
Naval Hydro Pack developed in FV framework of foam-extend, a community driven fork of OpenFOAM. Model
scale KCS resistance and wave pattern is evaluated with a detailed analysis of the bare hull wake field following
ITTC recommendations. Self propulsion simulation is carried out for the JBC case comparing the global features
of the flow with the experiment.

1. Introduction the Korea Research Institute for Ships and Ocean Engineering (KRISO)
Container Ship case. Experimental data for KRISO Container Ship (KCS)
For the last few decades computer technology has been ex- is openly available at (Chalmers University of Technology and
ponentially evolving in an unrestrained manner bringing massive Gothenburg, 2010) for a different number of cases ranging from simple
Central Processing Unit (CPU) power for acceptable price to the regular towing tank tests in calm water or head seas to wake and flow field
high-end users. Computational Fluid Dynamics (CFD) is greatly de- study (Kim et al. (2001)) and self propulsion (Hino (2005)).
pendent on CPU power since the basis in solving a fluid flow of any kind Starke et al. (2006) used a FD type solver PARNASSOS for Reynolds
is found in a non-linear Navier-Stokes (NS) equations with extensions, Averaged Navier Stokes (RANS) equations and obtained accurate re-
such as two phase-free surface flow. These require high computational sults when comparing computed and experimental wave elevation. In-
effort to acquire a satisfactory solution. teresting observation is made comparing viscous and non-linear po-
According to Larsson et al. (2013), on the Gothenburg 2010 tential flow solver on a stern wave system where a clear overestimation
Workshop approximately 80% of the participants used FV discretisa- of wave height in potential flow can be seen due to neglecting the
tion, while the rest relied on FD discretisation. Although the FV diffusion term in the NS equations. Also, in the study, it is obvious that
methods are only second-order accurate (see Jasak (1996)) simplicity of two equation turbulence model produces much better results for the
the method in developing complex geometries and straightforward wake field analysis then a simpler one equation model. Ozdemir et al.
implementation of complex mathematical models is deemed invaluable (2016) employed a commercial code STAR CCM+ and obtained sa-
since, in most engineering applications, second order accuracy is found tisfactory results for resistance and wave elevation using k-ε turbulence
sufficient. model (Jones and Launder (1972)). However, results for the wake field
In order to validate the FV based CFD a set of benchmark cases is are fairly diffused, most probably due to the first order discretisation
experimentally tested and data is made publicly available (e.g. Kim that was used for advection.
et al. (2001), Van et al. (1998), Olivieri et al. (2001)) which, together Enger et al. (2010) carried out a detail investigation with Star
with workshop gatherings in Gothenburg (Larsson et al. (2013)) or CCM + using different turbulence models where the best results were
Tokyo (Hino (2005), Larsson et al. (Tokyo, 2015 et al., 2015)), ensure obtained for k-ε model with a y+ value ranging from 40 to 60. Geo-
an improvement of numerical tools. metrical properties of the boundary layer were set with three para-
Wide range of publications is available on the topic of the analysis of meters: overall thickness (20 mm), expansion rate (1.5) and number of


Corresponding author. University of Zagreb, Faculty of Mechanical Engineering and Naval Architecture, Ivana Lučića 5, Zagreb, Croatia.
E-mail addresses: andro.bakica@fsb.hr (A. Bakica), inno.gatin@fsb.hr (I. Gatin), vuko.vukcevic@fsb.hr (V. Vukčević), h.jasak@wikki.co.uk,
hrvoje.jasak@fsb.hr (H. Jasak), nikola.vladimir@fsb.hr (N. Vladimir).

https://doi.org/10.1016/j.oceaneng.2018.12.043
Received 9 July 2018; Received in revised form 28 November 2018; Accepted 16 December 2018
0029-8018/ © 2019 Elsevier Ltd. All rights reserved.
A. Bakica, et al. Ocean Engineering 175 (2019) 149–162

layers (6). Through the given parameters the resulting first layer height actuator disk forces as a momentum jump condition for the normal and
is 0.9 mm. Test was made by lowering the number of layers to four, but tangential component at the location of the propeller mid-plane which
keeping the overall thickness and expansion rate the same. Discrepancy is applied only on the adjacent mesh faces which significantly simplifies
in results was evident on the coarsest mesh, increasing the error by the numerical implementation and simulation setup. Only the mean
1.6%. features of the flow are considered, where instead of using external
In Guo et al. (yu Guo et al., 2016) an experiment on KCS hull is algorithms (Badoe et al. (2014), Rijpkema et al. (2013), Krasilnikov,
made by measuring the wake field using Particle Image Velocimetry (Krasilnikov)), the entire computation is being done inside a single CFD
(PIV) and comparing the results with Kim et al. (2001), where a pitot simulation. In order to avoid the complex coupling procedure, com-
tube, stereoscopic camera and other equipment is used. Sufficient pletely different approach will be considered for self-propulsion case by
agreement was found between the data. Obtained results are then applying the correction on the inflow velocity based on the momentum
compared with the computed result from STAR CCM + using the k-ω transfer in the vicinity of the actuator disk which will be described in
Shear Stress Transport (SST) model (Menter (1994)) where good the Sec. 2.3 and by using the Proportional-Integral (PI) controller to
agreement was found between the wake fields, but no quantitative data achieve the desired thrust. Only the most important characteristics of
comparison was given beside the visual inspection. A somewhat dif- the self-propulsion condition will be evaluated in order to achieve high
ferent approach is seen in Banks et al. (2010) using an ANSYS CFX CPU efficiency, thus ignoring the local flow features near the propeller.
RANS flow solver on a structured grid. Results were compared for the k- The current study is divided into three sections. The first section is
ω SST model and the baseline model. Similar results are produced by dedicated to mathematical equations and numerical discretisation with
both models for resistance and wave elevation. Further on, for a self a short description of meshing procedure. The second section describes
propulsion simulation using body force model, it is concluded that in the KCS case setup from the Gotenburg 2010 Workshop for resistance
order to achieve accurate computation of propulsion forces a greater (Case 2.2a) on a coarse mesh with wave elevation and wake field (Case
level of detail is needed in the propeller inflow region. Wave elevation 2.1) evaluated on a more refined mesh in order to compute a more
is correctly captured in almost every study and depends mostly on mesh accurate result. The third section is dedicated to Japanese Bulk Carrier
grid density in the vertical direction near the free surface (also advised (JBC) resistance (Case 1.1a) and self-propulsion (Case 1.5a) from the
in ITTC (ITTC, 2011a)). Tokyo 2015 Workshop with only the global aspects of the flow being
Regarding self-propulsion cases, modeling the propeller behind the evaluated on a coarse mesh.
hull in sliding mesh technique (Yang et al. (2010)) or using dynamic KCS case is chosen for the wake field analysis due to invaluable data
overset grids (Caricca et al. (Carrica et al., 2015), Shen et al. (2015)) are quality with proven high degree of accuracy. However, for the KCS
obviously one of the first choices. However, problem is in dealing with there was no equal experimental setup for the resistance and self-pro-
different time scales of the flow around the hull and the propeller. In pulsion test without the rudder to properly evaluate the thrust deduc-
order to simulate physically correct interaction between the propeller tion coefficient. For the stated reason, JBC is chosen for the self-pro-
and the hull, simulation time-step must be adjusted to the propeller pulsion test leaving an opportunity to demonstrate the solver
rotation which drastically increases computational time. Max time-step consistency and accuracy on different hull forms and Froude numbers.
size in such cases corresponds to the 2–3 degrees of rotation of the Regarding resistance and wake field on the KCS case, similar geo-
propeller. If for model scale propeller rotation equals to 7–8 rps, global metrical properties of the boundary layer mesh are used as in (Enger
and local time scale can differ from O (10) to O (103) depending on the et al., 2010), but the effect of the overall layer thickness on the solution
solver. At the same time, another problem lies in creating a good quality is observed. In all of the previously mentioned studies, if wake field was
mesh topology in such cases. investigated, results were found sufficiently accurate using k-ω SST
Another approach is to model the flow around the hull and the model with second order discretisation schemes. Some authors (Kra-
propeller inflow as two separate problems, where a different solver can silnikov (Krasilnikov), Enger et al. (2010)) report that most precise
be used for each part of the simulation. As for the coupling procedure, results can be obtained with Reynolds Stress Model (RSM), but at the
Badoe et al. (2014) coupled blade element theory and the momentum cost of larger computation expense and decrease of simulation stability.
theory (BEMt) with RANS calculations in OpenFOAM. Coupling pro- For the current analysis, k-ω SST model is weighted as the best choice.
cedure is made by converting local thrust and torque into momentum Solver used for the study is navalFoam and it is a part of an in-house
sources which are added into the equations until convergence is code library Naval Hydro Pack developed in the foam-extend environ-
reached. Effective wake field is defined by total wake field and then ment. It is a community driven fork of OpenFOAM software for
subtracting the propeller induced velocities calculated from the BEMt Computational Continuum Mechanics using FV discretisation in an ar-
code. On the other hand, Hally (2015), proposed a Boundary Element bitrary polyhedral framework (see Weller et al. (1998)).
Method (BEM) and RANS coupling, also implemented in OpenFOAM. In
order to account for the blade blockage a mass source term is added in
the pressure and the momentum equation, where the effect of blade 2. Mathematical and numerical model
blockage is estimated at about 1% on the resulting thrust force.
Another example of RANS/BEM coupling is presented in Rijpkema This section contains the governing equations for the resistance and
et al. (2013). Computed nominal wake in RANS calculations without self-propulsion flow cases. Detailed mathematical background is given
the propeller in action is used as an input for BEM computation. Pres- for the two-phase incompressible flow followed by the actuator disk
sure loading distribution on propeller is then interpolated as a body (AD) model implemented (Šeb (2017)) in the Naval Hydro Pack. In-
force field in RANS simulation and scaled to provide equal thrust, if terface capturing is accomplished using the Volume-of-Fluid (VOF)
necessary. Total wake field is computed with RANS and the propeller method (Ubbink et al. (Ubbink and Issa, 1999)) while for the handling
induced velocities are subtracted to obtain the effective wake field. The of the free surface discontinuities, the Ghost Fluid Method (GFM)
process is repeated until there is no change in thrust and effective wake (Huang et al. (2007), Desjardin et al. (Desjardins et al., 2008), Lalanne
velocities in the coupling. This approach is very similar to one used in (Lalanne et al., 2015)) is used. To account for the turbulence in the
Krasilnikov (Krasilnikov), where a panel method is used to solve the flow, k-ω SST (Menter (1994)) model with standard wall functions is
pressure distribution on propeller blades followed by the implementa- used, where only the calculation of the main variables is given. Dis-
tion of calculated forces in the actuator disk region in RANS simulation. cretisation methods and coupling procedure for pressure and velocity
The process is repeated until the imbalance between propeller thrust are briefly mentioned. For the rigid body motion and flow-motion
and hull resistance becomes smaller than desired tolerance. coupling the reader is referred to Gatin el al. (Gatin et al., 2017), for
The present paper will introduce a different way of modeling the details.

150
A. Bakica, et al. Ocean Engineering 175 (2019) 149–162

2.1. Governing equations νe = ανe, w + (1 − α ) νe, a, (9)

where νe, w and νe, a are effective kinematic viscosities for water and air,
Mathematical model is based on the following assumptions which
respectively. It is important to notice that smearing of the interface due
are justified for large scale flows in naval hydrodynamics: surface
to usage of GFM does not affect the density field:
tension is considered negligible and tangential stress balance is taken
into account assuming continuous effective viscosity and velocity gra- ρw if α (x) ≥ 0.5
dient. However, the dynamic pressure jump due to the density jump ρ (x) = ⎧ ,
⎩ ρa if α (x) < 0.5
⎨ (10)
and continuity of the velocity field (kinematic boundary condition) is
obtained without simplifications. A brief description of the used turbulent model is needed to be able
Since both fluids are considered incompressible with constant den- to solve the given Reynolds Averaged Navier Stokes (RANS) equations.
sity and continuous velocity field, the continuity equation reads: k − ω SST model is fully described in Menter (1994), so here only the
∇ ·u = 0. (1) procedure to specify the far-field boundary condition for the variables
of turbulent kinetic energy k and specific dissipation rate ω will be
The momentum equation or the Navier-Stokes equation for the in- presented. Value of k is defined as:
compressible two-phase flow with the previously defined assumptions
taken into account is defined as: 3
k= (UI )2
2 (11)
∂u 1
+ ∇ ·(uu) − ∇ ·(νe ∇u) = − ∇p + g , x∈ Ωa ∪ Ω w ∪ Γwa, where U is the free stream velocity and I is the assumed turbulence
∂t ρ (x) (2)
intensity. Used values for turbulent intensity of KCS case is 3% and 2%
where Ω w and Ωa stand for water phase and air phase, respectively. Γwa for JBC simulation. Specific dissipation rate is obtained following Eça
denotes the free surface, x is the position vector, ∇p is the pressure and Hoekstra (2008):
gradient, g is gravitational acceleration and νe is the effective kinematic
U
viscosity.Using the decomposition of the pressure term on dynamic and ω = 10 ,
static component (Rusche (2002)): LPP (12)

p = pd + ρgx, (3) where LPP is ship length between perpendiculars. Next step is to prop-
erly discretise given equations so that the numerical solution adheres to
where pd is the dynamic pressure, yields a different formulation of the the solution of the proposed partial differential equations. For full text
combined two-phase flow in the gravitational field: concerning two-phase incompressible flow coupled with GFM and VOF
∂u 1 see Vukčević et al. (2017) and Vukčević (2016).
+ ∇ ·(uu) − ∇ ·(νe ∇u) = − ∇p .
∂t ρ (x) d (4)
2.2. Numerical model
Terms from left to right are: transient, convection, diffusion and
dynamic pressure gradient term. Free surface Γwa is labeled separately
Discretisation schemes used in this study will be briefly mentioned
due to usage of GFM (Huang et al. (2007)). GFM handling of the free
in this section. For details about FV discretisation of any arbitrary
surface in mathematical formulation is defined in the following equa-
polyhedral Control Volume reader is referred to Jasak (1996). Temporal
tions:
derivative is discretised using first order accurate Euler implicit scheme,
since quasi steady-state solution is sought. Convective term is dis-
- discontinuity of the density field
cretised using the Gauss theorem, while for the interpolation from cell
[ρ] = ρa − ρw , x ∈ Γwa, (5) center to face center linear upwind-biased interpolation is used. Explicit
correction, in order to achieve a second order accurate solution, is made
- continuity of pressure with the gradient extrapolation from an upwind cell. All gradients are
discretised using least square fit (see Jasak and Weller (2000)) with or
[p] = 0, x ∈ Γwa, (6)
without limiters to accomplish a second order accurate solution despite
of mesh skewness. Diffusion term is discretised using the Gauss theorem
- pressure gradient jump condition due to kinematic boundary con-
with linear interpolation using the over relaxed approach for the non-
dition (continuity of the velocity field at the free surface)
orthogonal correction (Jasak, 1996).
⎡ ∇p ⎤ = 0, Following Vukčević (Vukčević et al., 2017) a segregated pressure-
x ∈ Γwa, velocity coupling is accomplished and interface-corrected schemes are

⎣ ρ ⎥⎦ (7)
used in the vicinity of the free surface to account for the discontinuous
where [.] is a jump condition operator as used in Huang et al. (2007), jumps (Eqn. (5), Eqn. (6) and Eqn. (7)). To mark the interface faces,
while ρa and ρw are the air and water densities, respectively. For in- compact polyhedral computational stencil is used. Reader can refer to
terface capturing (VOF) method is used (Ubbink et al. (Ubbink and Issa, Vukčević (Vukčević et al., 2017) for more details regarding interface-
1999)) which is based on an indicator function α that takes the value of corrected approach.
0 or 1 inside the air phase or water phase, respectively. Transport In α transport Eqn. (8) Euler implicit scheme is used for temporal
equation for α reads: derivative, convective term is discretised using the Gauss theorem with
∂α Van Leer's Total Variation Diminishing scheme (Van Leer (van Leer,
+ ∇ ·(uα ) + ∇ ·ur α (1 − α ) = 0, 1977),) with deferred correction (Ferziger and Perić (Ferziger and
∂t (8)
Peric, 1996)) that ensures scalar field boundedness between zero and
where the terms are (from left to right) unsteady, convective and
one. The compressive term is discretised using the Gauss theorem
compressive term. Compressive term (see Rusche (2002) for details)
yielding compressive fluxes which are dependent on the compressive
prevents excessive smearing of the interface while the implementation
velocity ur , which is defined in Jasak et al. (Jasak et al. ) as:
in the current work will be described in more detail in Sec. 2.2. ur
stands for compressive velocity which is only used in the vicinity of the CFLref |df |
ur = cα nΓ ,
free surface due to the multiplier α (1 − α ) .Effective viscosity is as- Δt (13)
sumed continuous across the interface approximating the tangential
where cα denotes a compression constant with default value of one, nΓ
stress balance:
is a unit vector normal to the free surface, CFLref is a reference

151
A. Bakica, et al. Ocean Engineering 175 (2019) 149–162

compression Courant Friefrichs Levy number, df is the distance be-


tween neighboring control volumes and Δt is the time step. Interpola-
tion scheme in the compressive term is specially designed to be upwind
if the interface is sharp, while switching to central differences with
appropriate weighting factors if the interface smearing becomes too
large. For full details regarding the α equation see Rusche (2002) and
for the implementation details see Jasak et al. (Jasak et al. ).Solving
large sparse matrices produced by a FV mesh is achieved using a Krylov
Subspace method with Conjugate Gradient (CG) method and a Cholesky
preconditioner for a symmetric pressure equation, while for asymme-
trical matrices (three components of velocity, indicator function and
turbulence variables) Biconjugate Gradient Stabilized Method (BiCG-
Stab) is used (van der VorstBi-CGStab, 1992).Pressure, velocity and free
surface capturing are coupled with a combination of SIMPLE (Patankar
Fig. 1. KCS - coarse mesh with rudder.
and Spalding (1972)) and PISO algorithm where a PISO pressure-ve-
locity loop is nested inside an outer SIMPLE loop updating the pressure
multiple times per one momentum-equation update (Issa, 1986). Due to Table 1
prior experience with the solver, two outer (SIMPLE) and two inner KCS model particulars.
(PISO) loops are deemed sufficient for a steady state solution. For a Length between perpendiculars LPP (m) 7.2785
complete description of the solution algorithm the reader is referred to Maximum beam of waterline BWL (m) 1.0190
Vukčević (Vukčević et al., 2017).To avoid wave reflection from the Depth D (m) 0.6013
boundaries of the domain, relaxation zones are set up following the Draft d (m) 0.3418
procedure from (Jasak et al., 2015). Wetted surface area w/o rudder Sw / orudder (m2) 9.438
Wetted surface area of rudder Sw / rudder (m2) 0.115
Block coefficient CB 0.6505
2.3. Actuator disk
Model Speed U (m/ s ) 2.19699
Reynolds number Rn 1.4 × 107
Actuator disk model following the Goldstein optimum distribution Froude number Fr 0.26
((Goldstein, 1929)) will be explained in this section (for full details on Propeller diameter DP 0.25
the implementation of the theory in OpenFOAM see Šeb (2017)). A Hub ratio 0.18
mathematical model for an ideal propeller (infinitely thin disc) is based Propeller center, long. location (from FP) x / LPP 0.983
Propeller center, vert. location (below WL) − z / LPP 0.02913
on the modeling of only the two most important aspects of a functional
propeller: the torque and thrust forces. Regarding ship propulsion,
torque - Q and thrust - T are usually given in dimensionless form with
respect to the advance coefficient J which is defined as:
VA
J= ,
nD (14)
where VA is the advance speed, n is the propeller rotation rate and D is
the propeller diameter. Values of T (J ) and Q (J ) are obtained from the
open water test curve and used to determine pressure and tangential
velocity jump for thrust and torque, respectively. Since the advance
speed VA in an experimental set-up is equal to carriage speed for given
thrust, in order to properly compute the inflow velocity on the propeller
plane, correction of the axial velocity is needed. From the actuator disk
theory propeller thrust can be calculated as the mass flux at the pro-
Fig. 2. Convergence of drag coefficient for mesh with 6 layers and 0.20 m
peller plane and the difference in flow velocities (Jasak et al. (Jasak thickness.
et al. )):
T = ρw AD VD (V2 − V1), (15) single time-step in a segregated solution algorithm for pressure-velocity
coupling. Pressure jump is calculated as (Jasak et al. Lalović):
where AD is the actuator disk surface, VD is the axial speed at the pro-
peller plane while V1 and V2 are velocities in front and behind the pro- T (J )
105
peller, respectively. Note that the velocity VD , following the actuator Δp = f (r ),
8π (RP − RH )(3RH + 4RP ) T (18)
disk theory, can be defined as:
VD = 0.5(V2 + V1). (16) where RH is the hub radius, RP is the propeller radius and fT (r ) is de-
fined as (Jasak et al. ):
If V2 is expressed as the unknown and linked with the Eqn. (15), the
inflow velocity V1 = VA can then be computed, after rearranging the
fT (r ) = r * 1 − r * , (19)
terms, as:
T where r * is the normalized disc radius defined as:
VA = VD, CFD − ,
2ρw AD VD, CFD (17)
r ′ − r ′h
r* = ,
where the second term on the r.h.s. accounts for the velocity correction 1 − r ′h (20)
due to the propeller suction on the inflow side while VD, CFD is the
averaged velocity in the propeller axis direction gathered from the AD where r ′h = RH / RP and r ′ = r / RP . The tangential velocity jump that
faces in the computational domain.Non-linearity of the Eqn. (17) for the models the swirl caused by the propeller in action is calculated using
advance speed, since T is a function of J, is easily solved inside of a following expression (Jasak et al. ):

152
A. Bakica, et al. Ocean Engineering 175 (2019) 149–162

Table 2
KCS Resistance results with varying boundary layer.
Boundary layer parameters

Thickness Layer Expansion First layer height CT × 10−3 Error (%) CF × 10−3 Error (%)

20 mm 6 1.5 0.9 mm 3.508 1.38% 2.898 −2.33%


20 mm 5 1.5 1.5 mm 3.492 1.83% 2.899 −2.37%
15 mm 6 1.4 0.9 mm 3.493 1.81% 2.877 −1.60%
15 mm 5 1.4 1.4 mm 3.476 2.28% 2.881 −1.66%
10 mm 5 1.35 0.9 mm 3.398 4.47% 2.797 1.25%
Experiment 3.557 2.832

Fig. 3. KCS - Final mesh geometry.

Fig. 4. Convergence of drag force (left) and wetted area (right) for fine mesh in the KCS case.

105 Q (J ) obvious that the quality and the density of the mesh in the given region
Δut = f (r ),
8π u x (RP − RH )(3RH + 4RP ) Q (21) has a significant role in the definition of the AD. However, when per-
forming self-propulsion simulation mesh is expected to be thoroughly
where u x is the axial speed at the propeller plane and fQ (r ) is defined as
refined in the stern region so the location of the propeller mid-plane is
(Jasak et al. Lalović):
easily satisfied to an acceptable tolerance.
r* 1 − r* Regarding the numerical implementation, imposing a jump condi-
fQ (r ) = ,
r * (1 − r *) + r 'h (22) tions on the chosen set of faces is straightforward and much more
simplified compared to the volumetric source terms in the body force
Note that Δut is imposed tangentially at the propeller plane, in the model which are added to the right hand side of the momentum
circumferential direction. equation. In the following text the procedure of prescribing the pressure
FV framework makes it suitable to easily collect the set of faces on jump is described.
which computed pressure and velocity jumps are imposed. Four para- During the outer SIMPLE and inner PISO loop, boundary conditions
meters are needed to define the AD model: location of the propeller are updated several times in order to ensure the conservation of
plane, direction of the propeller thrust, propeller and hub radius. Using quantities in the domain. In the time instance of each boundary con-
the given parameters AD faces are collected and imposed with a special dition update, pressure jump calculated in Eqn. (18) is added to the
type of boundary condition. Reader should be aware that such current value of the pressure field on the chosen AD faces. Due to the
boundary faces are in fact internal faces of the mesh coinciding with the iterative type of solution algorithm, both the pressure field and the
position of the propeller mid-plane. If there are no faces present at the pressure jump converge to a certain value. Although, if it happens that
exact propeller mid-plane location, the imposed boundary condition the required thrust and computed thrust are not equal, pressure jump is
will be forced upon the next set of closest mesh internal faces. It is

153
A. Bakica, et al. Ocean Engineering 175 (2019) 149–162

Fig. 5. Wave elevation comparison.

Coupling between the body motion and the AD is fully resolved by


recalculating the thrust direction axis and AD faces position after each
mesh motion update.

3. KCS resistance and flow field study

One of the most studied and validated cases in modern naval hy-
drodynamics CFD is the KCS case, featured in both Gothenburg
((Larsson et al., 2013)) and Tokyo Workshops ((Hino, 2005), (Tokyo,
2015 et al., 2015)). A wide range of numerical results are validated
against the experimental data as shown in the introduction, with good
overall agreement.
Simulations around the KCS hull are divided into two sections. The
first section is dedicated to drag coefficient calculation by varying the
Fig. 6. Hull sections on the stern of the KCS case. boundary layer thickness in order to study the effect of the near-wall
layer mesh resolution on drag data and to approximate the optimal
simply scaled to account for the difference. Same procedure is applied boundary layer for the further wake field analysis. The goal is to find
on the velocity in the circumferential direction. Also, it should be noted the smallest possible boundary layer thickness while preserving the
that the AD boundary condition requires as an input a propeller open accuracy of the results in order to have a good quality mesh in the
water test data to compute the thrust and torque during the simulation. propeller region where high boundary layer thickness produces skewed
and non-orthogonal cells thus disrupting the downstream solution.

154
A. Bakica, et al. Ocean Engineering 175 (2019) 149–162

Fig. 7. Wake fraction contours at the propeller plane.

Fig. 8. EFD - wake field on the propeller plane.

155
A. Bakica, et al. Ocean Engineering 175 (2019) 149–162

Fig. 9. CFD - wake field on the propeller plane.

The study is performed on a coarse mesh to save CPU time. Suitable of layers (5 or 6) where the initial boundary layer parameters are taken
boundary layer is then chosen and the mesh is refined for the second from (Enger et al., 2010). Consequently, by changing the boundary
simulation, where the entire flow field is evaluated, giving special at- layer mesh, dimensionless distance y+ on the hull is being altered, but it
tention to detailed wake field analysis. The grid uncertainty is not retains a suitable value for all meshes, ranging from 40 to 80. Generated
evaluated in this study. grid consists of roughly 0.640 × 106 cells for each mesh, where most of
The mesh for all cases spans 2LPP from amidship section to the inlet, the cells are hexahedrons with a maximum non-orthogonality of 7070∘,
2.5LPP to the outlet and 2LPP to side boundaries. Only half of the ship is skewness of 4 and only 10 cells in the z direction near the free surface.
being simulated since the flow problem is symmetric. The mesh height Besides the residuals reaching a satisfactory tolerance and the
from the free surface to the top boundary is 1LPP , while the bottom continuity of mass and volume preserved in the whole domain, force
boundary is set at 1.5LPP below the free surface. acting on the hull is used a convergence criterion. Fig. 2 shows con-
vergence of the force acting on the hull after being converted into the
3.1. Steady resistance - case 2.2a (Gothenburg, 2010) dimensionless drag coefficient for a mesh with six layers and 20 mm
layer thickness. The drag coefficient is decomposed into viscous and
For this case, rudder is added to hull geometry in order to match the pressure components with the viscous component compared to the ITTC
experimental conditions (see Fig. 1). Fluid properties are set up to 1957 correlation line:
correspond the towing tank ((Chalmers University of Technology and 0.075
Gothenburg, 2010)) Reynolds and Froude number of 1.4 × 107 and 0.26, CF = .
(log Re − 2)2 (23)
respectively. Ship particulars are given in Table 1. Boundary layer
disretisation is varied in thickness (10 mm, 15 mm, 20 mm) and number Slight overshoot of the viscous component is considered adequate

156
A. Bakica, et al. Ocean Engineering 175 (2019) 149–162

Fig. 10. Comparison of circumferential wake.

Table 3
Comparison of radial wake distribution.
EFD CFD Error (%)

1 − wEFD 1 − wCFD

0.3 0.38 0.394 −3.68%


0.4 0.457 0.546 −19.47%
0.5 0.586 0.687 −17.24%
0.6 0.717 0.759 −5.86%
0.7 0.791 0.795 −0.51%
0.8 0.823 0.816 0.85%
0.9 0.841 0.828 1.55%
1.0 0.851 0.833 2.12%
1.2 0.852 0.832 2.35%
Nominal wake (from 0.3 to 1.0r / R )
1 − wnEFD 1 − wnCFD
0.753 0.772 2.52%

due to the hull coefficient impact on the viscous part of a few percent.
Results for all grids are presented in Table 2. Experimental friction
Fig. 11. Radial wake distribution.
coefficient CF in the Table 2 is computed from Eqn. (23).
The solution is in reasonable agreement with experimental results,
while the computed error becomes significant when lowering the 3.2. Flow field study - case 2.1 (Gothenburg, 2010)
boundary layer thickness to a value of 10 mm, most probably due to the
boundary layer being too thin to properly evaluate the flow near the Based on the above, the mesh for the flow field analysis is refined to
hull and resolve the velocity gradient. Evaluating the computed results, approximately 2 × 106 cells, also with mostly hexahedron cells and
a five layer mesh with a 15 mm thickness and 1.4 expansion rate is 16 cells in the vertical z direction near the free surface, Fig. 3. Also, the
regarded as sufficient for wake field analysis in the next section. rudder is removed from the mesh to match the experimental condition
(Chalmers University of Technology and Gothenburg, 2010). The jump
in the drag force time signal at around 45 s is visible in Fig. 4 due to
switching off the limiters on the gradient schemes, yielding a second

157
A. Bakica, et al. Ocean Engineering 175 (2019) 149–162

Fig. 12. Velocity components comparison at z / LPP = 0.0302 .

Table 4 surface elevation on the hull is reported together with three long-
JBC model particulars. itudinal cuts on the y plane. Results are also compared in Fig. 5. Cor-
Length between perpendiculars 7.000
relation with the experiment is found to be satisfactory. Loss of accu-
LPP (m)
Maximum beam of waterline BWL (m) 1.125 racy in the results after LPP = 1.5 is mostly due to mesh coarsening,
Depth D (m) 0.625 which is also the cause of slightly inaccurate comparison on the
Draft d (m) 0.4125 y / LPP = 0.4224 cut.
Wetted surface area w/o ESD Sw / oESD (m2) 12.223 Regarding the wake field study, a detailed analysis is performed
Block coefficient CB 0.858 following the ITTC, Practical Guidelines for RANS Calculation of
Metacentric height GM (m) 0.133 Nominal Wake Fields (ITTC, 2014). Data is examined in order to
Moment of Inertia Kxx / B 0.4
evaluate the flow pattern in the propeller location. In Fig. 6, hull sec-
Moment of Inertia Kyy / LPP , Kzz / LPP 0.25
tions and the propeller plane are shown for a more clear visual corre-
Model Speed U (m/ s ) 1.197
Reynolds number Rn 7.46 × 106
lation between the hull shape and the wake fraction contours. Wake
Froude number Fr 0.142 fraction contours are presented in Fig. 7, where the intermediate so-
Propeller diameter DP 0.203 lution with gradient limiters is also shown.
Hub ratio 0.18 Nonetheless, it is clear that the flow is adequately captured with the
Propeller center, long. location (from FP) x / LPP 0.985714
gradient clipping, hence it should always be weighted if the limiters
Propeller center, vert. location (below WL) − z / LPP −0.04042
Vertical center of gravity (from keel) 0.3323
should be removed, since the removal can introduce a high amount of
KG (m)
LCB(% LPP ), fwd+ 2.5475 instability in the simulation in the regions of insufficient mesh resolu-
tion and in the areas of low mesh quality.
From the contour plot in Fig. 7, overestimation of the so-called
order accurate solution. Limiters are used in the beginning in order to hook-like shaped vortices is observed, as well as that the vortex center
stabilize the simulation. Another measure of convergence shown in is moved vertically comparing to the experiment. Due to a fine stern
Fig. 4 is the wetted surface area that stabilizes, proving sharp and stable shape, bilge vortices are not clearly identified because fuller stern
resolution of the interface on the hull. frame-lines usually generate stronger vortices (as noted in Van et al.
Fig. 5 shows the wave elevation compared with the experiment (2006)). However, according to Kim et al. (2001), in the experimental
where accurate correlation between the data is observed. Surprisingly, setup a dummy hub was used to prevent the abrupt change in the flow
even the bow crest is well captured together with a complex stern wave after stern boss. To match the setup more correctly, a hub was placed
field caused by the immersed transom. Small diffusion in the results is inside of the computational mesh and the results are shown in Fig. 7.
most probably due to mesh coarsening in the given area. The graph is Vortex center is moved downwards, but the intensity of the hook-shape
obtained using a contour plot for α = 0.5.Apart from the wave field, free vortex remains overestimated.

Fig. 13. JBC bow (left) and stern mesh with actuator disk (right).

158
A. Bakica, et al. Ocean Engineering 175 (2019) 149–162

Fig. 14. Force convergence on JBC hull.

Fig. 15. Wave elevation on hull.

For further analysis, data is extrapolated for the experimental values Fig. 11. By integrating the circumferential wakes following the proce-
near the center of the propeller plane in order to get a fully resolved dure from ITTC (ITTC, 2011b) a nominal wake can be computed for a
plot of wake fraction contours. The values near the center were out of propeller plane from 30% to 100% of propeller diameter. The com-
the calibration range due to high vorticity in the experiment (Kim et al., parison of nominal wake is also shown in Table 3 with differences in the
2001). Experimental and numerical results are shown in Fig. 8 and inflow velocities on the propeller plane below 3%. The nominal wake in
Fig. 9, with the propeller hub and the tip of the propeller plotted as both cases should be somewhat larger because the flow from the hub to
dashed circles. Radius dimensions are in millimeters, while transverse r / R = 0.3 is ignored.
velocities are drawn every 15∘ degrees in the circumferential direction In the experiment, axial velocities along a vertical cut on the pro-
from the smallest radius of r / R = 0.3 to r / R = 1.2 according to ITTC peller plane were measured with a probe that was set-up along a z-
(ITTC, 2014), where R is the propeller radius. Numerical values with coordinate just below the center of propeller. Values are compared in
the dummy hub were used for comparison in order to be more con- Fig. 12 where good agreement is observed. Large scale aspects of the
sistent with the experimental setup. flow are computed correctly while the small scale flow disturbances are
To make a more detailed investigation of the flow field, each re- not captured.
ferenced radii is compared to experimental values. Circumferential To conclude, acceptable results have been computed at the propeller
mean wake comparison at each propeller radius is shown in Fig. 10. The plane using a k − ω SST model for model scale. However, it should be
error is clearly higher closer to the center of the propeller and the clear that the amount of turbulence present in model scale is not
reason is due to a more complex flow near the propeller hub. To further equivalent to full scale, hence in full scale the wake field is less de-
analyze the results, a mean wake fraction of every reference radii is also pendent on the turbulence model.
presented in Table 3. Errors are below 6% for the range of r / R= be-
tween 0.6 and 1.2, while for the r / R lower than 0.5 the error remains 4. JBC - resistance and self-propulsion
within 20%.
To achieve a more clear view of the comparison, a radial wake In the this section, AD is implemented inside of the computational
distribution shall be plotted together with the experimental results, mesh for JBC hull to evaluate bare hull resistance and capture the

159
A. Bakica, et al. Ocean Engineering 175 (2019) 149–162

Fig. 16. : UX at. x / LPP = 0.9625 Fig. 18. Ux at x / LPP = 0.9625 with AD working.

global features of the flow in self-propulsion condition. Propeller open 4.1. Bare hull
water test data are provided in (Tokyo, 2015) as an input for the AD.
The main dimensions of the JBC hull are presented in Table 4. No First, bare hull results are evaluated where the resistance is mea-
coupling is made with an external (BEM or BEMt) solver. Instead, the sured with the AD turned off. Wave elevation comparison along the hull
flow is resolved inside the FV mesh using the AD model with local flow is shown in Fig. 15, where good agreement between experimental
phenomena (such as blade blockage etc.) being neglected. ((Tokyo, 2015)) and numerical results is observed. In Fig. 16 and
To properly evaluate the asymmetry in the propeller flow due to the Fig. 17 contours of axial velocities are plotted on plane x / LPP = 0.9625
tangential velocity jump related to torque force modeling, full domain where the bilge vortices start to form, and on a plane x / LPP = 0.9843
is simulated. Mesh with 1.012 × 106 cells is used. Note that the mesh just before the propeller plane with hub and propeller radius indicated
models the whole ship instead of half, thus this resolution corresponds by dashed lines. Main features of the flow are accurately captured with
to approximately 0.5 × 106 cells in half of the domain. The hull shape certain amount of diffusion present in numerical results. Intensity of
extracted from the mesh with the AD present is shown in Fig. 13. The bilge vortices is underestimated, with the inflow axial velocities on
mesh is generated in the manner as in the KCS case. The AD is im- plane x / LPP = 0.9843 obviously exhibiting similar behavior as in the
plemented inside of the FV mesh in the last step, after the entire mesh is experiment with certain asymmetry present in the flow. Nonetheless,
completed. Similar approach as in Sec. 3.1 is applied to approximate good overall agreement is obtained.
the optimum boundary layer thickness which resulted in a four
boundary layers with expansion rate of 1.2 and a overall layer thickness 4.2. Self propulsion
of 12 mm, yielding an average y+ value of 40–60 on the hull. Mesh
metrics are similar as in the KCS case. Instead of performing two se- After 60 s of the simulation AD is turned on, measuring the com-
parate simulations for steady resistance and self-propulsion, a single puted resistance for self-propulsion condition. Initial rotation rate on
simulation is performed where the AD is being turned on when the drag AD is set to zero in order to avoid sudden jumps in the flow that may
force for the bare hull converges, thus saving CPU time. Convergence of cause instabilities in the simulation. Average mean velocity on the
the drag force is shown in Fig. 14. propeller plane is calculated from the simulation, while the effective

Fig. 17. Axial velocities x / LPP = 0.9843, dashed lines hub and propeller tip.

160
A. Bakica, et al. Ocean Engineering 175 (2019) 149–162

Table 5
JBC results.
w/o AD with AD

CT × 103 CF × 103 Sinkage (%LPP ) Trim (%LPP ) CSP × 103 KT KQ t

CFD 4.270 3.073 −0.092 −0.197 4.797 0.235 0.0299 0.1091


EFD 4.289 3.159 −0.086 −0.18 4.811 0.217 0.0279 0.1085
Error (%) −0.44 2.71 6.97 9.44 −0.29 8.3 7.16 0.56

wake field is established using an inflow velocity corrected approach step value of approximately 7 × 10−4s . Compared to the current simu-
described in Sec. 2.3. Results from the experiment and the simulation lation the CPU time would be increased by more than ten times at the
are used to compute the experimental and numerical thrust deduction t least. Also, for the real propeller simulation the grid density and size
coefficient to investigate the influence of the suction of the propeller on would have to be significantly increased in the stern region to properly
the hull stern pressure distribution. Thrust deduction is calculated as: evaluate the complex propeller flow and geometry, all of which ad-
Rt , SP − Rt ditionally increase the CPU time.
t= ,
Rt , SP (24)
5. Conclusion
where Rt , SP and Rt are the resistance force at self-propulsion and towing
tank condition, respectively. Propeller rotation rate is varied depending Numerical approach in the FV framework of Naval Hydro has
on the following coefficient (notation used as in (Krasilnikov)): proved accurate for evaluation of resistance and the flow field around
the ship. Ghost Fluid Method handling of interface discontinuities has
Δ= Rt , SP − SFC − T , (25)
produced a very high resolution of the free surface position together
where SFC is the skin friction correction equal to 18.2N from the ex- with the Volume-of-Fluid conservative treatment of the two-phase flow
perimental setup, Rt , SP is the resistance force on the hull and T is the preserving both mass and volume inside of the domain. For both of the
computed thrust on the AD. PI controller is used to control the propeller simulated hull geometries with different Froude numbers, CFD code has
rotation rate, while the intention is to keep the force imbalance Δ demonstrated consistency and reliability.
minimum as possible. Due to the unsteady nature of the flow behind the Regarding the wake field study, detailed qualitative investigation of
hull and varying resistance force in each time step Δ is able to converge the KCS wake field has shown good agreement with the experimental
on a minimum value of about 1% of the resistance force with a rotation results using k − ω SST model, yielding a total mean inflow velocity
rate of 7.4RPS opposed to the experimental 7.8RPS. Mean axial velocity difference of only 2.52%. Closer to the propeller hub complexity and
is integrated from 0.4R to propeller tip due to a loss of data near the hub turbulence of the flow increases with the relative error being under
in the experiment. For the same reason, only the left side of the axial 20%. For r / R ratio above 0.6 correlation between the data is very ac-
velocities in the EFD data is used for calculation. Lower rotation rate in curate. Since the flow in the propeller region is demanding to capture
the CFD simulation is expected due to the lower mean axial velocity on even experimentally, numerical results are found satisfactory.
the propeller plane of 0.445 compared to the 0.452 in the experimental Resistance on both ships (KCS and JBC) has shown good correlation
setup. Since thrust is modeled in terms of advance coefficient J (Eqn. between the numerical and experimental results. Both resistance tests
(14)) any decrease in the inflow velocity must be accounted by the are done on the coarse mesh, with the relative error on JBC case under
decrease in n to compute the same thrust. In Fig. 18 contour plot at 1% and for the KCS case around 1–2.5% depending on the mesh
x / LPP = 0.9625 cut shows increased axial velocities compared with boundary layer. Self-propulsion condition for JBC using an actuator
Fig. 16 where AD is not working. Results show the comparable amount disk computed with sufficient accuracy added amount of resistance to
of numerical smearing as in the bare hull simulation. the hull due to suction from the propeller, yielding a thrust deduction
Complete results for bare hull resistance and self-propulsion simu- coefficient relative error of 0.56%. Such formulation can be used in
lation are presented in Table 5. Experimental CF is computed using Eqn. early steps of the hull design process to improve propeller and hull
(23). Resistance coefficient for both resistance and self-propulsion si- interaction at low CPU cost.
mulation is in a good agreement with the experiment with relative error For future work, a more thorough investigation with multiple mesh
smaller than 1% for both cases. Propeller suction on the hull stern is refinements and inclusion of local flow effects should be considered for
properly modeled which can be seen as the small relative error of the use in the later stages of the ship and propeller design.
thrust deduction coefficient of 0.56%. High relative error of sinkage and
trim are due to small absolute values, which are below one cell height Acknowledgement
resolution in the current mesh. Notable error in computation of thrust
and torque coefficients is present and overestimated for about 8%, This research was supported by the Croatian Science Foundation
while the error is linked to the inflow velocities on the propeller plane under the project Green Modular Passenger Vessel for Mediterranean
being inaccurately computed as seen in Fig. 17. (GRiMM), (Project No. UIP-2017-05-1253).
Results for the entire JBC simulation are found to be satisfactory.
Use of a PI controller and inflow velocity momentum correction proved Appendix A. Supplementary data
to be reliable for simplification of the self-propulsion condition with an
idealized thin disc. In the given setup, AD activation is implemented as Supplementary data to this article can be found online at https://
a run time modifiable variable allowing the user to modify activation doi.org/10.1016/j.oceaneng.2018.12.043.
time while the simulation is running. Regarding CPU efficiency, a
straightforward comparison can be made with the real propeller si- References
mulation. Time step used for the current simulation is equal to
1 × 10−2s . As stated in the introduction in a real propeller simulation a Badoe, C., Phillips, A., Turnock, S.R., 2014. Ship wake field analysis using a coupled
maximum allowable time step is constrained by the approximately 2 BEMt-RANS approach. In: NuTTS ’14: 17th Numerical Towing Tank Symposium.
Chalmers University of Technology, Marstrand, Sweden 6pp.
degrees of propeller rotation. In the current case, propeller rotation rate Banks, J., Phillips, A., Bull, P., Turnock, S., 2010. RANS simulations of the multiphase
equals to 7.8RPS which can be easily used to derive the maximum time

161
A. Bakica, et al. Ocean Engineering 175 (2019) 149–162

flow around the KCS hullform. In: Gothenburg 2010: A Workshop on CFD in Ship 007-7189-5.
Hydrodynamics Gothenburg. Menter, F.R., 1994. Two-equation Eddy-viscosity turbulence models for engineering ap-
Carrica, P.M., Mofidi, A., Martin, E., 2015. Progress toward dircet CFD simulation of plications. AIAA J. 32 (8), 1598–1605.
Manoeuvres in waves. In: Proceedings of the MARINE 2015 Conference, pp. 327–338. Olivieri, A., Pistani, F., Avanaini, A., Stern, F., Penna, R., 2001. Towing Tank Experiments
Chalmers University of Technology, Gothenburg, 2010. A workshop on CFD in ship hy- of Resistance, Sinkage and Trim, Boundary Layer, Wake and Free Surface Flow
drodynamics. 2010. http://www.insean.cnr.it/sites/default/files/gothenburg2010/ Around a Naval Combatant INSEAN 2340 Model, no. IIHR Report No. 421. Iowa
index.html. Institute of Hydraulic Research, The university of Iowa, pp. 56.
Desjardins, O., Moureau, V., Pitsch, H., 2008. An accurate conservative level set/ghost Ozdemir, Y.H., Cosgun, T., Dogrul, A., Barlas, B., 2016. A numerical application to predict
fluid method for simulating turbulent atomization. J. Comput. Phys. 227 (18), the resistance and wave pattern of KRISO Container Ship. Brodogradnja/Shipbuilding
8395–8416. 67 (2), 47–65. https://doi.org/10.21278/brd67204.
Enger, S., Peric, M., Perić, R., 2010. Simulation of Flow Around KCS-Hull. In: Gothenburg Patankar, S.V., Spalding, D.B., 1972. A calculation procedure for heat, mass and mo-
2010: A Workshop on CFD in Ship Hydrodynamics, Gothenburg, Sweden, . http:// mentum transfer in three-dimensional parabolic flows. Int. J. Heat Mass Transf. 15,
www.cd-adapco.com/sites/default/files/conference{_}proceeding/pdf/ 1787–1806.
CDadapcoPaper.pdf. Rijpkema, D., Starke, B., Bosschers, J., May 2013. Numerical simulation of propeller-hull
Eça, L., Hoekstra, M., 2008. The numerical friction line. J. Mar. Sci. Technol. 328–345. interaction and determination of the effective wake field using a hybrid RANS-BEM
https://doi.org/10.1007/s00773-008-0018-1. approach. In: 3rd International Symposium on Marine Propulsors, pp. 421–429.
Ferziger, J.H., Peric, M., 1996. Computational Methods for Fluid Dynamics. Springer. Rusche, H., 2002. Computational Fluid Dynamics of Dispersed Two-Phase Flows at High
Gatin, I., Vukčević, V., Jasak, H., Rusche, H., 2017. Enhanced coupling of solid body Phase Fractions. Ph.D. thesis. Imperial College of Science, Technology & Medicine,
motion and fluid flow in finite volume framework. Ocean Eng. 143 (December 2016), London. https://doi.org/10.1145/1806799.1806850. http://portal.acm.org/
295–304. https://doi.org/10.1016/j.oceaneng.2017.08.009. citation.cfm?doid=1806799.1806850.
Goldstein, S., 1929. On the vortex theory of screw propellers. Proc. R. Soc. Lond. - Ser. A Šeb, B., 2017. Faculty of Mechanical Engineering and Naval Architecture. Master thesis.
Contain. Pap. a Math. Phys. Character 123 (792), 440–465. University of Zagreb, Croatia Numerical characterisation of a ship propeller.
Hally, D., June 2015. Propeller Analysis Using RANS/BEM Coupling Accounting for Blade Shen, Z., Wan, D., Carrica, P.M., 2015. Dynamic overset grids in OpenFOAM with ap-
Blockage. In: 4th International Symposium on Marine Propulsors (smp’15), pp. plication to KCS self–propulsion and maneuvering. Ocean Eng. 108, 287–306.
297–304. https://doi.org/10.1016/j.oceaneng.2015.07.035.
Hino, T. (Ed.), 2005. Proceedings of CFD Workshop Tokyo. NMRI report, pp. 2005. Starke, B., Windt, J., Raven, H.C., September 2006. Validation of viscous flow and wake
Huang, J., Carrica, P.M., Stern, F., 2007. Coupled ghost fluid/two–phase level set method field predictions for ships at full scale. pp. 17–22.
for curvilinear body–fitted grids. Int. J. Numer. Meth. Fluids 44, 867–897. https:// Tokyo, 2015. A Workshop on CFD in Ship Hydrodynamics. http://www.t2015.nmri.go.
doi.org/10.1002/fld.1499. jp/ 2015.
Issa, R.I., 1986. Solution of the implicitly discretised fluid flow equations by operator- Tokyo 2015: A Workshop on CFD in Ship Hydrodynamics, in: L. Larsson, F. Stern, M.
splitting. J. Comput. Phys. 62, 40–65. Visonneau, N. Hirata, T. Hino, J. Kim (Eds.), Tokyo 2015: A Workshop on CFD in Ship
ITTC, 2011a. Recommended Procedures and Guidelines: Practical Guidelines for Ship Hydrodynamics, Vol. vol. 3, NMRI (National Maritime Research Institute), Tokyo,
CFD. In: 26th International Towing Tank Conference, pp. 1–18. Japan, 2015.
ITTC, 2011b. Recommended Procedures and Guidelines - Nominal wake measurement by Ubbink, O., Issa, R.I., 1999. A method for capturing sharp fluid interfaces on arbitrary
a 5-Hole pitot tube, vol. 7. pp. 11 5-02-03-02.4 (Revision 01). meshes. J. Comput. Phys. 153, 26–50.
ITTC, 2014. Recommended Procedures and Guidelines - Practical guidelines for RANS van der Vorst, H.A., 1992. Bi-CGStab: a fast and smoothly converging variant of Bi-CG for
calculation of nominal wakes. 7.5-03-03-02 (Revision 00). pp. 9. the solution of nonsymmetric linear systems. SIAM J. Sci. Comput. 13, 631–644.
Jasak, H., 1996. Error Analysis and Estimation for the Finite Volume Method with Van, S.H., Kim, W.J., Yim, G.T., D. H, K., Lee, C.J., 1998. Experimental Investigation of
Applications to Fluid Flows. Ph.D. thesis. Imperial College of Science, Technology & the Flow Characteristics Around Practical Hull Forms. In: Proceedings of the 3rd
Medicine, London. Osaka Colloguium on Advanced CFD Applications to Ship Flow and Hull Form
Jasak, H., Weller, H.G., 2000. Application of the finite volume method and unstructured Design.
meshes to linear elasticity. Int. J. Numer. Methods Eng. 48, 267–287. Van, S.H., Kim, W.J., Yoon, H.S., Lee, Y.Y., Park, I.R., 2006. Flow measurement around a
Jasak, H., Vukčević, V., Gatin, I., 2015. Numerical simulation of wave loads on static model ship with propeller and rudder. Exp. Fluid 533–545. https://doi.org/10.1007/
offshore structures. In: CFD for Wind and Tidal Offshore Turbines. Springer Tracts in s00348-005-0093-6.
Mechanical Engineering, pp. 95–105. van Leer, B., 1977. Towards the ultimate conservative difference scheme. {IV}. {A} new
H. Jasak, V. Vukčević, I. Gatin, I. Lalović, CFD Validation and Grid Sensitivity Studies of approach to numerical convection. J. Comput. Phys. 23, 276–299.
Full Scale Ship Self Propulsion, International Journal of Naval Architecture and Vukčević, V., 2016. Numerical Modelling of Coupled Potential and Viscous Flow for
Marine Engineering (in press). Marine Applications. Ph.D. thesis. Faculty of Mechanical Engineering and Naval
Jones, W., Launder, B.E., 1972. The prediction of laminarization with a two-equation Architecture, University of Zagreb.
model of turbulence. Int. J. Heat Mass Transf. 15, 301–314. Vukčević, V., Jasak, H., Gatin, I., 2017. Implementation of the Ghost Fluid Method for
Kim, W.J., Van, S.H., Kim, D.H., 2001. Measurement of flows around modern commercial free surface flows in polyhedral Finite Volume framework. Comput. Fluids 153, 1–19.
ship models. Exp. Fluid 31 (5), 567–578. https://doi.org/10.1007/s003480100332. Weller, H.G., Tabor, G., Jasak, H., Fureby, C., 1998. A tensorial approach to computa-
V. I. Krasilnikov, Self-Propulsion RANS Computations with a Single-Screw Container Ship, tional continuum mechanics using object oriented techniques. Comput. Phys. 12,
in: Third International Symposium on Marine Propulsors smp’ 13, Launceston, 620–631.
Tasmania, Australia. Yang, C., Lu, H.D., Löhner, R., 2010. On the simulation of highly nonlinear wave-
Lalanne, B., Villegas, L.R., Tanguy, S., Risso, F., 2015. On the computation of viscous breakwater interactions. J. Hydrodyn. 22 (5 Suppl. 1), 932–938. https://doi.org/10.
terms for incompressible two-phase flows with Level Set/Ghost Fluid Method. J. 1016/S1001-6058(10)60055-8. https://doi.org/10.1016/S1001-6058(10)60055-8.
Comput. Phys. 301, 289–307. yu Guo, C., cheng Wu, T., Zhang, Q., Gong, J., 2016. Numerical simulation and experi-
Larsson, L., Stern, F., Visonneau, M., 2013. Numerical Ship Hydrodynamics: An assess- mental research on wake field of ships under off-design conditions. China Ocean Eng.
ment of the Gothenburg 2010 workshop. Springerhttps://doi.org/10.1007/978-94- 30 (5), 821–834. https://doi.org/10.1007/s13344-016-0053-3.

162

You might also like