Sie sind auf Seite 1von 10

Estimation of the maneuvering characteristics of the DTC containership

using URANS based simulations


N. Fournarakis, A. Papanikolaou, D. Chroni, S. Liu, T. Plessas
Ship Design Laboratory, National Technical University of Athens, Greece

ABSTRACT: The objective of the presented research is to explore the capability of CFD techniques and of
numerical methods in general, in the estimation of the hydrodynamic derivatives and ultimately in the simulation of the maneuvering performance. In this context, the DTC standard containership bare hull is subjected to
resistance, transverse force and yaw moment CFD calculations at various Froude Numbers and headings. Unsteady Reynolds Averaged Navier Stokes (URANS) computations were performed using the STAR-CCM+
code. The resulting hydrodynamic derivatives are validated against results of other approaches, such as semi
empirical methods as well as the NTUASDL's panel codes (NEWDRIFT, HYBRID) and the respective results
are discussed. Resulting hydrodynamic derivatives are then used to predict the maneuvering performance of the
hull, using the NTUASDL's HYBRIDMAN, in calm water and in adverse weather conditions.

1 INTRODUCTION
Nowadays, maneuvering simulations are used to
estimate and assess the operational capabilities of a
ship during the design phase. Zigzag and turning circle tests are, amongst others, standard IMO tests in
order to identify the compliance of a ship design with
the related design requirements. As the respective
process suggest, the maneuvering performance of the
ship for a given set of environmental loads and effects, is dependent on the design characteristics of the
ship. Using a solid mathematical modeling technique
(Hirano, 1980), simulators solve the equations of ship
motion, using the data set of forces and moments acting on the ship during its course. The key advantage
of this process is the strength of the simulator to estimate the ship performance in any arbitrary motion of
the ship. However, the accurate prediction of forces
and moments as source data is fundamental for the
development of a simulation model for ship maneuvering.
Traditionally, standard Planar Motion Mechanism
(PMM) tests, Circular Motion Tests (CMT) and other
tank tests provide the necessary data, yet in a costly
and time consuming manner. Experiments on a
Froudescaled hull model, enable a deeper insight
into the hydrodynamic effects on models maneuverability. Unfortunately, such tests are performed much

later in the design process, leading to long design cycles, costly prototype development and late estimation of maneuvering capabilities.
Recent advances in the field of numerical hydrodynamics, like (Stern, et al., 2011), (Simonsen et al.,
2012), suggest the use of Computational Fluid Mechanics as a way to assess the hydrodynamic forces
and moments used for the maneuvering simulation of
a hull. The evolution of computational power complemented by the development of sophisticated Finite
Volume Method codes, provide promising insights
for the hydrodynamic pressure and viscous phenomena occurring during a maneuvering action. Much of
the work performed so far is focused on simulating
the various PMM conditions through the modelscale
replication of basin tests like static drift, static rudder,
pure sway or pure yaw. Their aim is to identify the
actual hydrodynamic forces and moments, which are
ultimately used to feed a mathematical maneuvering
model based on the MMG standard. Latest efforts
(Yasukawa et al., 2015) pursue the replication of the
actual maneuvering motions, simulating zigzag, circle tests, etc.
The milestone work of (Abkowitz, 1964) facilitated
the development of several mathematical formulations in order to estimate the forces acting on a ship
during a maneuvering action, like Hirano et al. (1980)
with a 3DOF and Son and Nomoto (1981) with a 4
DOF equation model. Recently, within the context of

SIMMAN 2008 (Stern et al., 2011), the first work- 2.2 Maneuvering problem
shop on verification and validation of ship maneuverIn the present study, the ships motion is restricted to
ing simulation methods was contacted, presenting inthree degrees of freedom (surge, sway and yaw).
sights for the maneuvering abilities of the KVLCC1
Herein, heel motion is disregarded; although it may
and KVLCC2 standard tankers, as well as the KCS
has effects during maneuvering. For this case the
standard containership.
equations of motion are as follows:
Recent advances in the field suggested by Skejic
and Faltinsen (2008) in their review paper of the subject area, provide methods to incorporate second order wave forces and moments in the maneuvering
equations as well as a two time scale model to separate low frequency (maneuvering) from high frequency (seakeeping) motion components. Similarly,
Yasukawas (2006) two time scale model used a 6
DOF motion formulation, coupling seakeeping motion with the second order mean drift forces, computed by a momentum conservation far field approach.
In the present study, a 4DOF maneuvering
model, mathematically formulated and solved within
MATLABs Simulink environment, developed at
Figure 1: The coordinate systems
Ship Design Laboratory of National Technical University of Athens, is used (Chroni D. et al., 2015). The
nonlinear maneuvering equations are formulated by
(( ) 2 + (2 + )
applying the Newtonian laws. Forces and moments
= + + + () + +
acting on the hull are computed using computational
(( ) + + (( 2 + 2 ) ))
fluid dynamics. Moreover, the resulted forces and
= + + +
moments during a maneuver induced by the propeller

( + )

and rudder are taken into account, as well as, environ=


+ + +
mental forces, such as wind and waves. The maneu2
( + 2 ) + ( 2 ) +
vering motion results are validated against experi ( + ) + ( ) = + + +
mental data of the DTC hull, in calm water as well as

(1)
under the influence of winds and waves.
2 FORMULATION OF THE MANEUVERING
PROBLEM
2.1 Coordinate system
Two types of coordinate systems will be used: Fixed
systems (relative to earth) and moving systems. As
shown in Figure 1, the earth fixed, right handed coordinate system O(i,j,k) with the kaxis pointing downwards, is used for the identification of the position
and orientation of the vessel, during a maneuver. The
body fixed o(x,y,z), advances with the ships forward
speed V and rotates with rotational speed r and it is
used for the calculation of the forces and moments
which are acting on the ship during a maneuver. Another, earth fixed coordinate system is defined in order to express the existence of mean second order
wave forces, which is right handed with the kaxis
pointing upwards. Finally, as shown in Figure 1, is
the rudder angle (negative for rudder to starboard)
and , and are ships heading, drift and incident
wave angles respectively.

where, m is the ship mass, and are the moments


of inertia about x and z axes, and are the center
of gravity coordinates with respect to the body fixed
coordinate system (i.e. CG = [ ,0, ]). , v and r are
the time varying accelerations which are defined with
respect to the body fixed coordinate system. Their integration in time, leads to u, v and r velocity components and their double time integration expresses
ships position in the earth fixed coordinate system.
X,Y,K,N represent the surge, sway, heel and yaw directional components, respectively, and subscript H,
R, prop and e indicates forces and moments due to
x ,
hull, rudder, propeller and wind. In addition,R

R y and
Mz are the mean second order wave forces
and moments.
The hydrodynamic forces and moments acting on the
hull are modeled as nonlinear functions of the accelerations, the velocities and the Euler angles which can
be expressed in a series expansion of coefficients,
called hydrodynamic derivatives. Hull forces for each
degree of freedom in HYBRIDMAN inhouse code
are expressed as follows:

= + + 2 + 3 + + 2
+ 3
= + + 2 + 3 + + 2
+ 3
= + + 2 + 3 + + 2 +
3
(2)

In the present study, the hydrodynamic derivatives


were determined from CFD computations, as will be
presented in section 3.
2.3 Propulsion and Resistance
There are several references (Lewis, 1989) which
provide good overviews on hull resistance forces. In
case of absence of resistance test data of the examined
vessel, the water resistance can be estimated as follows:
= 0.5 2
(3)
Where , , are the sea water density, the wetted
surface area of the hull and the ship forward speed
respectively. is the resistance coefficient.
However, in the present study, resistance is predicted
from CFD calculations for the subjected vessel at various Froude numbers. Then, the non dimensional resistance force as a function of Froude number is estimated as a third order polynomial function of ships
speed in x direction (body fixed system). Finally, the
resistance is imported in HYBRID MANs code, as
follows (in surge hull forces component):
= + 2 + 3

(4)

Propeller force is obtained by using the following


equation:
= 2 4 ()

(5)

where Z, , n, D, () are the number of propellers,


the propeller revolution per second, the propeller diameter and thrust coefficient, respectively. Thrust coefficient is a function of the advance coefficient J and
it is dependent on the propeller design and Reynolds
number Rn. For, the approximation of () the Wagenigen B series were used (Lammeren et al., 1969)
(Oosterveld, 1975).
The initial value of the propeller revolutions
should be adjusted so that the desired ship velocity is
obtained for the condition of still water and constant
forward speed (ships speed at the beginning of the
maneuver).
2.4 Rudder force
The rudder force and moment acting on the ship hull
by rudder action is calculated as follows (Lewis
1989):
= (1 )
= (1 + )

=
= ( + )

(6)

where , and denote the rudder angle and the x


and z directional center of normal force acting on it.
is the ratio of the hydrodynamic force induced on
the ship hull by the rudder action and can be estimated
by the following empirical formula:
= 0.62( 0.6) + 0.227
(7)
is the normal rudder force, which is estimated
as
6.13

= 0.5 2 +2.25

(8)

where and are the rudder area and the aspect


ratio of the rudder respectively. stands for the
speed and the angle of the effective inflow into the
rudder. The effective rudder inflow speed can be
calculated as follows:
= (1 )[1 + 2 ()]0.5
(9)
Where 2 = 1.065 for the port rudder and 0.935 for
the starboard rudder.
Finally, which is the effective inflow angle of the
rudder, is calculated as follows:
= +

(10)

2.5 Wind force


The wind force and moment acting on the ship hull is
estimated as follows:
2
= 0.5
2
= 0.5
2
= 0.5 ( 2 /)
2
= 0.5

(11)

where = ( + )2 + ( + )2 is the resultant airflow velocity felt by the ship. The dimensionless , , coefficients are functions
of ships water profile and of relative wind angle. In
HYBRIDMAN code, these coefficients are obtained
from published, model experimental data for the specific vessel type, which are given in tabulated form,
as a function of the relative wind angle, as presented
by Blendermann (2001).
2.6 Calculation of mean second order wave forces
The second order mean forces are calculated by either
near field method (Papanikolaou et al., 1987) or far
field method (Liu et al., 2011) using the first order
quantities calculated with potential flow method (Papanikolaou et al., 1990). Noting that in relatively

short waves (/LPP<0.5), it is necessary to consider


some complications of the ensuing physical problem
of the added resistance (Liu et al. 2011),
(Papanikolaou et al., 2015), (Papanikolaou et al.,
1990). In the Figures 2, 3 and 4 the profiles of the
mean second order wave forces are presented
(/Lbp=0.5).

Figure 4: Mean second order yaw moment /Lbp=0.5

3 HYDRODYNAMIC FORCES AND MOMENTS


COMPUTATION USING CFD

Figure 2. Mean second order surge force /Lbp=0.5

The CFD computations presented herein are supposed to replace tank test conditions used traditionally to estimate the required bare hull forces and moments used in the maneuvering simulation method.
3.1 Governing equations
The governing equations for the case studied are
the Navier Stokes equations for an incompressible
and laminar flow of a Newtonian fluid. Using a compact form and Einsteins notation the equations are
written:

Figure 3: Mean second order sway force /Lbp=0.5

+ = + +

(12)

where ui represent the components of the velocity


vector, p is the fluid pressure and ij the components
of the viscous stress tensor. In order to simplify the
solution of the system of equations, by removing all
fluctuations arising from turbulence, an averaging
process at a time scale larger than the largest scale of
turbulence, is followed. Practically, the velocity components may be expressed as the sum of a mean term
and an instantaneous deviation term and then,

the system of equations rewrites in the following


form:

( )

+ +

1
(

(13)

Equation 13 is the basis for Reynolds Averaged


Navier Stokes (RANS) computational methods,
where turbulence fluctuations are removed below a
certain turbulence length scale. However, due to the
final term of Equation 13 which represent the so

called Reynolds stresses, the introduction of a turbulence model is required in order to resolve the correwhere ui are the components of the velocity field and
lation symmetric tensor arising from the velocity flucc is the volume fraction of water in the cell which vartuation components.
ies from 0 to 1, full of air to full of water, respectively.
Boussinesqs eddy viscosity assumption stated that
the Reynolds stresses may be approximated from the
rate of strain tensor Sij, as in the case of viscous 3.2 Computational method used
stresses, by substituting molecular viscosity with its
The CFD computations performed in the context
turbulence equivalent, turbulent viscosity T. Rewritof the present research used the Reynolds Averaged
ing the final term of equation 13, using the Boussinesqs assumption, we have:
2

= 2 = 3

(14)

where T is the turbulent viscosity, Sij the fluid strain


tensor and k the turbulence kinetic energy per unit
mass.
Amongst the various models to calculate the value
of turbulent viscosity, in the frame of the current research, the two equation NASAs realizable k turbulence model is used (Shih et al., 1997), which is
based on Jones and Launder (1972) standard k
model. According to the model, the eddy viscosity is:
=

(15)

where C is a function of mean strain and rotation


rates whereas is the kinetic energy rate of dissipation.
Complementing the RANS equations, the fluid
mass conservation is described by the continuity
equation:

=0

(16)

Typically, in order to resolve the free surface phenomena, either interface tracking or interface capturing techniques may be used. Each method chosen has
certain advantages, however in the present research
the Volume of Fluid (VOF) interface capturing tech-

Figure 5: The computational domain

+ ( ) = 0

Figure 6: The DTC hull mesh

NavierStokes (RANS) solver STAR-CCM+ from


CDAdapco (CDAdapco). The code solves the
RANS and continuity integral equations on an unstructured mesh using the Finite Volume technique.
For the case studied, even though it could theoretically be solved using the steady RANS equations and
modeling, the Volume Of Fluid technique implemented by STAR-CCM+ is far more robust when
solving the unsteady equations for such a problem
type. Thus, the Unsteady RANS equations have been
used herein, where the timestep, iterative convergence and overall simulation time have been selected
to achieve a converged steady solution. A second order Euler difference temporal discretization scheme
was used. Spatial discretization was performed using
second order schemes for both convective and viscous terms. The SIMPLE method was used by the
solver, coupling pressures and fluid velocities, while
the RANS stress tensor closure was achieved using a
realizable k turbulence model, using an all Y+ wall
treatment that is capable to treat the flow near the wall
depending on the Y+ value achieved by the respective
mesh and flow conditions.
In order to analyze the resulting forces and moments, two coordinate systems were used: an earth
fixed inertial reference coordinate system and a COG
ship fixed coordinate system. All ship motions and
hull forces and moments are calculated with reference
to the ship fixed frame. For the case studied, the ship
is considered free to move in heave and pitch, while
all other ship motion DOFs are constrained.
3.3 Computational domain

nique is used. According to the method, the fraction


of water and air is calculated in each volume cell,
solving the following equation for the entire grid:

Figure 2: The DTC hull mesh

(17)

The computational domain presented in Figure 5


was developed using the STAR-CCM+ meshing
tools. The mesh is dominated by hexahedral cells,
while near the hull, the cells are polyhedral as a result
of the trimming to follow the hull lines. In order to

resolve the boundary layer flow (Simonsen et al.,


2012) and the near field viscous phenomena, a number of prism layers are attached on the hull wall,
where a noslip condition is applied.
Typical boundary conditions for the subject case
are used. The inlet boundary is located two ship
lengths in front of the ship. A Dirichlet condition is
used at the inlet of the fluid flow, describing the inlet
velocity as well as the volume fractions of the corresponding fluid phases.
The sides are located two ship lengths from the
hull. On the sides, Neumman symmetry conditions
are used, ensuring the field gradient continuity on the
sides.
The outlet is located three ship lengths from the
hull. A Dirichlet pressure condition is used at the outlet, ensuring zero gradients of velocity and volume
fraction of fluids, as well as the continuity of hydrostatic pressure.
Bottom and top sides are located one and half a
ship length from the hull, respectively. Top and bottom sides are considered as inlets with similar boundary conditions.
The resulting hull mesh is presented in Figure 6,

each studied parameter are required and therefore,


three different grid sizes have been used, namely a
695K cell coarse (No. 3), a 1.2M medium (No. 2) and
a 2.5M fine grid (No. 1) to provide three solution sets.
However, a systematic grid refinement is not possible
when using an unstructured grid, since the refinement
areas are not explicitly controlled. In order to overcome this challenge, a grid base size has been used

Figure 8: Wave contours ate various drift angles, at 20 kts

and all grid areas have been defined according to this


size.
Table 2: PMM static drift nondim forces and moments

Figure 7: The near field mesh

while the near field meshing is presented in Figure 7.


The standard DTC hull was modeled using a scale
1:80 and thus, its main particulars are presented in Table 1.
Table 1: DTC model particulars
Lbp
Lwl
Bwl
T

5.577m
5.684m
0.801
0.228m

3.4 Verification and validation of results


CFD analysis is a laborious task, yet often the process results to either poor stability or poor accuracy.
Both require care and attention in order to exploit the
respective results. When it comes to verification, grid
sensitivity, i.e. the perturbation of the results when assuming a slight change on grid sizing, is of major importance.
According to ITTC (2002), an uncertainty analysis
is required to verify and validate CFD results. In order
to perform such an analysis, three distinct values for

Drift
angle
0
5
10
15
20
Drift
angle
0
5
10
15
20
Drift
angle
0
5
10
15
20

v'
0,000
0,087
0,174
0,259
0,342
v
0,000
0,087
0,174
0,259
0,342
v'
0,000
0,087
0,174
0,259
0,342

20 kts
-6,14E-04
-6,51E-04
-6,56E-04
-8,49E-04
-9,99E-04
20 kts
0,00E+00
4,75E-04
1,46E-03
2,87E-03
4,73E-03
20 kts
0,00E+00
3,67E-04
7,25E-04
1,20E-03
1,75E-03

X'
16 kts
-6,39E-04
-6,94E-04
-6,71E-04
-7,78E-04
-9,30E-04
Y'
16 kts
0,00E+00
5,36E-04
1,54E-03
2,87E-03
4,60E-03
N'
16 kts
0,00E+00
3,70E-04
7,03E-04
1,14E-03
1,65E-03

6 kts
-6,62E-04
-6,95E-04
-7,20E-04
-7,58E-04
-8,14E-04
6 kts
0,00E+00
5,38E-04
1,50E-03
2,15E-03
4,08E-03
6 kts
0,00E+00
3,54E-04
6,79E-04
1,07E-03
1,33E-03

Refinement is performed by limiting the grid base


size using a refinement factor of rG2=2, as suggested
by ITTC (2002). For the different grid sizes, in order

From the results obtain, a good agreement between


the CFD and the EFD data from MARINTEK
(Sprenger F. et al., 2015) is observed, especially at the
6,00E-03

Y'

to subsequently estimate the hydrodynamic derivatives, the longitudinal Xforce, the transverse Y
force and the yaw moment N around the vertical axis
Z with reference at midship, were computed.
The results from the CFD PMM static drift analysis are presented in Table 2, while for the CMT tests
are presented in Table 3. Forces and moments are
nondimensionalized with the appropriate factors for
the forces and moments respectively
PMM static drift analysis was performed for three
different ship speeds, namely 20 kts, 16 kts and 6 kts,
in order to evaluate how the viscous phenomena affect the calculated hydrodynamic derivative values.
Drift angle ranged from 0 to 20 degrees. Furthermore,
CMT analysis was performed at two different ship
speeds, namely 15.5kts and 6kts with a rotating radius
varying from 20 to 50 meters at model scale.

6kts
16kts

4,00E-03

20kts
2,00E-03

-0,400

-0,300

-0,200

Exp

0,00E+00
-0,100
0,000

v'
0,100

0,200

0,300

0,400

-2,00E-03

-4,00E-03

-6,00E-03

Figure 10: The non-dimensional Y force vs the non-dimensional vertical ship speed

Table 3: CMT nondim forces and moments


R

15.5 kts
4,52E-04
7,24E-04
1,09E-03
1,49E-03

6kts speed. Increased speed leads to improved nondimensional forces and moments at higher drift
angles, a matter that may be attributed to the
contribution of free surface phenomena as well as the
viscous phenomena observed at higher Froude
numbers.

N'
6 kts
6,81E-05
7,81E-05
9,93E-05
1,59E-04

15.5 kts
7,28E-04
1,15E-03
1,63E-03
2,76E-03

6 kts
1,27E-04
1,62E-04
2,27E-04
3,78E-04

2,00E-03

N'

50
40
30
20

Y'

Following the ITTC process, a parameter and iterative convergence study is required, with systematic
refinement ceteris paribus. Such an analysis was performed and the respective results validated the accuracy of the CFD findings. However, the further analysis of this process is neglected within the context of
this research.
Production runs performed, identified the values
for the nondimensional forces and moments required as inputs for the simulation models. At the following figures, a comparison is made between the
CFD calculated data from PMM static drift analyses,
at the various drift angles and rotating radius, respec-0,400

-0,300

-0,200

-0,100
0,000
0,000E+00

0,100

0,200

0,300

0,400

v'
-2,000E-04

20kts
16kts

-4,000E-04

6kts
Exp

-6,000E-04

-8,000E-04

X'

-1,000E-03

Figure 9: The non dimensional X force vs the non-dimensional vertical ship speed

tively, for the case of 6 kts.

6kts

1,50E-03

16kts

1,00E-03

20kts
5,00E-04

Exp
-0,400

-0,300

-0,200

0,00E+00
-0,100
0,000

0,100

0,200

0,300

0,400

-5,00E-04
-1,00E-03
-1,50E-03

v'

-2,00E-03

Figure 11: The non-dimensional N moment vs the vertical


ship speed

In order to estimate the ship maneuvering behavior,


the mathematical model used requires the bare hull
hydrodynamic derivatives. Using the data from non
dimensional forces and moments coming from the
numerical simulations, the hydrodynamic derivatives
of the hull are calculated. These data are assumed to
replace the experimental data that could be measured
during a full set of costly towing tank campaigns. The
predicted hydrodynamic derivatives are presented in
Table 4. However, only the static drift and the rotating
arm tank tests were simulated with the CFD solver
and thus, the hydrodynamic derivatives set does not
inlcude the coupling terms (Y'rv, Y'vr, etc.). Therefore,
the simulated maneuvering motion is performed
without the above coupling terms, assuming that their
effect is herein limited; it will be elaborated in
planned future CFD investigations of the authors.

5 NUMERICAL RESULTS AND DISCUSSION


Table 4: The DTC bare hull hydrodynamic derivatives
(x105)
X'u
X'uu
X'uuu

610
-1740
1060

Y'u
Y'uu
Y'uuu

X'v
X'vv
X'vvv

-50
150
400

Y'v
Y'vv
Y'vvv

X'r
X'rr
X'rrr

Y'r
Y'rr
Y'rrr

N'u
N'uu
N'uuu
-970
3290
-11400
-90
340
-830

N'v
N'vv
N'vvv

350
580
-1310

N'r
N'rr
N'rrr

170
1340
-2390

4 IMPLEMENTED CODE
The above outlined method has been practically implemented by coupling NTUA-SDL's 3D seakeeping
codes (NEWDRIFT, HYBRID) with a newly developed code to simulate the maneuverability of ships,
i.e. HYBRID MAN.
The new code has been developed in MATLAB's
Simulink environment. As shown in Figure 12 the
maneuvering module enables the time domain simulation of different maneuvering scenarios, as speci-

The DTC standard container hull (El Moctar et al,


2012) has been selected as a validation example of the
maneuvering numerical simulations. The principal
particulars of the vessel are presented in the table 5
below:
Table 5: Principal particulars of the DTC hull
Model scale =63.65
L

5,684

ship length [m]

0,801

ship breadth [m]

0,228

mean draft [m]

Cb

0,661

block coefficient

672,7

ship mass [kg]

Sw

5,534

wetted surface [m2]

Propeller information
D

0,14

propeller diameter [m]

number of blades

P/D

0,959

Propeller pitch ratio

Ae/Ao

0,8

Port and starboard turning circle tests were performed


in calm water, in waves and with wind.
The tested conditions, such as vessel's speed, wind direction and speed, and the wave characteristics are
presented at the beginning of each case study. The numerical results are compared with experimental data
which were conducted in MARINTEK (Sprenger F.,
2015).

Figure 12: Flow chart of the maneuvering simulation procedure

fied by IMO, taking in addition into consideration the


influence of external forces due to wind and waves.
Each examined hull is subjected to virtual tank tests
by using Star-CCM+ commercial code. From the resulted forces and moments acting on the hull, the
nondimensional hydrodynamic coefficients are calFigure 13: Calculated turning circle trajectory compared
with experimental data
culated in order to be imported in HYBRID MAN
maneuvering code. Hydrodynamic components of the
hull, i.e. added masses, as well as the mean second
order wave loads are being calculated by the seakeep- 5.1 Turning circle test in calm water
ing module. For each time step, depending of the vesIn the turning circle test, the initial ship velocity is
sel position and orientation with respect to the earth
equal to 6 knots (full scale) and the rudder angle is set
fixed coordinate system, the mean second order wave
to 35 degrees after the end of a phase in procedure.
forces and moments are calculated by interpolation
The helm rate of the rudder is 2.25o/s in full scale. The
from precalculated response surfaces.

propeller revolutions were initially set to attain ships


speed in calm water.
In Figure 13, the calculated trajectory results are
compared with model experimental data disposed by
MARINTEK in the frame of the SHOPERA project
(Sprenger F. et al., 2012). Moreover, in Figure 14 the
turning circle trajectories, both for both port and starboard rudder angles (35deg), are presented.
5.2 Turning circle test in waves
In this test, DTC hull is subjected to the turning circle
maneuver in the presence of waves. The initial ship
velocity is again equal to 6 knots (full scale) and the
rudder angle is set to -35 degrees after a phase in procedure. The propeller revolutions are initially set up
to achieve the initial ship speed. The incident waves
are regular, with a wave height of 2m, and 12,5sec

6 CONCLUSIONS
The conducted research study and the presented numerical simulation results for the DTC standard containership have shown that, the employed theoretical
and numerical approaches to the determination of the
maneuvering equation components are satisfactory
and the overall agreement of the obtained theoretical/numerical results with corresponding experimental data very good, even though some coupling
terms in the equations of motion were not included in
the hydrodynamic derivatives set. The same conclusion is generally valid for the simulated turning circle
maneuvering trajectories, even though the effect of
waves on the trajectories is less satisfactorily captured
with increasing simulation time. The reason for this
deviation is being investigated and will be elaborated
in future publications of the herein presented research
work.
7 AKNOWLEGEMENTS

Figure 14. Port and starboard turning circles for DTC hull,
calculated via HYBRID MAN.

period. The initial encounter angle is 0 (head waves).


A comparison between calculated and experimental data is presented in Figure 15 below.

The work presented in this paper is supported by the


Collaborative Project (Grant Agreement number
605221) SHOPERA (Energy Efficient Safe SHip OPERAtion) cofunded by the Research DG of the European Commission within the RTD activities of the
FP7 Thematic Priority Transport / FP7-SST-2013RTD-1/ Activity 7.2.4 Improving Safety and Security
/ SST.2013.4-1: Ships in operation. The European
Community and the authors shall not in any way be
liable or responsible for the use of any knowledge, information or data of the present paper, or of the consequences thereof. The views expressed in this paper
are those of the authors and do not necessary reflect
the views and policies of the European Community.
8 REFERENCES

Figure 15: Calculated turning circle trajectory in waves,


compared with experimental data

Abkowitz, M. (1964). Lectures on Ship Hydrodynamics


Steering and Manoeuvrability, Report No HY-5,
Hydro-Og Laboratorium, Lyngby, Denmark.
Blendermann, W. (2001). Probabilistic and spectral modelling
of the wind loads on ships. Technische Univ. Hamburg,
Harburg.
Carlton, J. (2007). Marine Propellers and propulsion. USA:
Elsevier.
CDAdapco (n.d.). Retrieved from Star CCM+: www.cd
adapco.com
Chroni, D., Liu, S., Plessas, T., & Papanikolaou, A. (2014). Ship
maneuvering in waves: Background and validation of
simulation software (HYBRID_MAN). Technical
Report of Ship Design Laboratory, National Technical
University of Athens.
Chroni, D., Liu, S., Plessas, T., Papanikolaou, A. (2015).
Simulation of the maneuvering behavior of ships under
the influence of enviromental forces, IMAM 2015,
Croatia.

Hirano, M. (1980). On the calculation method of ship


maneuvering motion at initial design phase (in
Japanese). Journal of the Society of Naval Architects
of Japan 59, pp. 7181.
International Towing Tank Conference (ITTC). (2002).
Uncertainty Analysis in CFD: Verification and
Validation Methodology and Procedures. Quality
Manual 7.5-03-01-01.
International Towing Tank Conference (ITTC) (2011).
Guideline on Use of RANS tools for Manoeuvring
Prediction, Recommended Procedures and Guidelines
Manual, 7.5-03-04-01.
Jones, W., & Launder, B. (1972). The prediction of
laminarization with a twoequation model of
turbulence. Journal of Heat and Mass Transfer, pp.
301304.
Lammeren, A., Mannen, & Oosterveld. (1969). The Wagenigen
B-screw series. Monograph, SNAME.
Lewis, E. V. (1989). Principles of Naval Arctitecture Vol III:
Motion in waves and controllability. Jersey City, NJ:
SNAME.
Liu S., Papanikolaou, A., & Zaraphonitis, G. (2011). Prediction
of added resistance of ships in waves. Ocean
Engineering, pp. 641650.
Moctar, O., Shigunov, V., & Zorn, T. (2012). Duisburg Test
Case: Post Panamax Container Ship for
Benchmarknig. Ship Technology Research Vol.5 No.3.
Oosterveld, O. (1975). Further Computer Analysed Data of the
Wagenigen B-srew Series. ISP.
Papanikolaou, A., & Zaraphonitis, G. (2015). Practical
approach to the added resistance of a ship in short
waves. Proceedings of the 25th International Offshore
and Polar Engineering Conference KONA- USA.
Papanikolaou, A., Zaraphonitis, G., & Schellin, T. (1990). On a
3D method for the evaluation of motions and loads of
ships with forward speed in waves. Proceedings of 5th
International Congress on Marine Technology.
Shih, T., Zhu, J., & Lumley, J. (1997). A realizable Reynolds
Stress Algebraic Equation Model. NASA Lewis
Research Center.
Simonsen, C., Otzen, J., Klimt, C., & Larsen, N. (2012, August
2631). Maneuvering predictions on the early design
phase using CFD generated PMM data. 29th
Symposium on Naval Hydrodynamics.
Skejic, R., & Faltinsen, O. (2008, August). A unified seakeeping
and maneuvering analysis of ships in regular waves.
Journal of Marine Science and Technology, pp. 371
394.
Son, K., & Nomoto (Sprenger, Maron, Delefotrtie, &
Hochbaum, 2015), K. (1981). On the coupled motion of
steering and rolling of a high-speed container ship.
Journal of Society of Naval Architects 150.
Sprenger, F., Maron, A., Delefotrtie, G., & Hochbaum, A.
(2015). Mid Term Review of Tank Test Results.
SHOPERA (Grant Agreement number 605221).
Stern, F., Agdrup, K., Kim, S., Hochbaum, A., Rhee, K.,
Quadvlieg, F., Gorski, J. (2011, June). Experience from
SIMMAN 2008- The First Workshop on Verification
and Validation of Ship Maneuvering Simulation
Methods. Journal of Ship Research, pp. 135147.
Yasukawa, H. (2006). Simulations of ship maneuvering in
waves. Journal of the Japan Society of Naval Architects
and Ocean Engineers.
Yasukawa, H., & Yoshimura, Y. (2015). Introduction of MMG
standard method for ship maneuvering predictions. J
Mar Sci Technol.

Das könnte Ihnen auch gefallen