You are on page 1of 6

NUMERICAL SIMULATIONS OF SHIP SELF-PROPULSION AND MANEUVERING

USING DYNAMIC OVERSET GRIDS IN OPENFOAM


Zhirong Shen (American Bureau of Shipping, USA)
Richard Korpus (American Bureau of Shipping, USA)

1.

SUMMARY

Dynamic overset grids provide a powerful tool for


CFD calculations that model relative motion between
body components. This paper presents a new overset
method utilizing OpenFOAM to simulate the selfpropulsion and maneuvering of ships. The method
uses one set of overset cells fixed to the Earth, a
second set fixed to the hull, a third set rotating with
the propellers, and a fourth set rotating with the
rudders. Prescribed motions move the propeller and
rudder blocks relative to the hull, and the entire hull
assemblage moves relative to Earth using 6-DOF
motions derived from CFD forces. A digital autopilot
controls rudder rotation for course keeping. Overset
boundary conditions and hole cutting are updated
dynamically at run time for every time step.
The method is demonstrated using two ship models
for which extensive validation data exists: the Japan
Bulk Carrier (JBC); and the Office of Naval Research
Tumblehome ship (ONRT). JBC self-propulsion
results include cases with and without an Energy
Saving Device (ESD). ONRT results include free
running maneuvers in regular head seas and
quartering waves.
2.

INTRODUCTION

Predictions of ship self-propulsion, maneuvering and


course keeping, whether in calm water or waves, are
one of the most demanding challenges in ship
hydrodynamics. Interactions between the hull, rudder,
and propeller all have to be accurately resolved, and
this proves impossible for potential flow codes and
other linear methods. Computational Fluid Dynamic
(CFD) is one of the most effective alternatives
because it resolves the complex viscous turbulent
flow around a stern using the Naiver-Stokes
equations. However, relative motion between the
various components (propellers, rudders, heaving
hulls) remains a critical challenge because traditional
dynamic grids distort (causing a loss of accuracy),
and sliding mesh methods lack robustness (e.g.
rudders moving close to rotating propellers).

The dynamic overset grid technique has a long


history of accurate, efficient and robust applications
(Rogers, et al., 1994, Korpus, et al., 1997, 2005, 2007,
Meakin, 1999, Suhs, et al., 2002). Individual
geometry elements are usually moved using a nested
hierarchy approach wherein any group of objects
moves relative to its parent using prescribed motions,
free dynamics, or some combination of the two.
Propellers, for example, can rotate around one
ship-fixed axis while rudders rotate about a different
one. The whole collection of elements can either be
made to translate at a known speed (such as during a
ship resistance test), or under the influence of forces
from waves, propellers, or towing forces. The type of
movement can even change part way through a
simulation as might be required for a stores
separation problem. The potential permutations are
endless.
More recent developments with dynamic overset
grids include demonstrations of hull-propeller-rudder
interactions and zig-zag maneuvering for the KCS
CFD validation model (Mofidi and Carrica 2014),
Carrica et al. (2015) extended that work to include
the ability to perform maneuvering simulations in
waves, and demonstrated a great step forward in
computational ship hydrodynamics. Shen et al.
(2014a, 2014b, 2015) implemented overset grid
technique in the open source toolkit OpenFOAM by
coupling with the SUGGAR library (Noack, 2005).
The resulting dynamic overset grid technique was
validated for self-propulsion, seakeeping and
maneuvering using the KCS, KVLCC2 and DTMB
5415M CFD validation data sets.
The present paper introduces a new implementation
of dynamic overset gridding for OpenFOAM solvers.
Domain Connectivity Information (DCI), hole cutting,
and overlap minimization are provided using a
standalone overset grid assembler. Dynamic linking
of the grid assembler and OpenFOAM solvers allows
computations of complex ship hydrodynamic
problems including self-propulsion and free-running
maneuvering simulations.

III - 221

Two categories of simulations are performed to


demonstrate the method. This first category covers
self-propulsion simulations of the Japan Bulk Carrier
(JBC) with and without an Energy Saving Device
(ESD). The second category covers course-keeping
simulation of the ONR Tumblehome ship (ONRT) in
waves. The ONRT model is equipped with twin
propellers, twin rudders, and a digital autopilot for
course keeping.
3.

COMPUTATIONAL METHODS

All simulation results presented in this paper were


created using OpenFOAM version 2.3.1. Minor
modifications to the standard interDyMFoam solver
are necessary to enable inclusion of dynamic overset
grids. InterDyMFoam discretizes the incompressible
Reynolds-Averaged Navier-Stokes (RANS) equations
using a finite volume technique, and captures the free
surface using the Volume of Fluid (VOF) method to
resolve interfaces between air and water. The
SST model provides turbulence closure.

The method uses a stand-alone DCI grid assembler to


establish connectivity among independent moving
overset grids. Coding uses the new C++11 standard
and dynamic library calls OpenFOAM to fetch and
return grid information for fast DCI searches and
boundary condition specifications. The software
currently handles up to two-levels of grid block
hierarchy, which is sufficient for the self-propulsion
and maneuvering simulations presented herein. An
overlap minimization procedure reduces duplicated
grid regions to the smallest allowable for a given
order of accuracy, and therefore improves
interpolation accuracy. The assembler is capable of
processing OpenFOAM grids with arbitrary cell
shape, including general polyhedron, and therefore
allows grid importation from a wide range of mesh
generation tools. Additions are under development to
improve stability and run time efficiency.
4.

4.2 Overset Grids


The computational grids consist of several overset
component cell groups. Each component grid is
created using HEXPRESS from NUMECA, and then
assembled using stand-along DCI software. Figure 1
illustrates an assembled grid for the case with ESD
and propeller. The grid is composed of separate
components for the hull (including ESD), propeller
and background each shown in a different color.
Tables 2 and 3 demonstrate the results of grid sizes
from a grid convergence study performed for JBC.
Results for three grid levels are shown to demonstrate
how a refinement ratio of 2 affects accuracy. Total
cell counts range from O(1.0e+6) to O(5.7e+6) for
the conditions without propeller (Cases 1.1 and 1.2),
and from O(1.5e+6) to O(7.4e+6) for the conditions
with propeller (Cases 1.5 and 1.6).
Simulating the four separate cases demonstrates one
advantage of the overset technique. Changing
between cases only requires swapping out one or two
component grids. Switching from Case 1.5 to Case
1.6, for example, requires changing only the hull grid.
Reusing the propeller and background grids (as
indicated in Table 3) saves immeasurable time.

(a) Surface Grid of ESD and Propeller

JAPAN BULK CARRIER (JBC)

4.1 Test Conditions


JBC is a new self-propulsion benchmark model for
the Tokyo 2015 Workshop. The model has a length of
7.00 meters and test speed of 1.179 meters per second
(corresponding to Fr = 0.142). Test conditions
include self-propelled cases with and without a
duct-like ESD mounted upstream of the propeller.
Table 1 summarizes the test conditions. The test
series also includes resistance measurements made in
calm water with the propeller removed.
Corresponding CFD simulations include both the
resistance and self-propelled configurations, each
performed with and without the ESD. In all cases, the
model is free to sink and trim and the computational
domain moves forward at ship speed.

(b) Global View

(c) Close Up of Stern Region


Fig. 1 Overset Grids of JBC, ESD and Propeller

III - 222

Table 1 Summary of JBC Test conditions


Case ID Case 1.1 Case 1.2
ESD
No
Yes
Propeller No
No

Case 1.5 Case 1.6


No
Yes
Yes
Yes

target, the propeller RPS defines the self-propulsion


point. In order to capture blade-rate force fluctuations
accurately, the time step for self-propulsion
simulations is set to 0.0006 seconds (~ 200 time steps
per propeller revolution).

Table 2 Overset Grids for JBC without Propeller


Grid
Coarse
Medium
Fine
Coarse
Medium
Fine

Hull
Background
Without ESD (Case 1.1)
461,728
562,500
1,097,591
1,208,380
2,663,514
2,818,520
With ESD (Case 1.2)
500,747
562,500
1,142,667
1,208,380
2,908,232
2,818,520

Total
1,024,228
2,305,971
5,482,034
1,063,247
2,351,047
5,726,752

Table 3 Overset Grids for JBC with Propeller


Grids

Hull
Propeller Background
Without ESD (Case 1.5)
Coarse
608,906 349,517
562,500
Medium 1,219,223 843,611 1,208,380
Fine
2,673,358 1,879,165 2,818,520
With ESD (Case 1.6)
Coarse
773,460 349,517
562,500
Medium 1,383,455 843,611 1,208,380
Fine
2,754,647 1,879,165 2,818,520

Total
1,520,923
3,271,214
7,371,043
1,685,477
3,435,446
7,452,332

4.3 Results
Table 4 shows resistance predictions for JBC without
ESD (Case 1.1). All quantities come from after the
solution converges to a steady state solution. Errors
of CT for medium and fine girds are less than 1%.
Table 5 shows resistance predictions with the ESD
(Case 1.2), and demonstrates that CT, CP and CV show
monotonic convergence. Predictions of sinkage and
trim are diverged, but are notably small values at this
low Froude number. They will have limited impact on
resistance. The medium grid obtains good resistance
predictions with less than 1% relative error for CT.
Fine grid results are only marginally better.
For the self-propulsion tests, a PI controller adjusts
propeller speed during the simulations in order to
achieve target speed and thrust/drag balance. The
controller monitors the difference between ship speed
and target speed and updates propeller rotation rate
accordingly. Figure 2 demonstrates the method for
the self-propelled JBC with ESD (Case 1.6), and
shows how ship and propeller speed eventually reach
a steady state. The initial condition is with the ship at
target speed (1.179 m/s) and propeller at zero
revolutions per second (RPS). Note that ship speed
first drops due to insufficient propeller rotation. After
a few seconds, the thrust catches up and ship speed
approaches its target. Once ship speed reaches its

Fig. 2 Time History of Propeller RPS and Ship Speed


for Self-Propelled JBC with ESD (Case 1.6, Medium
Grid)
Table 4 Resistance of JBC without ESD (Case 1.1)
CT (x103)
Error
CP (x103)
CV (x103)
Sinkage [%LPP]
Error
Trim [%LPP]
Error

EFD
4.29

Fine Medium Coarse


4.2621 4.2928 4.6197
-0.65% 0.064% 7.684%
1.1195 1.1851 1.4489
3.1426 3.1077 3.1708
-0.086 -0.0873 -0.0882 -0.0895
1.547% 2.588% 4.069%
-0.18 -0.1915 -0.1925 -0.1936
6.405%
6.94% 7.573%

Table 5 Resistance of JBC with ESD (Case 1.2)


CT (x103)
Error
CP (x103)
CV (x103)
Sinkage [%LPP]
Error
Trim [%LPP]
Error

EFD
4.26

Fine Medium Coarse


4.275
4.287
4.479
0.351% 0.637% 5.146%
1.175
1.184
1.346
3.1
3.103
3.134
-0.085 -0.0871
-0.088 -0.0888
2.513% 3.582% 4.507%
-0.182 -0.195
-0.194 -0.191
7.027% 6.344% 5.173%

Table 6 Self-Propulsion of JBC without ESD


(Case 1.5)
CT (x103)
Error
n (RPS)
Error
KT
Error
KQ
Error

EFD
4.81
7.8
0.217
0.0279

Fine
Medium
4.768
4.866
-0.88%
1.17%
7.842
7.962
0.54%
2.07%
0.2125
0.2138
-2.06%
-1.48%
0.02836 0.02835
1.68%
1.60%

Coarse
5.039
4.76%
8.194
5.05%
0.2148
-0.99%
0.02825
1.24%

Table 6 summarizes results of self-propulsion tests


for the JBC without ESD (Case 1.5). CT, propeller
speed (n) and KQ all show monotonic convergence.
Although KT fails to converge, the maximum error is
less than 2.1%. As for the total resistance and
propeller speed, both medium and fine grids obtain

III - 223

better results than the coarse grid. Table 7 lists the


self-propulsion results with ESD (Case 1.6), and
shows a similar trend as Case 1.5. Both total
resistance and propeller speed are well predicted by
the medium and fine grid. KT shows oscillatory
convergence and KQ is diverged, indicating that it is
relatively more difficult to achieve convergence for
propeller forces and moments than for hull resistance.

global view of the complete grid, and close ups of the


relevant details. Note that a different background grid
in the shape of a cylindrical is used in Case 3.13 for
quartering wave simulations

Table 7 Self-propulsion of JBC with ESD (Case 1.6)


CT (x103)
Error
n (RPS)
Error
KT
Error
KQ
Error
5.

EFD
4.762
7.5
0.233
0.0295

Fine
Medium
4.786
4.752
0.50%
-0.22%
7.673
7.632
2.30%
1.75%
0.2266
0.2256
-2.75%
-3.18%
0.02966 0.02933
0.54%
-0.56%

Coarse
5.024
5.50%
7.96
6.10%
0.2286
-1.89%
0.02952
0.07%

(a) Side View

ONR TUMBLEHOME (ONRT)

5.1 Test Conditions

(b) Hull Surface Grid

Similar to the JBC, ONRT is a new benchmark ship


model for Tokyo 2015. The model has a length of
3.147 meters and speed of 1.11 meters per second
(corresponding to Fr = 0.20). ONRT has twin rudders
and twin propellers. Table 8 lists the three
free-running maneuvering cases simulated using the
overset system described herein. The comparable
model tests tow a model at a fixed attitude for a
prescribed period and then release it from the carriage
into six degree-of-freedom motion controlled only by
the rudder autopilot. Rudders are activated by a PID
controller for course keeping. For Cases 3.12 and
3.13, the model is sailing in regular waves of length
3.147 meters and wave height of 0.06294 meters.
Table 8 Free-Running Test Conditions for ONRT
Case No. Case 3.9
Case 3.12
Condition Calm water Head wave
Motion
6DoF
Propellers Yes
Rudders Active

6DoF
Yes
Active

Case 3.13
Quartering
wave (135o)
6DoF
Yes
Active

5.2 Overset Grids


The computational grid for ONRT consists of six
overset cell groupings as listed in Table 9. In addition
to the background grid, the moving components
include the hull, two rudders and two propellers.
Each overset grouping of cells can move relative to
all the other groupings and is generated using
HEXPRESS. The components are prepared using the
described overset grid assembler. Figure 3 shows a

(c) Overset Component Grids at Stern Region


Fig. 3 Overset Grids for ONRT
Table 9 Summary of Overset Grids for ONRT
Grid Name
Hull
Propeller x 2
Rudder x 2
Background
Total

Case 3.9 & 3.12


1,911,782
1,178,626
424,951
1,208,380
6,327,316

Case 3.13
1,876,649
1,178,626
424,951
1,383,300
6,467,103

5.3 Results
Before any free-running simulations start, we
initialize the calculation by making self-propelled
simulations to determine the propeller speed needed
for thrust balance at a ship speed of 1.11 m/s. The
simulation follows the same procedure as described
in Section 4.3. For ONRT the final predicted
propeller speed is 8.80 RPS, or 1.9% less than the
model-test value of 8.97 RPS.
For free-running simulation in calm water (Case 3.9),
the ship model has full 6-DOF motion and a constant
propeller speed of 8.80 RPS. Table 10 shows the

III - 224

results, and indicates a non-dimensional final speed


of 1.001 (1.27% less than the experimental value of
1.01). Predicted sinkage is 2.89% larger than the
experimental result and trim angle has a slightly
worse error at 6.6%. As with JBC though, the
absolute differences in value are small.
Table 10 Free-Running Test of ONRT in Calm Water
Case 3.9
Speed
Sinkage 102(m)
Trim (deg)
n (RPS)

EFD
1.01
0.226
-0.0386
8.97

CFD
1.001
0.2327
-0.0411
8.8

Error
-1.27 %
2.89%
6.63%
-1.90%

For Case 3.12, the model is sailing in head waves and


Figure 4 shows the time history of ship motions. The
solid lines represent CFD predictions and the circles
experimental measurements. The predicted heave and
pitch motions match well with measurements. The
speed loss is due to added resistance induced by the
incident waves, and matches well given the limited
data acquisition rate. Figures 5 and 6 show four
instantaneous snapshots of free-surface motion and
vortices near the stern (evenly spaced over one
encounter period). Waves breaking at the bow and
violent motions at the stern are both apparent in
Figure 5. Figure 6 demonstrates the complex
interactions between propeller vortices, rudders, ship
motions and incoming waves.

Fig. 6 Four Snapshots of Vortical Structures near the


Stern during One Encounter Period

Fig. 7 Wave Generating Zone for Quartering Waves


For Case 3.13, the ONRT model sails in oblique
quartering regular waves. OpenFOAMs third party
library, waves2Foam (Jacobsen et al., 2012), creates
waves in the far field region and applies a relaxation
zone to transmit them into the near field. A relaxation
method blends the far field analytic solution with the
near field computed one. Figure 7 demonstrates the
wave generation region and computational domain.

Fig. 4 Time History of Ship Motions (CFD: solid


line; Experiment: circle)

Fig. 5 Four Snapshots of Free-Surface and Ship


Motion for One Encounter Period

Fig. 8 Time History of Ship Motion and Rudder


(CFD: solid line; Experiment: circle)

III - 225

REFERENCES
Carrica, P.M., Mofidi, A., Martin, E., (2015),
Progress toward direct CFD simulation of
manoeuvres in waves, Proceedings of MARINE
2015, Rome, Italy.

Fig. 9 Free-Surface and Ship Motion at the Moment


of Maximum Roll Angle

Jacobsen, N.G., Fuhrman, D.R., Fredse, J., (2012),


A wave generation toolbox for the open-source CFD
library: OpenFoam, International Journal for
Numerical Methods in Fluids, Volume 70, No 9, pp
10731088.
Korpus, R., Hubbard, B., Jones, P., Stromgren, C.,
and Bennett, J., (1998), Hydrodynamic Design of
Integrated
Propulsor/Stern
Concepts
by
Reynolds-Averaged
Navier-Stokes
Technique,
Proceedings of the 7th PRADS, The Hague,
Netherlands.
Korpus, R., and Liapis, S., (2005), Active and
Passive Control of SPAR Platform Vortex-Induced
Motions, Proceedings of OMAE 2005, Halkidiki,
Greece.

Fig. 10 Vortical Structures near the Stern: Vorticity


Represented by ISO-Surface of Q.
Figure 8 shows time histories of ship motion and
rudder angle resulting from the simulation. The heave
and pitch motions are in good agreement with
measurements. The amplitude of roll motion is
over-predicted, which may be due to an inaccurate
vertical center of gravity. The angle of yaw deviation
reaches a maximum value right after model release,
but the deviation angle returns to an average of zero
(with oscillations) due to the countering effects of
waves and autopilot. Since the rudder angle is
dependent on the yaw deviation angle, the curve
presents a similar trend as that of the yaw angle.
Figure 9 presents the evolution of free surface and
ship motion during one encounter wave period. The
ship is undergoing large-amplitude roll motions.
Figure 10 shows a close-up view of the vortical
structures in the stern region. The vortices are
represented using an ISO-surface of Q and colored by
the axial velocity. The axial velocity is projected onto
the x-axis of ship coordinate system.
6.

CONCLUSIONS

This paper presents numerical simulations of two new


benchmark ships using an overset grid technique
developed
for
OpenFOAM.
Self-propulsion
simulations of the JBC model indicate total resistance
and propulsion points are in good agreement with
experimental results. Free-running course keeping
simulations with the ONRT model in both calm water
and regular waves reveal a good match with model
measurements. The comparisons demonstrate that
dynamic overset grids work well with OpenFOAM
and greatly simplify marine and offshore simulations.

Korpus, R., (2007), Performance Prediction without


Empiricism: A RANS-Based VPP and Design
Optimization Capability, Proceedings of the 18th
CSYS, Annapolis, Maryland, USA.
Meakin, Robert L., (1999), Composite Overset
Structured Grids, Chapter 11, Handbook of Grid
Generation, CRC Press.
Mofidi, A., Carrica, P.M., (2014), Simulations of
zigzag maneuvers for a container ship with moving
rudder and propeller, Computers & Fluids, Volume
96, pp 191203.
Noack, R.W., (2005), SUGGAR: a general
capability for moving body overset grid assembly,
AIAA Paper 2005-5117.
Rogers, S. E. and Pulliam, T. H., (1994), Accuracy
Enhancements for Overset Grids Using a Defect
Correction Approach, AIAA Paper 94-0523.
Shen, Z., Carrica, P.M., Wan, D., (2014a), Ship
motions of KCS in head waves with rotating
propeller using overset grid method, Proceedings of
OMAE 2014, San Francisco, USA.
Shen, Z., Wan, D., Carrica, P.M., (2014b), RANS
Simulations of Free Maneuvers with Moving Rudders
and Propellers Using Overset Grids in OpenFOAM,
Proceedings of SIMMAN 2014, Lyngby, Denmark.
Shen, Z., Wan, D., Carrica, P.M., (2015), Dynamic
Overset Grids in OpenFOAM with Application to
KCS Self-Propulsion and Maneuvering, Ocean
Engineering, Volume 108, pp 287-306.
Suhs, Norman E. and Rogers, Stuart E. and Dietz, W.
E., (2002), PEGASUS 5: An Automatic
Pre-Processor for Overset-Grid CFD, AIAA Paper
2002-3186.

III - 226

You might also like