Вы находитесь на странице: 1из 26

Proceedings of the 11th International Conference on

Hydrodynamics (ICHD 2014)


October 19 – 24, 2014
Singapore

RECENT PROGRESS IN CFD FOR NAVAL ARCHITECTURE AND OCEAN


ENGINEERING (KEYNOTE SPEAKER)

FREDERICK STERN, ZHAOYUAN WANG, JIANMING YANG, HAMID SADAT-HOSSEINI, MAYSAM


MOUSAVIRAAD, SHANTI BHUSHAN*, MATTEO DIEZ†, SUNG-HWAN YOON, PING-CHEN WU, SEONG MO
YEON, TIMUR DOGAN, DONG-HWAN KIM, SILVIA VOLPI, MICHAEL CONGER, THAD MICHAEL, TAO XING‡,
ROBERT S. THODAL§ AND JOACHIM L. GRENESTEDT**
IIHR-Hydroscience and Engineering, University of Iowa, Iowa City, IA, USA

An overview is provided of CFDShip-Iowa modeling, numerical methods and HPC, including both current V4.5 & V5.5 and
next generation V6. Examples for naval architecture highlight capability and needs. High fidelity V6 simulations for ocean
engineering and fundamental physics describe increased resolution for analysis of physics of fluids. Uncertainty quantification
research is overviewed as first step towards development stochastic optimization.

1. Introduction
CFD capabilities continue to advance at ever-faster speed and ever-more-impressive accomplishments, as
recently reviewed for ship hydrodynamics by Stern et al. (2013). None-the-less CFD is slow in its adaptation by
industry since most users are at universities and R&D laboratories (ITTC, 2011). However, slowly-but-surely
CFD is transforming engineering design as the build-and-test design spiral approach transforms to the SBD
approach offering innovative out-of-the-box 21st century design concepts with improved safety, energy and
economy. First generation SBD capability has focused more on functionality than high fidelity and exascale
computing requiring significant advancements to achieve the next generation SBD capability for fully resolved,
fully coupled, sharp-interface, multi-scale, multi-phase, multi-disciplinary turbulent ship flow including fluid
structure interactions and utilizing billions of grid points.
Herein, recent progress in CFD for naval architecture and ocean engineering is overviewed based
specifically on CFDShip-Iowa URANS/DES toolbox, as an example of the current state-of-the-art. The
emphasis is on the latest research since Stern et al. (2013). For a more complete list of references with regard to
the development and applications of CFDShip-Iowa URANS/DES toolbox within the field of computational
ship hydrodynamics, the readers are referred to Stern et al. (2013). Iowa science and technology paradigm for
the development of the SBD capability is described. An overview is provided of CFDShip-Iowa modeling,
numerical methods and HPC, including both current V4.5 & V5.5 and next generation V6. Examples for naval
architecture highlight capability and needs. High fidelity V6 simulations for ocean engineering and fundamental
physics describe increased resolution for analysis of physics of fluids. Uncertainty quantification research is
overviewed as first step towards development stochastic optimization. Recent progress deterministic and
stochastic optimization research is not reviewed herein since recently provided by Campana (2013).

2. Paradigm for Development SBD for Ship Hydrodynamics


Rapid advancements in simulation technology are revolutionizing engineering practice, as SBD and ultimately
virtual reality are replacing current reliance on experimental observations and analytical methods. It is expected
that a major shift in how scientific method forms its basis of conceptual truth, a shift from reliance on
observations, based on experiments, to reliance on logic, based on simulations supported by experiments. SBD

*
Currently at Center for Advanced Vehicular Systems, Mississippi State University, Starkville, MS 39759

Also affiliated CNR-INSEAN, Rome, Italy

Currently at Department of Mechanical Engineering, University of Idaho
§
Lehigh University, Bethlehem, PA, 18015
**
Lehigh University, Bethlehem, PA, 18015
1
Copyright © 2014 by ICHD
covers a broad range from computerized systems based methods to solutions of physics based initial boundary
value problems (IBVP). Present interest is in solutions of physics based IBVP for ship hydrodynamics. SBD for
ship hydrodynamics merges traditional fields of resistance and propulsion, seakeeping, maneuvering, open-
ocean and littoral environmental effects, and offers new opportunities for future ships to meet challenges of the
21st century. Development SBD involves new paradigm for hydrodynamics research in which CFD,
experimental fluid dynamics (EFD), and uncertainty analysis (UA) are conducted simultaneously for benchmark
geometries and conditions using an integrated approach along with optimization methods, all of which serve as
internal engine guaranteeing simulation fidelity. International collaborations with other research institutions and
organizations include participation in ITTC and NATO AVT working groups and naval engineering educational
consortium (NEEC), organizing international CFD workshops and current NICOP projects. Those activities are
mutual-beneficial and magnifying individual institute capabilities, which has been foundational in the
unprecedented achievements of computational ship hydrodynamics.

3. CFDShip-Iowa URANS/DES/LES SBD Toolbox


CFDShip-Iowa is general-purpose CFD simulation software developed at the University of Iowa’s IIHR—
Hydroscience & Engineering for support of student thesis and project research as well as transition to Navy
laboratories, industry, and other universities. CFDShip-Iowa has been a leading ship hydrodynamics CFD code
for over 20 years, which has been verified and validated for many applications in ship flows. The current
versions include CFDShip-Iowa V4.5, V6.1, and V6.2, with V5.5 and V6.3 under development.

3.1. V4.5 & 5.5 modeling, numerical methods, HPC


CFDShip-Iowa V4.5 is an incompressible URANS/DES solver designed for ship hydrodynamics (Huang et al.,
2008). The equations are solved in either absolute or relative inertial non-orthogonal curvilinear coordinate
system for arbitrary moving but non-deforming control volumes. Turbulence models include blended k–/k–
based isotropic and ASM (Algebraic Stress Model) based anisotropic RANS, and DES approaches with near-
wall models or wall functions. A single-phase level-set method is used for free-surface capturing. Captive, semi-
captive, and full 6DOF capabilities for multi-objects with parent/child hierarchy are available. The fully
discretized propeller or body-force propeller model can be employed for propulsion. The water-jet propulsion
can be included using actual water-jet with detailed simulation of the duct flow or water-jet model with the
reaction forces and moments. Incompressibility is enforced by a strong pressure/velocity coupling, achieved
using either PISO or projection algorithms. The fluid flow equations are solved in an earth-fixed inertial
reference system, while the rigid body equations are solved in the ship system. Other modeling capabilities
include semi-coupled two phase air/water modeling, environmental waves and winds, bubbly flow, and fluid-
structure interaction.
Numerical methods include finite difference discretization on body-fitted curvilinear grids, with high order
upwind schemes for the convection terms and second-order centered for the viscous terms. The temporal terms
are discretized using a second-order backward difference Euler scheme. Since the solver is designed for high-
Reynolds number flows, the transport and re-initialization equations are weakly elliptical and thus pentadiagonal
line solvers in an alternate-direction-implicit (ADI) scheme are used. A MPI-based domain decomposition
approach is used, where each decomposed block is mapped to one processor. The resulting algebraic equation is
solved with the PETSc toolkit using block Jacobi incomplete factorization (ILU) pre-conditioners and bi-
conjugate gradients stabilized (BCGSL). All equations of motion are solved in a sequential form and iterated to
achieve convergence within each time step.
Extension of CFDShip-Iowa Version 4.5 to Version 5.5 with a fully coupled two-phase flow solver using
the Volume-of-Fluid (VOF) method is in progress. The approach includes implementing the highly accurate
geometric VOF interface tracking method developed for V6, developing fully-coupled two-phase flow solver,
implementing cavitation and mixture models for air/water/vapor three-phase interaction, and developing
capabilities for the necessary applications. The numerical methods, HPC, and SBD functional areas are similar
to Version 4.5.

2
Copyright © 2014 by ICHD
3.2. V6.1, 6.2 & 6.3 modeling, numerical methods, HPC
The next-generation high-fidelity SBD tools, CFDShip-Iowa V6, are already under development for milestone
achievement in increased capability focusing on orders of magnitude improvements in accuracy, robustness, and
exascale HPC capability.
In Version 6.1, Cartesian grids are used with immersed boundary methods for complicated geometries
(Yang and Stern, 2009), and the level set based ghost fluid method is used for sharp interface treatment and fully
two-phase coupling with the VOF method for interface tracking. Extension to orthogonal curvilinear grids was
made in V6.2 (Wang et al., 2012a) with enhanced technologies for the interface modeling (Wang et al., 2012b,
c) and similar numerical methods and HPC capabilities as V6.1.
A finite-difference method is used to discretize the governing equations on a non-uniform staggered grid, in
which the velocity components are defined at the cell face centers. All other variables are defined at the cell
centers. Time advancement is based on the semi-implicit four-step fractional step method. The diagonal
diffusion terms are advanced with the second-order Crank–Nicholson method and the other terms by the second-
order explicit Adams–Bashforth method. The pressure Poisson equation is solved to enforce the continuity
equation. The convective terms are discretized using the fifth-order WENO scheme. The other terms are
discretized by the second-order central difference scheme. The pressure Poisson equation is solved using a semi-
coarsening multi-grid solver from the HYPRE library.
The code is parallelized via a domain decomposition (in three directions) technique using the MPI library.
All inter-processor communications for ghost cell information exchange are in non-blocking mode. Parallel I/O
using MPI2 have been implemented such that all processors read from and write to one single file
simultaneously (Yang et al., 2008). In order to speed up the computations and improve the accuracy and
efficiency for very large grid simulations (billions of grid points), some enhanced technologies have been
implemented such as semi-Lagrangian advection schemes and optimized memory usage. The water/air interface
is extracted as PLY polygon file format for post-processing. A multi-block grid capability has been recently
incorporated into CFDShip-Iowa Version 6.2.
Development of the general curvilinear grid solver, V6.3, is in progress, which is built on the success of
V6.1 and V6.2 to achieve all functionalities of V4.5 and beyond. CFDShip-Iowa V6.3 is aimed at the high-
fidelity, high-resolution simulations of fully coupled, multi-scale, multi-phase, turbulent ship flows with fluid-
structure interactions utilizing billions of grid points. The approaches include finite volume method, multi-
block, body-fitted, general non-orthogonal curvilinear structured grids, overset background Cartesian grids, and
highly modularized, developer-friendly code structure written in Modern Fortran (2008) and MPI.
The second-order finite volume method with accurate geometric approximations for non-smooth, non-
orthogonal structured grids is used for the discretization. A generic transport equation is solved for momentum
components and scalars with central difference and high-order upwind schemes used for face-centered value
reconstruction. Exact projection method is implemented for machine-accuracy mass conservation where central
difference and high-order upwind schemes for contra-variant volume flux reconstruction at cell face centers.
Scalable MPI communication using new MPI-3 features will be implemented and MPI sub-array data type is
extensively utilized for scalable MPI communication and I/O in V6.3.

4. Naval Architecture

4.1. Resistance and seakeeping, captive and free running maneuvering, free running course keeping,
and intact and damaged stability
Resistance and seakeeping predictions are included in Gothenburg 2010 (G2010) and upcoming Tokyo 2015
(T2015) workshops. Prediction of resistance is the oldest application of CFD in ship hydrodynamics and its
accuracy has been significantly improved since Gothenburg 1980 (G1980), the first CFD workshop held in
1980. In G2010, 89 submissions of resistance prediction are documented, which is the largest number in the
workshop series (Larsson et al., 2014). More than 90% of the simulations were conducted using grids smaller
than 10M points. The resistance prediction simulations were carried out for a wide range of applications and
conditions. Other than resistance, sinkage and trim, local flow fields such as boundary layer and wake, and wave
patterns were also predicted by many simulations. Different geometries including tankers, container ships, and

3
Copyright © 2014 by ICHD
surface combatants were studied at a range of very small to large Froude numbers (Fr). The simulations showed
average error of 3.3% for resistance for both low and high Fr while sinkage and trim showed less errors for high
Fr. The average error for sinkage/trim at low and high Fr was 9.7% /11% and 35%/55%, respectively. For
seakeeping, several seakeeping test cases were included in G2010 with numerous contributions for each case.
CFD computations of seakeeping have been rapidly increasing since Tokyo 2005 CFD workshop (T2005) in
which there was only one forward-speed diffraction case with no motions. The applications for seakeeping
predictions included a wide range of wave conditions, Froude numbers, and motion conditions. Similar to the
resistance test cases, different geometries including tankers, container ships, and surface combatants were
studied. Grid sizes ranging from 0.4 M to 71 M points were used with a clear trend toward increasing accuracy
with grid size. The CFD predictions are assessed separately for 1st order vs. 2nd order terms. The mean value of
resistance and the amplitude of motions were considered 1 st order terms whereas the amplitude of resistance and
mean value of motions were considered 2nd order terms. The simulations showed large average error for the
second order terms (44%D) while the average error was less than 15% for the first order terms.
Captive and free running maneuvering simulations are included in the SIMMAN 2008 workshop (Stern et
al., 2011) and upcoming SIMMAN 2014. The applications for captive predictions included PMM-type forced
motions such as static rudder, static drift, pure sway, pure yaw, and yaw & drift conditions for different
geometries. For SIMMAN 2008, 16 submissions were received for the forced motion simulations, comprising
different CFD-based methods such as RANS, URANS, and DES. Grid sizes ranging from 2.1 M to 250 M
points were used. It was concluded that finer grids were needed especially for the rudder and appendages and in
regions of large vortices, as well as more advanced turbulence and propeller models for improvements in the
CFD predictions of static and dynamic PMM maneuvers. Overall, the average error for captive maneuvering
simulation was 13.6%D. The largest error values were generally observed for pure yaw and static rudder
simulations. For linear derivatives, the average error was much larger for yaw moment (40%D) than sway force
(15%D). For nonlinear derivatives, the average error value was about 40%D. Free running maneuvering
simulations were reported for limited cases in the SIMMAN 2008 workshop. The maneuvering simulation
included standard maneuver test cases such as turning circle and zigzag. The results showed 6%D error for
trajectories for turning maneuver prediction while larger errors (13%D) were obtained for zigzag maneuver. The
grid sizes were from 0.4M to 14.9 M points for these simulations. For most SIMMAN 2008 computations, the
propulsion was implemented as an axisymmetric body force distributed in the propeller disk. The body force
was specified in a non-iterative manner in which the ship wake on the body force was neglected. Recently, Wu
et al. (2013) used Yamasaki propeller model coupled with the RANS code to give a model that interactively
determines propeller-hull interaction without requiring detailed modeling of the propeller geometry. Yamasaki
model is based on a potential theory formulation, in which the propeller is represented by bound vortex sheets
on the propeller disk and free vortices shed from them downstream of the propeller. Wu et al. (2013) showed the
Yamasaki propeller model could predict successfully the asymmetric wake field. In addition, the propeller rpm
was predicted with less than 0.5% error for Yamasaki compared to 12% for non-iterative axisymmetric body
force. Free running simulations are also conducted with more advanced propulsion system such as water-jet.
Sadat-Hosseini et al. (2013) performed maneuvering simulations for a catamaran and validated the results
against the experimental data (See Fig. 1). The simulations were conducted either for bare hull with integral
force models for water-jet or with actual water-jet with body force impeller defined by pump curves. Turning
maneuver simulations showed average error of 9-22.6%D for CFD simulations with minimum error for the
actual water-jet simulation. Zigzag maneuvers showed larger errors. In addition, the extremely large overshoot
angles in zigzag showed the deficiency of water-jet propulsion system for maneuvering. Since CFD is
computationally expensive for maneuvering in comparison to system based (SB) methods, some studies have
focused on improving the SB mathematical model by using CFD with system identification methods. Araki et
al. (2013) employed CFD free running outputs to improve a 4DOF mathematical model developed for
maneuvering in calm water and following waves. The CFD predictions were first validated against the
experimental data from different facilities including IIHR wave basin (Sanada et al., 2013&2014). For calm
water, it was shown that the average system based prediction error drops from 16% to 8% using the
maneuvering coefficients and rudder forces estimated from CFD free running instead of those from captive
experiments. For waves, Araki et al. (2013) showed that the mathematical model with wave loads estimated
from CFD outputs provide better prediction for maneuvering in moderate following and quartering waves,
4
Copyright © 2014 by ICHD
compared to the original mathematical model with the wave loads computed from slender body theory.
However, the improved mathematical model was too stable in severe waves and unable to predict the
instabilities such as periodic motion or broaching. For upcoming SIMMAN2014 workshop, Sadat-Hosseini et
al. (2014a) conducted simulations for free running maneuvers of KVLLCC2 in calm water using body-force
propeller model and actual propeller (see Fig. 2). The grid size was 6.8M-8.4M for different cases. The
computational cost was 3-5 times higher for the simulations with the actual propeller. The results for turning
maneuver showed E=6.6%D using propeller model and much less error (E=2.2%D) using actual propeller.
Similarly, zigzag simulations showed better prediction using actual propeller. Sadat-Hosseini and Stern (2014)
performed maneuvering simulations for 5415M test cases of SIMMAN 2014 using twin counter-rotating
propellers based on body-force propeller model with total grid size of 6.7M points (see Fig. 3). The results
showed about E=12%D for turning and zigzag 2020 while larger errors were shown for zigzag1010. In addition,
Sadat-Hosseini and Stern (2014) conducted system-based simulations for 5415M maneuvering in calm water.
The maneuvering coefficients were found from system identification using CFD outputs. To estimate the
coefficients, parallel processing technique was used in which CFD free running data for several turning and
zigzag maneuvers were first combined and then used to estimate one set of maneuvering coefficients. The
system based predictions showed an average error of 5.30, 12.64 and 4.67%D for trajectories for turning 35,
zigzag 1010 and 2020, respectively.
Among free running maneuvering simulations, there are very limited studies on local flow. Recently, Sadat-
Hosseini et al. (2014b) studied DES predictions of the local flow including transom wave field and vortex
structures in turning maneuver. Similar study was previously conducted only for straight-ahead condition
(Bhushan et al., 2012). The mean and unsteadiness of transom wave field were predicted with 9% and 11%D
error while the trajectories were predicted with <3%D. The asymmetry of mean wave field was significantly
under predicted due to surprisingly large asymmetry of EFD data. The unsteadiness spectra at few points in the
transom wave field showed f-1.5 scaling. The resolved turbulence kinetic energy was 86% in the transom region.
The simulations showed Karman-like instability at transom, horseshoe vortices at the juncture of strut-hull and
strut-shaft, and shear layer instability at the strut-hull intersection. Figure 4 shows the predicted transom wave
field and vortex structures. Compared to straight-ahead condition, the Karman-like frequencies were 3% higher
while others were 8-35% lower for turning. In addition, the predicted frequency for Karman-like, horseshoe and
shear layer vortex shedding in turning showed 2.4%, 3.7-7.7% and 8.6% asymmetry, respectively.
There are few simulations conducted to investigate free running course keeping and instability. Stern and
Toxopeus (2013) and Sadat-Hosseini et al. (2014c) performed course keeping simulations in calm water, regular
and irregular waves for the fully appended 5415M ship hull, in collaboration with NATO AVT 216 Evaluation
of Prediction Methods for Ship Maneuvering and Control. The results were validated against the experiments
not only for the ship motions but also for the loads on the appendages. The results showed good prediction for
the trajectories and loads on the appendages (<10%D) even for very complex geometries with dynamic
stabilizer and rudders (see Fig. 5). Comparing the irregular wave results with the results computed from regular
wave simulations at several discrete wavelength conditions showed that the ship has similar motion in both
regular and irregular waves with same wavelength condition. The course keeping simulations focusing on intact
instability are summarized in Stern et al. (2013), showing good prediction for different instabilities including
parametric roll, broaching and capsize, surf-riding, and periodic motion. For damaged stability, Sadat-Hosseini
et al. (2014d) showed good prediction for both ship motions and water heights inside the compartment for
damaged ship in calm water and waves.
Overall, free running simulations have been increasing in past few years and it is expected that the future
challenges and method development efforts for modelling, numerical methods and HPC will focus on free
running rather than captive simulations. In addition, more research will focus on improving the SB mathematical
model by using CFD since CFD is computationally expensive in comparison to SB methods.

4.2. Turbulence
Prediction of turbulent viscous flow for ship hulls is of central importance and focused topic at CFD Workshops
since G1980 to most recent G2010. Verification and validation of CFD predictions have been performed for
tanker KVLCC2, container KCS and surface combatant 5415 hull forms at straight ahead conditions. In in
recent workshop extensive local-flow analysis was performed for KVLCC2 (bluff body) and 5415 (slender
5
Copyright © 2014 by ICHD
body) focusing on the effect of turbulence modeling. URANS with anisotropic turbulence model perform better
than isotropic model. For KVLCC2, URANS under predicted axial velocity and vortical strength by 10% and
over predicted turbulent structures by 35%, when compared with the experimental data. DES predicted unsteady
flow with up to 95% resolved turbulence. DES mean flow predictions were quantitatively comparable to that of
URANS, but were over predictive for both velocity and vortical and turbulent structures. DES showed grid
induced separation inside the boundary layer and modeled stress depletion, and the former was resolved by
using delayed DES approach, whereas the latter issue was unresolved. For 5415, URANS provided reasonably
good agreement with the experimental data, but under predicted the vorticity magnitude and boundary layer
bulge, and over predicted turbulent structures at nominal wake plane. In DES, the resolved TKE levels were less
than 3%, thus the results were unacceptable. Nonetheless, provided for the first time plausible description of the
overall vortex structures, and helped in understanding the sparse experimental data. Overall firm conclusions
were not possible since grid and turbulence modeling errors could not be separated and sparseness of
experimental data, especially for turbulence variables and onset and progression of 5415 vortices.
NATO AVT-183: Reliable Prediction of Separated Flow Onset and Progression for Air and Sea Vehicles
research effort for the sea facet focused on procurement of detailed experimental data using PIV techniques
(Yoon et al., 2014; Maksoud et al., 2014; Broglia et al., 2014), and evaluation and validation of CFD predictions
using different codes by NATO members. The study focused on three ship hulls: KVLCC2 at static drift  = 30
deg; 5415 with bilge keels at straight ahead and  = 20 deg; and Delft Catamaran at static drift conditions. Note
that for 5415 cases, EFD data were procured for both planar sections and volumes surrounding the primary
vortices. This allowed evaluation of Q-criteria along the vortex, which enabled validation for the vortex core
predictions for the first time. Validation of 5415 case has been largely completed, and discussed below.
CFDShip-Iowa simulations for the 5415 cases were performed using anisotropic URANS and DES models
using finest adaptive grids to date, to reduce grid errors. In both the cases, URANS results do not improve when
the grid is refined beyond 50M points. The best URANS predictions showed excessive decay of the vortices as
shown in Fig. 6, and resulted large errors for the progression of the vortices as shown in Fig. 7(a), when
compared with experimental data. Considering that the results did improve with grid refinement, the large errors
in URANS predictions were attributed to modeling errors. DES predictions for the straight-ahead case showed
very low resolved turbulence levels, similar to G2010. For the static drift case, DES predictions improved with
grid resolution. On the finest 84M grid, the resolved turbulence levels were >95%, and the flow predictions
compared better with experimental data than those obtained using URANS. However, as shown in Fig. 7(b),
they predicted stronger vortex strength at onset and weaker vortex strength downstream. Note that the large
errors could be partly due to grid resolution issues. CFD submissions using other codes were mostly using
URANS, and one submission for the straight ahead case was using DES. The URANS results from other codes
were very similar to that of CFDShip-Iowa for the static drift case. However for the straight ahead case, solvers
predicted different decay rate of the vortices for similar size grids. The differences could be due to differences in
numerical methods, grid topologies or turbulence model implementation, which needs to be investigated. The
DES submission for the straight-ahead case, showed significantly high vortex strengths than experiments,
similar to CFDShip-Iowa prediction, affirming the limitations of DES models.
Vortex onset and separation in the straight ahead case was identified due to open-type cross flow
separation, wherein the vortex separates from the surface due to the presence of adverse axial pressure gradient
along converging streamline, and is identified from the peak of div(w). The vortex separation patterns for the
static drift case included both open-, closed- and open-closed type separations. The separation pattern and
topology were consistent with those available in the literature. The closed-type separation satisfied the
topological rules expected for a close-separation formed over an isolated body or body intersecting a wall or
free-surface.
Overall, turbulence modeling is a roadblock for improved prediction of viscous flow for ship hulls, as
URANS is too dissipative and DES has limitations for both slender and bluff bodies. For the slender body at
straight ahead condition, DES fails to trigger resolved turbulence. For slender bodies at static drift and bluff
bodies at both straight ahead and static drift conditions, DES predicts sufficient resolved turbulence levels and
the predictions are better than that of URANS, but shows large comparison errors for the progression of the
vortices probably due to modeled stress issue. LES are ideal for accurate CFD predictions, as they have less
dependence on modeling; however, they are prohibitively expensive due to grid resolution requirements in the
6
Copyright © 2014 by ICHD
boundary layer. Hybrid RANS/LES models provide a reasonable alternative, wherein URANS is used in the
boundary layer and LES in the wake. However, more advance hybrid RANS/LES models should be investigated
to address the DES modeling issues. The research should particularly focus on investigation of: turbulence
trigger models to enable transition from RANS to LES for slender body simulations; blended RANS-LES
models as they include explicit LES modeling that are more rigorously validated than the LES mode of single
parameter models, such as DES; and physics based RANS-LES blending rather than grid based blending to
address modeled stress depletion issues.

4.3. Ship-Ship Interaction


CFD computations of ship-ship interaction have been reported in the recent International Conference on Ship
Maneuvering in Shallow and Confined Water. Mousaviraad et al. (2014a) used CFDShip-Iowa to study Hope
and Bobo in replenishment condition in calm water and waves, and in overtaking maneuver in waves. The
average error against EFD was 21%D for calm water, and 10%D for replenishment in waves. The sheltering
effect was significant for oblique waves, with 105% difference between mid-mid and mid-bow configuration.
The separation distance effect was more important for head waves than oblique waves, being 43% and 23%
respectively. During the overtaking, the interaction effect decreases motions and increases sway forces, roll and
yaw moments, being more significant for the smaller vessel. Sadat-Hosseini et al. (2011) and Wu et al. (2013)
investigated the interaction between two different tankers Aframax and KVLCC2 using CFDShip-Iowa. The
ships were free to heave and pitch advancing in shallow water with same speed and fixed separation distance.
Both ships were appended with rudder and operating propellers, which were modelled by axisymmetric body
force propeller model with same RPM as experiments. Overall, the simulations showed large errors for
predicted forces, moments and motions compared with the experimental data. Later, it was found the
longitudinal positions of the two ships in the experiments were not reported correctly. Therefore, the simulations
were repeated with the revised conditions but the errors were still large and thus more studies should be
conducted to evaluate the experimental setup. In addition, the accuracy of the axisymmetric body-force
propeller model for propulsion in shallow water should be investigated.

4.4. Advanced hull forms and fluid-structure interaction: ACV/SES, WAM-V, planing hulls
CFD studies of advanced hull forms impose significant challenges due to complex and multi-disciplinary
modeling requirements, very high speeds introducing different physics than conventional ships, and difficulties
in validation studies due to limitations in model testing and limited measurements in sea trials. Modeling
requirements are different for specific hulls, e.g. fluid-structure interactions (FSI) including multi-body
dynamics (MBD) for suspension systems and finite element (FE) modeling for flexible hulls.
ACV/SES capabilities are implemented in CFDShip-Iowa including cushion models, seal models, air-flow
over the above water seals and superstructure, decoupled cushion cavity flow, waterjet propulsion with side
forces and yaw moments induced by nozzle rotations and reverse buckets, and air-fan propulsion model.
Validation simulations are carried out for a combined SES/ACV ship (T-Craft) for captive tests in deep and
shallow water. Free-running simulations of T-Craft in turning and zigzag maneuvers in deep and shallow water
and in calm water and waves are also carried out. Recent analyses showed that the resistance and moment due to
cushion pressure distribution inside the cavity is significant for seakeeping cases while not considered in the
initial simulations. The improved results will be published for captive validation studies and free-running
demonstration simulations.
The Wave Adaptive Modular Vessel (WAM-V) is an ultra-light flexible catamaran that conforms to the
surface of the water through a collective suspension and is modularly designed enabling a wide variety of
applications. The springs, shock absorbers, and ball joints articulate the vessel such that the hulls can move
semi-independently and along with the inflatable pontoons adapt to the water surface/waves to mitigate
structural stresses and reduce drag. WAM-V capabilities are implemented in CFDShip-Iowa including: LS_IBM
(level-set immersed boundary) method for treatment of the gap between pontoon and hinged pod, a two-body
dynamics model for hinged pod motions, and a jet force model moving with hinged pod for free-running
simulations. Captive calm water verification and validation studies are carried out with average error of 5.7%D
(Mousaviraad et al., 2013a). Validation against full-scale sea trial data and coupling with MBD modeling are

7
Copyright © 2014 by ICHD
carried out in collaboration with Prof. Mehdi Ahmadian of Virginia Tech University. Free-running validation
studies are carried out against sea trial data in calm water and seas (Conger et al., 2014). For simulations in
waves, statistical analysis of the sea trial data in waves is conducted to provide an estimate of the dominant
encounter frequency. CFD regular head waves simulations (Fig. 8) are carried out at the dominant encounter
frequency and with a wave height over wavelength of H/λ=1/64, the typical value for sea state 3. The results are
compared with sea trial data in Fig. 9 and show that although the EFD data have large peaks, their standard
deviation (SD) values converge to values very close to CFD. The CFD regular head wave results are then used
by Virginia Tech University as inputs to run a 2-post shaker rig testing and a SIMULINK virtual shaker rig
modeling and the results for the payload suspension motions are shown in Fig. 10 with very good agreement.
MBD modeling for the suspended payload is carried out for a 2DOF cylinder drop as a first step to WAM-V
suspended payload modeling. The SIMULINK MBD code is coupled with CFDShip-Iowa in 1-way weak
coupling and 2-way strong coupling approaches and the results are shown in Fig. 11. The 2-Way coupling
results show significant improvement over 1-Way results: for the un-sprung mass displacement the initial slope
after the pontoon hits the water free surface is more accurately predicted, the double hump at the first peak is
predicted, and the frequency of occurrence is maintained correctly through the displacement curve; for the
sprung mass displacement, 2-Way results follow the EFD displacement curves both in magnitude and
frequency, especially for the first second, while after 1 second the sprung mass displacement is slightly over-
predicted. Overall the results are validated with acceptable agreement. Future work will couple the CFDShip-
Iowa and the MBD model for WAM-V, and perform validation studies against measurements of the
motions/accelerations of the suspended payload during the full-scale sea trials.
Planing hull capabilities including hydrodynamic performance and structural loads and slamming are
implemented and validated for calm water, regular waves, and irregular waves for the Fridsma geometry
(Mousaviraad et al., 2013b; Mousaviraad et al., 2014b). CFD and EFD studies are carried out to validate the
hydrodynamic forces, moments, hull pressures, accelerations, motions, and the multiphase free-surface flow
field generated by the USNA planing craft at high-speed (Fr=1.8 - 2.1) in calm water and regular and irregular
waves (Fu et al., 2014). The work is conducted by collaborations with Carolyn Judge of United States Naval
Academy (USNA). CFDShip-Iowa simulations for calm and regular waves were carried out blind, before EFD
data was available. Calm water spray root at Fr=1.83 is compared in Fig. 12 with underwater surface photo from
EFD indicating very close agreement. CFDShip-Iowa V5.5 simulations with volume of fluid free surface solver
showed negligible effects on resistance and motions, while the extension of the jet spray flow was resolved
better than V4.5 level-set solver. Regular wave results for USNA experiments (Run 43 and 44) and CFD
simulations using CFDShip-Iowa V4.5 and NFA solvers are compared in Fig. 13 for motions and slamming
pressure. The phase of the heave and pitch is well predicted, while the amplitude of the numerical simulations is
greater than measured experimentally. Pitch motions at twice the lowest frequency are not evident in
simulations performed using either CFDShip-Iowa or NFA. Single point pressure measurements show good
agreement for slam duration while the re-entering pressure amplitudes are under-predicted for both codes. A
smaller time step may be needed to capture the peak pressure. The emerging peak pressures are missed in NFA
simulations while captured in CFDShip with close agreement. Irregular waves simulations are validated with
good agreement in terms of expected values and standard deviations of motions, accelerations, and slamming
pressures. Slamming statistical studies are carried out for both experimental data and simulation results and
validation results are shown in Fig. 14 for slamming pressure. Extreme slamming events are studies both for
EFD and CFD by examining the standard score for re-entering pressure ( = ( )⁄ ). For EFD, 4 slam
events with >2 and 6 events with 1< <2 are detected. These events are found to correlate with ship motions,
namely the vertical velocity of the ship bow at the time of impact. CFD studies are carried out to provide further
insight by correlating the extreme slam events with relative bow/wave motions as well as history of previous
zero crossing waves. The CFD extreme events are grouped in 3 categories: >1.5 (3 events), 1< <1.5 (4
events), and 0< <1 (14 events). For each slam event, wavelength over ship length (λ/L) and wave height over
wavelength (H/λ) values for the immediate wave, as well as averaged values for the last 2, 3, 4, and 5 waves are
calculated. In group 1, slam pressures correlate 100% with smaller λ/L and larger H/λ for the last 3 waves. For
groups 2 and 3, strongest correlations are for larger H/λ averaged over the last 2 and 3 waves, respectively.
Considering all the slams in all 3 groups, strongest correlation is found for smaller λ/L from the last 3 waves and

8
Copyright © 2014 by ICHD
larger H/λ from the last 2 waves. Type-2 slams characterized by containing only one pressure peak (re-entering
pressure) with smaller peak values and shorter duration are identified both in EFD and CFD.
Fluid-structure interactions (FSI) studies (Volpi et al., 2015) are carried out for the Numerette planing hull
(slamming load test facility at Lehigh University) to provide a better understanding of slamming using
benchmark full-scale validation EFD data. The studies are conducted in collaboration with Dr. Joachim
Grenestedt of Lehigh University. Initially rigid body CFD simulations are conducted for both bare hull and
appended hull with sterndrive unit and body-force propeller model excluding the superstructure. The predicted
motions and loads are used for one way coupling with FE model for composite bottom panels to evaluate
displacement, strain, and stress. CFDShip-Iowa is used for CFD simulations and the commercial FE code
ANSYS is used as structural solver. Studies are carried out in calm water (Fr=0.7) and different regular head
waves conditions at Fr=0.7, 2.24 and 2.9. CFD/FE results show good prediction for displacement, strain, and
stress distribution for both starboard and bottom panels. Fig. 15 shows the panel force and displacement for a
regular head wave simulation with Sea-State 3 most probable wave conditions at Fr=0.7. Fig. 16 shows EFD
and CFD-FE strain for a regular head wave simulation with sea state 3 most probable wave condition at Fr=2.9.
Two-way coupling will be implemented by first using modal analysis with added mass modeling, and then fully
coupled CFD-FEA. FSI V&V studies are also planned for slamming loads on Athena semi-planing hull.

5. Ocean Engineering
Simulations of 3D unsteady separation (vortex shedding) around offshore structures and wave run-up induced
by ocean waves are still challenging for ocean engineering applications. Recently, the capabilities of state-of-
the-art CFD codes for vortex shedding and wave run-up are investigated in ITTC ocean engineering workshop
held in Nantes, France October 17-18, 2013. The capabilities of CFDShip-Iowa V4.5 and V6.2 for these
applications are reported in Yeon et al. (2013) and Yoon et al. (2013). The studies focused on the flow around
single/multiple cylinder(s), a typical geometry for both applications.

5.1. Single- and two-phase vortex shedding


In Koo et al. (2014) the two-phase turbulent flow past an interface-piercing circular cylinder was studied
using large-eddy simulation with a Lagrangian dynamic subgrid-scale model. It was shown that the air-water
interface makes the separation point more delayed for all regimes of Re and the air-water interface structures are
remarkably changed with different Froude numbers. However, the deep flow did not display the correct single-
phase flow behavior due to the deficient grid resolution and non-conservative convection scheme, among other
issues, employed with CFDShip-Iowa V6.2. Yeon et al. (2013) conducted a detailed study of the single-phase
vortex shedding around a circular cylinder for ITTC ocean engineering workshop test cases. The simulations are
conducted using CFDShip-Iowa V6.2, covering sub- to super-critical Re. A careful verification and validation
study were carried out. The effects of aspect ratio/span length, conservative vs. non-conservative convection
schemes, and grid resolution were investigated. The mean velocity, mean pressure, Reynolds stresses, and TKE
distribution were obtained and discussed. The snapshot POD method was employed to analyze flow structures
in the boundary layer, shear layer and wake. Fig. 17 shows coherent flow structures visualized with iso-surface
of Q criterion colored by the non-dimensional eddy viscosity. The wake width and amplitude of the shedding is
large for the sub-critical Reynolds number (Re) and become smaller as Re increases. Energy spectra of the
streamwise velocity in the shear layers are also shown in Fig. 17. At lower wavenumbers the energy spectra give
scaling exponents close to the Kolmogorov slope, which verifies that the large-eddy simulations properly
modeled the turbulence and preserved the correct energy decay behavior, although the ranges of wavenumbers
with the -5/3 spectral slope become narrower as Re increases. On the other hand, for larger wave numbers, the
rates of energy decay are faster than Kolmogorov’s decay law and the scaling slopes become much steeper. A
main cause of this rapid decay is the numerical dissipation from the upwind convection schemes used in the
simulations. The Kolmogorov wavenumbers estimated from the local velocity fluctuations are smaller than the
grid cut-off wavenumbers. This indicates the grid resolution is adequate in the shear layers, where the
turbulence intensity is usually lower than that in the wake. Fig. 18 shows comparisons for CD, CLRMS, LS/TS,
and –Cpb. The drag crisis is well predicted, although more cases in the critical and post-critical regime are
desirable. The LES CLRMS is close to the most reliable data for sub-, critical and super-critical Re. The angle of

9
Copyright © 2014 by ICHD
separation is close to the experiments for sub- and critical Re, but substantially under predicted for super-critical
Re. The base suction pressure shows good agreement with the experiments for sub- and super-critical Re, but is
under predicted for critical Re. The largest difference is for critical Re, where the drag drops sharply with small
changes in Re resulting in large changes in CD between facilities and likely simulations. The grid resolution,
convection scheme, and the effect of upstream disturbance ubiquitous in experiments, but missing in the
simulations, are most likely responsible for the under-predicted separation angle for the critical Re.

5.2. Wave run-up


CFD simulations of wave run-up around single/multiple truncated vertical cylinder(s) for ITTC ocean
engineering workshop tests cases were conducted by Yoon et al. (2013). The simulations were conducted in
regular head waves for various wave conditions including /D=4.7 and /D=21.9 with H/λ=1/30, 1/16 and 1/10.
Sensitivity studies are conducted for the effects of grid distribution, domain size and turbulence model.
Validation studies focused on averaged wave height at crest/trough and 0th, 1st and 2nd harmonics for wave
elevation and horizontal force. CFD predictions were assessed separately for 1st and 2nd order variables. The
averaged wave height at crest/trough and the 1st harmonics were considered as the 1st order variables, whereas
the 0th and 2nd harmonics were considered as the 2nd order variables. In addition, the wave field pattern around
cylinder(s), vortex shedding, and interaction among cylinders were analyzed. The grid sensitivity for the 1st and
2nd order variables was 3.17% and 77.74%, respectively, both less than the facility bias estimated from the
provided experimental data from two facilities. Nonetheless, the 2nd order variable sensitivity was large
indicting the need for finer grids to resolve 2nd order terms. The domain size sensitivity was also very small,
1.14% and 2.34% for 1st and 2nd order variables. The turbulence model sensitivity was conducted using
URANS and DES and the sensitivity for 1st and 2nd order variables was 1.64% and 6.55%, respectively,
suggesting that the URANS turbulence model is sufficient for the validation studies. The validation studies
showed 10% error for wave crest/trough, 7% error for the 1st harmonic of wave elevation; and 70% error for the
2nd order variables including the 0th and 2nd harmonics of wave elevation. The horizontal forces also showed 9%
error for the 1st harmonic amplitude while larger errors are predicted for the mean and 2 nd harmonics. The
detailed study of the wave field showed that the mean wave field elevations are similar to the free surface
elevations for a cylinder in a steady flow due to the large wave induced current (up to 15% of the orbital
velocity for the steepest wave). The results showed larger effects of the wave steepness on the wave mean and
2nd amplitude than on the 1st harmonic amplitude. The wave steepness effect was also more prominent for
/D=4.7 than /D=21.9. The studies were also conducted on the total wave field to evaluate the diffracted wave
pattern. The nonlinearities in the incident wave caused difficulties extracting the diffraction wave from the total
wave field. However, the total wave field could show the diffraction wave at upstream which was more
dominant for /D=4.7 and steeper waves. Fig. 19 and 20 show the wave field for single and four cylinder cases.
The studies on vortex structures showed more vortex shedding for longer wave conditions as its longer wave
period provides enough time to develop vortices around the cylinder. For both wavelength conditions, the vortex
shedding is more at instants the wave crest is located near the cylinder as the flow field velocity is larger. Lastly,
the comparison of four and single cylinder cases shows that the interaction of cylinders increases wave trough
for 4-10% while the wave crest increases about 9-25%. The largest interaction effect is found for the shoulder
side of the cylinders.

6. Fundamental Physics

6.1. IBM for idealized and practical geometries


Immersed boundary methods are simple and efficient approaches for many problems with complex geometries
and moving boundaries, thanks to the relaxation of the requirement of generating boundary-fitting grids in
numerical simulations. CFDShip-Iowa V6.1 is a Cartesian grid solver utilizing a direct forcing immersed
boundary method. The research focus is on efficient strong coupling schemes for fluid-structure interactions and
the extension to wave-body interaction problems in naval architecture and ocean engineering. In Yang & Stern
(2012) an efficient strong coupling scheme for 6DOF motion prediction was developed. The predictor-corrector
loop in each time step includes the adjustments of the structure displacements and velocities, but the fluid flow

10
Copyright © 2014 by ICHD
solver was excluded. Then in Yang & Stern (2013) an efficient and robust immersed boundary setup procedure
was developed for further accelerating the strongly coupled simulations of fluid-structure interactions. This
approach can be a viable choice for particulate flows as shown in Yang & Stern (2014). Currently a non-
iterative strong coupled scheme has been developed. Fig. 21 shows the flow charts of three different strong
coupling schemes in one time step for fluid-structure interaction problems. Compared with the scheme in Yang
et al. (2008) with a complete iterative loop including multiple Poisson solves and the scheme in Yang & Stern
(2012) with one Poisson solve but multiple local reconstruction steps, the present scheme utilizes an
intermediate step with a non-inertial reference frame (NIRF) attached to a solid body and no iterative loop is
involved. The improved efficiency and reduced algorithm complexity is evident. The next step of development
will be combination of this new scheme with an efficient two-phase flow solver for ultra-scale simulations of
6DOF motions in naval architecture and ocean engineering. It should be pointed out that the development of
wall models in immersed boundary methods is necessary if high Reynolds number flows are the target
application and a reasonable approximation of the turbulent boundary layers is required.

6.2. Bubble, droplet, and spray in breaking waves


Air entrainment, bubbles, droplets, jets, and spray in breaking waves are of great importance to ship
hydrodynamics. Previous experimental and computational studies are mainly focused on the global structures of
the wave breaking. With the development of the CFD technology, detailed studies of the small scale structures,
such as water droplets and air bubbles, in the two-phase region become possible. In the study by Wang et al.
(2012d), wave breakings around a wedge-shaped bow and over a submerged bump are simulated using very
large grids (1.0~2.2 billion grid points). This study is the first attempt to directly simulate the unsteady and
energetic wave breaking flows to the scale of micrometers. In Wang et al. (2014), even large grids (up to 11.8
billion) are used in order to resolve the bubbles/droplets in breaking Stokes waves at the scale of several
micrometers. Fig. 22a shows the wave profile at time t = 1.76 when the splash-ups are being generated after the
wave plunging. The 3D interface instability in the spanwise direction is clearly demonstrated in Fig.22b. The
study of the flow over a bump in a shallow water flume by Gui et al. (2014a,b) showed that the Görtler type
centrifugal instability is the most relevant mechanism for the free surface instabilities. Fig. 22c and d show the
applications of the Görtler inviscid instability and Rayleigh instability theories in the stream-wise central plane,
respectively. In the wave breaking region, Görtler stability criterion is violated in most locations and Rayleigh
stability criterion is broken only in small regions. These results support the idea that breaking wave instabilities
are mainly due to Gortler type centrifugal instability. Fig. 22e shows the formation of bubbles/droplets in the
process of wave breaking. Power-law scaling for the bubble size distribution was obtained with two different
slopes separated by a Hinze radius of 1.2 mm as shown in Fig. 22f. The simulation results are in good agreement
with the experimental findings.

6.3. Cavitation
Cavitation degrades the performance of lifting surfaces found on ships, such as propeller blades and rudders and
may cause erosion. Past computational models have generally been homogenous mixture models, which average
the effects of many bubbles, or discrete bubble models, which model only a limited number of bubbles. In
Michael et al. (2014) a new sharp interface cavitation model was described within the framework of CFDShip-
Iowa V6.2. The interface is advected using a volume of fluid method with the addition of an additional velocity
due to phase change. The phase change component of the interface velocity is modeled using a simplification of
the Rayleigh-Plesset equation computed through a volume source term included semi-implicitly in the pressure
Poisson equation. A marching cubes method is used to compute the interface area in each computational cell
and for the determination of the phase at the cell and face centers. Fig. 23a shows a time series of cavitation on
a 2D NACA66 hydrofoil at a 6° angle of attack. The details of the shedding process can be seen. Fig. 23b
shows the bubble growth, merging, and advection process in the simulation of the same foil in 3D shortly after
cavitation inception. This type of high fidelity simulation offers the opportunity for deeper insight into the
physics of cavitating flows.

11
Copyright © 2014 by ICHD
7. Uncertainty Quantification
Initial research focused on development and application of deterministic verification and validation (V&V)
methodologies and procedures for high-fidelity CFD simulations. Initial studies for validation methodologies
(Coleman and Stern, 1997) were subsequently extended to verification procedures for deterministic uncertainties
stemming from iterative, grid and time-step convergence (Stern et al., 2006; Xing and Stern, 2010). V&V
methodologies and examples were presented at the AVT-147 Symposium on Computational Uncertainty (Stern,
2007).Recently, the research focus moved to stochastic uncertainty quantification (UQ) methods as an essential
part of stochastic design optimization for real ocean environment and operations, such as robust design
optimization (RDO) and reliability based design optimization (RBDO). UQ research was undertaken within
NATO AVT 191 “Application of Sensitivity Analysis and UQ to Military Vehicle Design”. The objective was
the development and validation of efficient UQ methods for application to realistic ship hydrodynamic
problems. Non-intrusive UQ methods were addressed with high-fidelity physics-based CFD solvers. Evaluation
metrics for efficient UQ methods were developed, based on deterministic and stochastic convergence criteria
and validation versus numerical benchmark (Mousaviraad et al., 2013c; Diez et al., 2013), and efficiency of
overall UQ procedure by assessing the number of CFD simulations required to achieved prescribed error
thresholds (Volpi et al., 2014). Numerical benchmarks were provided by statistically convergent MC simulation
with direct use of CFD computations. UQ methods included metamodel-based Monte Carlo (MC) simulation,
quadrature formulas, and polynomial chaos methods. Applications covered unit studies and advanced industrial
problems. Specifically, a unit problem for a NACA 0012 hydrofoil with variable Reynolds number was
presented and assessed in Mousaviraad et al. (2013c). The high-speed Delft catamaran (DC) advancing in calm
water with variable Froude number and geometry was presented and studied in Diez et al. (2013). DC in
stochastic irregular and regular head waves (see Fig. 24) with variable speed and geometry was assessed in He
et al. (2013a). A combination of UQ problems for the DC was selected from Diez et al. (2013) and He et al.
(2013a) and used for further investigation in He et al. (2013b), focusing on the polynomial chaos method, and
Volpi et al. (2014), focusing on dynamic metamodels.
In conclusion, stochastic UQ methods were found mature for application to realistic stochastic optimization
problems. Based on the evaluation metrics, MC with dynamic metamodels was found the most promising
method overall. The high computational efficiency of dynamic metamodels, by auto-tuning and adaptive
sampling, makes the approach also recommended for stochastic optimization. Metamodel-based UQ has been
applied to stochastic design optimization of DC in real ocean environment and operations, as shown in Diez et
al. (2013) and Tahara et al. (2014). Future extensions include the application of metamodel-based UQ and
optimization to multi-disciplinary analysis and optimization (MDA, MDO) of FSI problems.

8. Future Research
The oncoming exascale HPC era is to change our approaches to grand scientific and engineering challenges and
to transform modeling and simulation into a specified discipline of predictive science. Current mainstream
RANS solvers for ship hydrodynamics are expected to continue performing well for even larger grids of up to a
few billions of points. However, there will be a threshold that further increase of grid size cannot improve the
results anymore because of the inherently limited RANS/DES turbulence models and the widely-used lower-
order discretization schemes. High-fidelity, first-principles-based simulations with unprecedented resolution can
reveal vast unknown temporal-spatial correlations in multi-scale and multi-physics phenomena that are beyond
today’s computing and experimental capabilities. With comprehensive verification and validation procedures,
assisted by targeted physical experiments, and rigorous uncertainty quantification, they are to revolutionize ship
hydrodynamics research and, along with optimization techniques, the ship design process for greatly reduced
design cycles and cost and much improved operation safety and economy. Therefore, the next-generation, high-
fidelity ship hydrodynamics solvers have to be developed aiming at the oncoming exascale computing
platforms, and addressing modeling issues, discretization schemes, and HPC memory and scalability restraints
at the same time.

12
Copyright © 2014 by ICHD
Acknowledgments
This work was supported by research Grants from the Office of Naval Research (ONR), with Dr. Patrick Purtell,
Dr. Ki-Han Kim, Dr. Thomas Fu, Ms. Kelly Cooper; Dr. Roshdy Barsoum, and Dr. Robert Brizzolara as the
program managers. The fluid-structure interaction studies are performed in collaboration with Dr. Joachim
Grenestedt of Lehigh University. The WAM-V studies are conducted in collaboration with Dr. Mehdi
Ahmadian of Virginia Tech University. The planning hull studies are conducted in collaboration with Dr.
Carolyn Judge of USNA. The simulations were performed at the Department of Defense (DoD) Supercomputing
Resource Centers (DSRCs) through the High Performance Computing Modernization Program (HPCMP).

References
1. Araki Motoki, Sadat-Hosseini Hamid, Sanada Yugo, Umeda Naoya, and Stern Frederick, “System
Identification using CFD Captive and Free Running Tests in Severe Stern Waves,” Proceedings 13th
International Ship Stability Workshop, Brest, France, September 23 – 26, 2013.
2. Bhushan S., Xing, T., Stern, F., “Vortical Structures and Instability Analysis for Athena Wetted Transom
Flow with Full-Scale Validation”, J. Fluids Eng. 134(3), 031201, 2012.
3. Bhushan, S., Yoon, H., Stern, F., Guilmineau, E., Visonneau, M., Toxopeus, S., Simonsen, C., Aram, S.,
Kim, S-E., Grigoropoulos, G., Petterson, K. and Fureby, C. “CFD Validation for Surface Combatant 5415
Straight Ahead and Static Drift 20 Degree Conditions.” IIHR Report 2014.
4. Broglia, R., G. Aloisio, M. Falchi, S. Grizzi, S. Zaghi, M. Felli, M. Miozzi, F. Pereira, F. Di Felice and F.
Stern, Measurements of the Velocity Field Around the DELFT 372 Catamaran in Steady Drift, 29th
Symposium on Naval Hydrodynamics, Gothenburg, Sweden, 26-31 August 2012.
5. Campana, E.F., Ship Design under Uncertainty via high-fidelity stochastic optimization, Annual General
Meeting of the Schiffbautechnische Gesellschaft e.V., Berlin, Germany, 2013.
6. Coleman, H.W., Stern, F., Uncertainties in CFD Code Validation, ASME J. Fluids Eng., Vol. 119, pp. 795-
803, 1997.
7. Conger, M., Mousaviraad, S.M., Stern, F., Peterson, A., Ahmadian, M., “URANS CFD for Two-Body
Hydrodynamic Simulation of Wave Adaptive Modular Vessels (WAM-V) and Validation against Sea
Trials”, accepted for publication in Naval Engineers Journal, special edition: “The current fleet, the next
class and the new prototypes”, 2014.
8. Diez, M., Chen, X., Campana, E.F., Stern, F., “Reliability-Based Robust Design Optimization for Ships in
Real Ocean Environment”, 12th International Conference on Fast Sea Transportation, FAST2013,
Amsterdam, the Netherlands, December 2013.
9. Diez, M., He, W., Campana, E.F., and Stern, F., “Uncertainty Quantification of Delft Catamaran
Resistance, Sinkage and Trim for Variable Froude Number and Geometry Using Metamodels, Quadrature
and Karhunen-Loève Expansion,” J. Marine Science Technol., Vol. 19, No. 2, pp. 143-169, 2013, DOI:
10.1007/s00773-013-0235-0.
10. Fu, T.C., Brucker, K.A., Mousaviraad, S.M., Ikeda, C.M., Lee, E.J., O’Shea, T.T., Wang, Z., Stern, F., and
Judge, C.Q., “A Computational Fluid Dynamics Study of the Hydrodynamics of High-Speed Planing Craft
in Calm Water and Waves,” 30th Symposium on Naval Hydrodynamics, Hobart, Tasmania, Australia, 2-7
November 2014.
11. Fujiwara, T., Ueno, M., Nimura, T., “An Estimation Method of Wind Forces and Moments Acting on
Ships,” Mini Symposium on Prediction of Ship Maneuvering Performance, Tokyo, Japan, 2001.
12. Gui, L., Yoon, H., Stern, F., Experimental and theoretical investigation of Instabilities for flow over a bump
in a shallow water Flume with steady downstream wave train, IIHR Technical Report 487, 2014a.
13. Gui, L., Yoon, H., Stern, F., “Techniques for measuring bulge–scar pattern of free surface deformation and
related velocity distribution in shallow water flow over a bump”, Experiments in Fluids, 55:1721, 2014b.
14. He, W., Diez, M., Zou, Z., Campana, E.F., and Stern, F., “URANS study of Delft catamaran total/added
resistance, motions and slamming loads in head sea including irregular wave and uncertainty quantification
for variable regular wave and geometry,” Ocean Engineering, Vol. 74, 2013, pp. 189-217, 2013, DOI:
10.1016/j.oceaneng.2013.06.020.
15. He, W., Diez, M., Campana, E.F., Stern, F., and Zou, Z., “A Polynomial Chaos Method in CFD-Based
Uncertainty Quantification Study for Ship Hydrodynamic Performance,” Journal of Hydrodynamics, Vol.
25, No. 5, 2013, pp. 655–662.
16. Huang J., Carrica P., Stern F., “Semi-coupled air/water immersed boundary approach for curvilinear
dynamic overset grids with application to ship hydrodynamics”, International Journal Numerical Methods
Fluids, Vol. 58, 2008, pp. 591-624.

13
Copyright © 2014 by ICHD
17. ITTC 2011, The specialist Committee on computational fluid dynamics, Proceedings of 26th International
Towing Tank Conference, Rio de Janeiro, Brazil, 28 August – 3 September, 2011.
18. Koo, B., Yang, J., Yeon, S., Stern, F., Reynolds and Froude number effect on the flow past an interface-
piercing circular cylinder, Int. J. Nav. Archit. Ocean Eng., in press, 2014, DOI: 10.2478/IJNAOE-2013-
0197.
19. Larsson, L., Stern, F., and Visonneau, M. (editors), “Numerical Ship Hydrodynamics: An Assessment of the
Gothenburg 2010 Workshop,” Springer, 2014, ISBN: 978-94-007-7188-8, 318 pages.
20. Maksoud, M-A. et al. “Experimental and Numerical Investigations on Flow Characteristics of the KVLCC2
at 30° Drift Angle.” NATO AVT 183 report, 2014.
21. Michael, T., Yang, J., Stern, F., Modeling Cavitation with a Sharp Interface, 30 th Symposium on Naval
Hydrodynamics, Hobart, Tasmania, Australia, 2-7 November 2014
22. Mousaviraad, S.M., Bhushan, S., and Stern, F., “URANS Studies of WAM-V Multi-Body Dynamics in
Calm Water and Waves,” Third International Conference on Ship Maneuvering in Shallow and Confined
Water, Ghent, Belgium, 3-5 June 2013a.
23. Mousaviraad, S.M., Wang, Z., and Stern, F., “URANS Studies of Hydrodynamic Performance and
Slamming Loads on High-Speed Planing Hulls in Calm Water and Waves for Deep and Shallow
Conditions,” Third International Conference on Ship Maneuvering in Shallow and Confined Water, Ghent,
Belgium, 3-5 June 2013b.
24. Mousaviraad, S.M., Wei, H., Diez, M., and Stern, F., “Framework for convergence and validation of
stochastic UQ and relationship to deterministic verification and validation,” International Journal for
Uncertainty Quantification, Vol. 3, No. 5, pp. 371-395, 2013c.
25. Mousaviraad, S.M., Sadat-Hosseini, S.H., Carrica, P.M., Stern, F., “URANS Studies and Validation of
Ship-Ship Interactions in Calm Water and Waves for Replenishment and Overtaking Conditions,”
Submitted to Journal of Ocean Engineering, 2014a.
26. Mousaviraad, S.M., Wang, Z., Stern, F., “URANS Studies of Hydrodynamic Performance and Slamming
Loads on High-Speed Planing Hulls in Calm Water and Waves for Deep and Shallow Conditions,”
submitted to Applied Ocean Research, 2014b.
27. Sadat-Hosseini, H., Wu, P.-C., Toda, Y., Carrica, P., Stern., F. ‘URANS studies of ship-ship interactions in
shallow-water’, 2nd Intl Conf. on Ship Manoeuvring in Shallow and Confined Water, Norway, 2011.
28. Sadat-Hosseini, H., Chen, X., Kim, D.H., Milanov, E., Georgiev, S., Zlatev, Z., Stern, F., “CFD and
System-Based Prediction of Delft Catamaran Maneuvering and Course Stability in Calm water”, 12th
International Conference on Fast Sea Transportation, 2013.
29. Sadat-Hosseini, H., Wu, P.C., Carrica, P.M., Stern, F., “CFD simulations of KVLCC2 maneuvering with
different propeller modeling”, In preparation for SIMMAN2014 workshop, 2014a.
30. Sadat-Hosseini, H., Kim, D.H., Taylor, G.L., Fu, T., Terrill, E., and Stern, F., “Vortical structures and
instability analysis for Athena in turning maneuver with full-scale validation,” 30th Symposium on Naval
Hydrodynamics, 2014b.
31. Sadat-Hosseini, H., Stern F., Toxopeus S., “CFD simulations of course keeping in irregular waves for
5415M”, In preparation for Ocean Engineering, 2014c.
32. Sadat-Hosseini, H., Kim, D.H, Lee, S.K., Rhee, S.H., Carrica, P., Stern, F., “CFD and EFD study of
Damaged Ship Stability in Regular Waves”, In preparation for Ocean Engineering, 2014d.
33. Sadat-Hosseini, H., Stern F., “System based and CFD simulations of 5415M maneuvering”, In preparation
for SIMMAN2014 workshop, 2014.
34. Sanada, Y., Elshiekh, H., Toda, Y., Stern, F., “Effects of Waves on Course Keeping and Maneuvering for
Surface Combatant ONR Tumblehome,” 30th Symposium on Naval Hydrodynamics, 2014.
35. Sanada, Y., Tanimoto, K., Takagi, K., Gui, L., Toda, Y., Stern, F., “Trajectories for ONR Tumblehome
maneuvering in calm water, ” Ocean Engineering 72, 45-65, 2013.
36. Stern, F., “Quantitative V&V of CFD Solutions and Certification of CFD Codes with Examples for Ship
Hydrodynamics,” Symposium on Computational Uncertainty, AVT-147, December 2007, Athens, Greece.
37. Stern, F., Agdrup, K., Kim, S.Y., Hochbaum, A.C., Rhee, K.P., Quadvlieg, F., Perdon, P., Hino, T., Broglia,
R., and Gorski, J., “Experience from SIMMAN 2008—The First Workshop on Verification and Validation
of Ship Maneuvering Simulation Methods”, Journal of Ship Research, Vol. 55, No. 2, pp. 135-147, 2011.
38. Stern, F., Toxopeus, S., “Chapter 1 – Experimental and Computational Studies of Course Keeping in Waves
for Naval Surface Combatant”, NATO AVT-161 report, 2013.
39. Stern, F., Wilson, R., and Shao, J., 2006. Quantitative approach to V&V of CFD simulations and
certification of CFD codes, Intl. J. for Numerical Methods in Fluids, Vol. 50, pp. 1335-1355.
40. Stern, F., Yang, J., Wang, Z., Sadat-Hosseini, H., Mousaviraad, M., Bhushan, S., Xing, T., “Computational
Ship Hydrodynamics: Nowadays and Way Forward”, 29th Symposium on Naval Hydrodynamics
Gothenburg, 2012.
14
Copyright © 2014 by ICHD
41. Stern, F., Yang, J., Wang, Z., Sadat-Hosseini, H., Mousaviraad, M., Bhushan, S., Xing, T., “Computational
ship hydrodynamics: nowadays and way forward,” International Ship Building Progress, Invited paper, Vol.
60, No.1-4, pp. 3–105, 2013.
42. Tahara, Y., Diez, M., Volpi, S., Chen, X., Campana, E.F., Stern, F., “CFD-based multiobjective stochastic
optimization of a waterjet propelled high speed ship”, 30th Symposium on Naval Hydrodynamics, Hobart,
Tasmania, Australia, 2-7 November 2014.
43. Volpi, S., Diez, M., Gaul, N.J., Song, H., Iemma, U., Choi, K.K., Campana, E.F., Stern, F., “Development
and validation of a dynamic metamodel based on stochastic radial basis functions and uncertainty
quantification,” Structural Multidisciplinary Optimization, 2014.

44. Volpi, S., Sadat-Hosseini, H., Kim, D.H., Diez, M., Stern, F., Thodal, R., and Grenestedt, J.G.,
“Validation high-fidelity CFD/FE FSI for full-scale high-speed planing hull with composite bottom panels
slamming”, Abstract submitted International Conference on Coupled Problems in Science and Engineering,
San Servoro Island, Venice, Italy, 2015.
45. Wang, Z., Suh, J., Yang, J., and Stern, F., “Sharp Interface LES of Breaking Waves by an Interface Piercing
Body in Orthogonal Curvilinear Coordinates,” 50th AIAA Paper, January 2012a.
46. Wang, Z., Yang, J., Stern, F., “A new volume-of-fluid method with a constructed distance function on
general structured grids,” Journal of Computa-tional Physics, Vol. 231, Issue 9, 2012b, pp. 3703-3722.
47. Wang, Z., Yang, J., Stern, F., “A simple and conservative operator-splitting semi-Lagrangian volume-of-
fluid advection scheme,” Journal of Computational Physics, Vol. 231, Issue 15, 2012c, pp. 4981-4992.
48. Wang, Z., Yang, J., Stern, F., “High-Fidelity Simulations of Bubble, Droplet, and Spray Formation in
Breaking Waves”, HPC Insights, 2012d, Fall Issue.
49. Wang, Z., Yang, J., Stern, F., “High-Fidelity Simulations of Bubble, Droplet, and Spray Formation in
Breaking Waves,” 30th Symposium on Naval Hydrodynamics, Hobart, Tasmania, Australia, 2-7 November
2014.
50. Wu, P.C, Sadat-Hosseini H., Toda, Y., Stern, F., “URANS Studies of Ship-Ship Interactions in Shallow
Water”, 3rd Intl Conf. on Ship Manoeuvring in Shallow and Confined Water, Belgium, 2013.
51. Xing, T. and Stern, F., “Factors of safety for Richardson extrapolation”, J. of Fluids Engineering, Vol. 132,
2010.
52. Yang, J., Bhushan, S., Suh, J., Wang, Z., Koo, B., Sakamoto, N., Xing, T., Stern, F., Large-eddy simulation
of ship flows with wall-layer models on Cartesian grids, Proc. 27th Symposium on Naval Hydrodynamics,
2008, Seoul, Korea.
53. Yang, J. and Stern, F., A sharp interface direct forcing immersed boundary approach for fully resolved
simulations of particulate flows, ASME Journal of Fluids Engineering, 136(4), 040904 (10 pages), 2014.
54. Yang, J. and Stern, F., Robust and efficient setup procedure for complex triangulations in immersed
boundary simulations, ASME Journal of Fluids Engineering, 135(10), 101107(11 pages), 2013.
55. Yang, J. and Stern, F., A simple and efficient direct forcing immersed boundary framework for fluid-
structure interactions, Journal of Computational Physics, Vol. 231 (15), pp. 5029-5061, 2012.
56. Yang, J. and Stern, F., Sharp interface immersed-boundary/level-set method for wave-body interactions,
Journal of Computational Physics, Vol. 228 (17), pp. 6590-6616, 2009.
57. Yang, J., Preidikman, S., Balaras, E., A strongly-coupled, embedded-boundary method for fluid-structure
interactions of elastically mounted rigid bodies, Journal of Fluids and Structures, 24, pp. 167-182, 2008.
58. Yeon, S., Yang, J. and Stern, F., 2013. Large eddy simulation of drag crisis in turbulent flow past a circular
cylinder. ITTC workshop on wave run-up and vortex shedding, Nantes France, 17-18 October 2013.
59. Yoon, H., Gui, L., Bhushan, S. and Stern F. “Tomographic PIV Measurements For Surface Combatant 5415
Straight Ahead and Static Drift 10 and 20 Degree Conditions.” IIHR Report, 2014.
60. Yoon, S.H., Kim, D.H., Sadat-Hosseini, H., Yang, J. and Stern, F., 2013. High-fidelity CFD simulation of
wave run-up around vertical cylinders in monochromatic waves. ITTC workshop on wave run-up and
vortex shedding, Nantes, France, 17-18 October 2013.

15
Copyright © 2014 by ICHD
inlet

Figure 1: Turning maneuver simulation with water-jet propulsion (Sadat-Hosseini et al., 2013)

Figure 2: The grid topology and propeller vortices for KVLCC2 free running simulations with fully discretized propeller (Sadat-Hosseini et
al., 2014a)

(a)

15
EFD
CFD
10 SB_SI
y (m)

0
0 10
x (m)

(b) (c)
Figure 3: 5415M maneuvering simulations in calm water: (a) vortex structures for maximum drift angel during zigzag 1010; (b) free surface
profile for turning/pull-out; (c) CFD and SB trajectories compared with the experimental data (Sadat-Hosseini and Stern, 2014)

16
Copyright © 2014 by ICHD
Figure 4: The predicted transom free surface and vortex structures for turning maneuver simulation (Sadat-Hosseini et al., 2014b)

Figure 5: Course keeping simulation in irregular oblique waves (Sadat-Hosseini et al., 2014c)

EFD URANS DES

(a)

(b)
Figure 6: Overall vortical structures predicted by CFDShip-Iowa URANS (Middle) and DES (Right) predictions on adapted 84M grid for
5415 with bilge keels at (a) straight ahead and (b)  = 20 deg conditions.

17
Copyright © 2014 by ICHD
Variation QPeak Variation of TKECore
1000 6.5E-03 SDV: EFD
SDV: 84M URANS

Vortex Core TKE


FBKV: EFD
100 4.5E-03 FBKV: 84M URANS
Q

SDV: EFD
10 SDV: 84M URANS 2.5E-03
FBKV: EFD
FBKV: 84M URANS
1 5.0E-04
0 0.2 0.4 x/L 0.6 0.8 1 0 0.2 0.4 x/L 0.6 0.8 1
(a)
0.14
100000 Variation of QPeak Variation of TKEcore
0.12

0.1

Core TKE
10000
Q

0.08
SDTV: EFD
SDTV: 84M, DES
0.06
BKTV: EFD
SDTV: EFD
1000 0.04 BKTV: 84M, DES
SDTV: 84M, DES
BKTV: EFD
0.02
BKTV: 84M, DES

100 0
0 0.2 0.4 x/L 0.6 0.8 1 0 0.2 0.4 x/L 0.6 0.8 1
(b)
Figure 7: Variation of QPeak (LEFT) and TKE (RIGHT) at primary vortex cores predicted by best CFDShip-Iowa simulations are compared
with EFD data. (a) URANS predictions on 84M grid for sonar dome vortex (SDV) and fore body keel vortex (FBKV) cores is compared for
straight ahead case. (b) DES predictions on 84M grid for sonar dome tip vortex (SDTV) and bilge keel tip vortex (BKTV).

Figure 8: WAM-V CFDShip-Iowa regular head wave results in most probable conditions of SS3

18
Copyright © 2014 by ICHD
Figure 9: WAM-V hydrodynamic modeling for CFD simulations in regular head waves compared with EFD sea trials in random seas

Figure 10: WAM-V 2-post testing (left) and 6-post suspension simulation (middle) using CFD results as inputs. Comparison of payload
accelerations is shown (right) for 6-post simulation (yellow) and 2-post test data (purple)

Figure 11: CFD-MBD 2DOF 1-Way and 2-Way coupling results for cylinder drop compared with EFD data for pontoon (top) and sprung
mass (bottom) motions

19
Copyright © 2014 by ICHD
Figure 12: Underwater surface photo from EFD for USNA planing model at Fr=1.83 (Top) and comparison with CFDShip-Iowa predictions

Figure 13: USNA planing hull regular wave simulations using CFDShip-Iowa and NFA compared with experiments for: (a) Heave, (b) Pitch
and (c) Pressure at probe P13

20
Copyright © 2014 by ICHD
Figure 14: Irregular wave slamming pressures for USNA planing hull: (a) EFD slamming events aligned by peak pressure; (b) CFDShip-
Iowa validation showing expected value (EV) and +/- standard deviation (SD) bars for re-entering and emerging pressures and duration

Figure 15: CFD-FEA simulation results for FSI studies showing force and displacement distribution for the bay 4, port panel on the full-
scale Numerette planing vessel in regular head waves corresponding to sea state 3 most probable wave condition at Fr=0.7

Figure 16: Average, min and max of EFD (Experimental Fluid Dynamics) strain with its expected value (EV) and standard deviation (STD)
at peak compared with CFD/FE predicted strain for sea state 3 most probable wave condition and Fr=2.9

21
Copyright © 2014 by ICHD
Figure 17: Vortical structures with Q-criterion (left) and energy spectra of the streamwise velocity in the shear layer (right): sub-critical
(Re=1.26x105) (top), critical (Re=2.52x105) (center) and super-critical (Re=7.57x105) (bottom).

22
Copyright © 2014 by ICHD
Figure 18: Drag coefficient, RMS lift coefficient, separation angle, and base pressure vs. Re.

λ/D=4.7, λ/D=4.7, λ/D=4.7,


H/λ=1/30 H/λ=1/16 H/λ=1/10

λ/D=21.9, λ/D=21.9, λ/D=21.9,


H/λ=1/30 H/λ=1/16 H/λ=1/10

Figure 19: EFD and CFD comparison of mean wave field for single cylinder cases

23
Copyright © 2014 by ICHD
Mean 1st

h0
0.18
0.15
0.12
0.09
0.06
0.03
-0.01
-0.04
-0.07
-0.10

Mean 1st

Figure 20: Mean and 1st harmonic amplitude of the wave field for single cylinder (λ/D=4.7, H/λ=1/10) and four cylinder (λ/D=4.7, H/λ=1/30)
cases

Figure 21: Flow charts of three different strong coupling schemes in one time step for fluid-structure

24
Copyright © 2014 by ICHD
(a) (b)

(c) (d)

(e) (f)
Figure 22: Stokes wave breaking. (a) perspective view; (b) detailed 3D interface structures; (c) Görtler inviscid instabilities; (d)
Rayleigh/Taylor instabilities (red: stable; blue: unstable); (e) side view showing bubbles and droplets; and (f) bubble size distribution.

25
Copyright © 2014 by ICHD
(a) (b)

Figure 23: (a) Cavitating NACA66 hydrofoil with a 6° angle of attack; 2D solution showing cavity growth and shedding. (b) Close up
views of a sharp interface simulation of a cavitating NACA66 hydrofoil with a 6° angle of attack showing bubble growth, merging, and
advection shortly after inception at the leading edge.

benchmark (empirical) benchmark (Normal)


UQ (empirical) UQ (Normal)
0.16
0.14
0.12
0.1
PDF

0.08
0.06
0.04
0.02
0
-25 -20 -15 -10 -5 0 5
X 103

(a)
benchmark (empirical) benchmark (Normal) benchmark (empirical) benchmark (Normal)
UQ (empirical) UQ (Normal) UQ (empirical) UQ (Normal)
35 12

30 10

25
8
20
PDF
PDF

6
15
4
10

5 2

0 0
-0.08 -0.06 -0.04 -0.02 0 0.02 0.04 0.06 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 0.2 0.25
z/L q [rad]
(b) (c)
Figure 24: Comparison of time history distributions from irregular wave (benchmark) and regular wave UQ for the Delft Catamaran in head
waves, at sea state 6 and Fr=0.5. Empirical and Normal density functions are shown for (a) force X, (b) heave z/L, (c) pitch θ (He et al.,
2013a).

26
Copyright © 2014 by ICHD

View publication stats

Вам также может понравиться