Sei sulla pagina 1di 13

THE 18th CHESAPEAKE SAILING YACHT SYMPOSIUM

ANNAPOLIS, MARYLAND, MARCH 2007

RANSE Calculation of Laminar-to-Turbulent Transition-Flow around Sailing


Yacht Appendages

Christoph Böhm, R&D-Centre Univ. Applied Sciences Kiel, Yacht Research Unit, Germany
Kai Graf, Institute of Naval Architecture, University of Applied Sciences Kiel (UAS), Germany

K flow acceleration coefficient


ABSTRACT L reference length
A new method to simulate laminar-turbulent p pressure
transition has been used to study flow around sailing yacht PK production term of k equation
~
appendages. Based on an empirical correlation, the method PK modified production term of k equation
allows to predict transitional flow in a fully 3D- Pγ1 production term of transition sources
environment using unstructured grids and massive Pγ2 production term of relaminarization sources
parallelization. Reynolds number of the flow around the ~ ~
PΘ,t source term to make Re Θ,t match ReΘ,t
appendages of sailing yachts is of an order where
transitional flow plays an important role. The method thus Pω production term of ω equation
serves as a highly valuable tool for realistic optimisation of Re Reynolds number LU∞/ν
yacht appendages. Rec critical Reynolds number
ReΘ momentum thickness Reynolds number ΘU0/ν
The paper describes briefly Menters γ-ReΘ- ReΘ,c critical momentum thickness Reynolds number
transition-model and shows its validity for the given ReΘ,t transition onset momentum thickness Reynolds
purpose. 2D-analysis of NACA-profiles has been chosen to number ΘtU0/ν
validate the transition model. The resulting lift and drag ~
Re Θ,t local transition onset momentum thickness
coefficients from the RANSE calculations have been
compared with experimental data and results from the 2D- Reynolds number
boundary layer code XFoil, showing reasonable RT viscosity ratio
agreement. The method then has been used for a bulb Rey Reynolds number based on wall distance
length optimisation of an appendage configuration of an Reν vorticity Reynolds number
ORC GP42 level racer. Here the differences between a Sij strain rate tensor
fully turbulent and a laminar turbulent optimisation are S absolute value of strain rate, wetted surface
shown and discussed. Tu turbulence intensity
U∞ freestream velocity
ui local velocity component
NOTATION U0 velocity outside of boundary layer
y wall distance
(1+k) form factor y+ dimensionless wall distance
ARef reference surface
γ Intermittency
cD drag coefficient
cF,Hughes friction line according to Hughes
δ boundary layer thickness
D drag force δ* displacement thickness
DK destruction term of k equation δij Kronecker delta
~ ε turbulence dissipation rate
DK modified destruction term of k equation
θ boundary layer momentum thickness
Dω destruction term of ω equation λ0 pressure gradient coefficient
Eγ1 destruction term of transition sources μ =ν ρ
Eγ2 destruction term of relaminarization sources μt =νT ρ
gi component of mass normalized body force μτ friction velocity
k turbulent kinetic energy
ν kinematic viscosity
νT Turbulent viscosity determination of optimal geometric properties of the
ρ Density ballast bulb.
σi constants in diffusion coefficients
τ wall shear stress The following chapters give an introduction to the
τij Reynolds stress tensor theory of both transitional flow and γ-ReΘ-Turbulence
ω specific turbulence dissipation rate model and finally two examples which demonstrate the
Ω Absolute value of vorticity capability of the method to predict laminar-to-turbulent
Ωij vorticity tensor transition.

INTRODUCTION TRANSITIONAL FLOW


RANSE simulations of flow around sailing yacht Also the complex mechanism of transition from
appendage systems are nowadays common practice in the laminar to turbulent flow, as well as turbulence itself is still
design stage of high performance sailing yachts. They are not fully researched, it is known, that the transition from
used by many flow scientists and yacht designers to laminar to turbulent flow is caused by viscous instabilities,
optimize the shape of yacht keels, ballast bulbs, rudders which are developing in the shear layer. These instabilities
and wings. Modern RANSE codes offer great flexibility always increase downstream and depend only on the
taking into account viscosity, turbulence and free surfaces, Reynolds number. At the beginning of the transition
providing results that can compete with towing tank and process turbulent and laminar regions are mixed in the
wind tunnel test results. boundary layer. The degree of this mixing is defined as the
intermittency factor γ. Figure 1 shows a scheme of the
A common procedure in the analysis of flow transition process obtained from experiments on a flat
around appendages is to assume fully turbulent flow plate.
allowing to use conventional turbulence models. However
the Reynolds number of typical yacht appendages is of an
order of magnitude where large laminar regions can be
expected. In fact designers usually try to extend the
laminar region of the appendage set by suitable form
modelling to decrease viscous resistance. Consequently
flow analysis not taking into account laminar to turbulent
transition may lead to erroneous design results.

Late developments of turbulence models address


this problem. The most promising of these enhanced
turbulence models, the γ-ReΘ-Turbulence model which is
implemented in the commercial RANSE solver CFX, uses
a correlation-based transition model to predict production
and dissipation of turbulent kinetic energy and wall shear
stress, MENTER (2004). The correlation used in the
transition model is based on experiments to correctly
describe the transition location. The principal idea behind
this method is to compare free stream turbulence intensity Fig. 1: By-Pass-Transition on a flat plate
with the local Reynolds number based on boundary layer
momentum thickness. Integral quantities needed for this In case of low levels of turbulence intensity in the
transition criteria are approximated by local variables to free stream, a laminar boundary layer is developing
avoid numerical problems within a RANSE code. downstream of the leading edge of the flat plate. This
laminar region extends until x=x(Rec) is reached, where
This transition model is used in a research project the first instabilities occur as so called Tollmien-
at YRU-Kiel to investigate the laminar-to-turbulent Schlichting (T-S) waves. Shortly afterwards, the mainly
transition flow around the appendages of an ORC 42 level 2D T-S waves start to develop 3D fluctuations of speed
racer. The achieved results are compared with results for and pressure, which open out into longitudinal eddies, so
fully turbulent flow. The ORC 42 has been used as a called Λ-vortices. The breakdown of these vortices at
benchmark since it features a small fin-keel and a medium locations of high vorticity results in the loss of the natural
length ballast bulb, where laminar flow can be expected on frequency of the T-S waves and the developing of high-
about 50% of it’s wetted surface. Goal of this investigation frequency vortices. These vortices break down into even
is to show whether the conventional assumption of fully smaller vortices, leading to the growth and forming up of
turbulent flow is appropriate or if it may lead to wrong
wedge-shaped turbulent spots. Turbulent spots move equation depending on the transition onset
downstream with the surrounding laminar boundary layer momentum thickness Reynolds number
and lead to an alternation between laminar and turbulent • The intermittency function is used to manipulate
flow, the intermittency as mentioned above. The production and dissipation of turbulent kinetic
unification of the turbulent spots finally leads to a fully energy.
turbulent boundary layer after exceeding the transition
Reynolds number Ret. The following chapter outlines the basic equations
of this model and its integration into common RANSE
Transition to turbulence is mainly influenced by: methods.

• Local Reynolds number


• Free stream turbulence intensity
• Pressure gradient FLOW SIMULATION METHOD
• Surface roughness The theory given here is a very brief outline of
• Separation Menters method. For details see the paper of
MENTER (2004).
For common yacht appendages the transition
length is usually much shorter than the overall length of
the appendage element. Consequently accurate prediction Governing Equations
of transition length is of less importance than prediction of RANSE solver use a volume based method to
transition onset. Engineering methods for the prediction of solve the time-averaged Navier-Stokes equations in a
transition thus usually do not take into account the details computational domain around the investigated body. The
of the transition process. They rather predict the critical RANS equation evolves from time averaging mass and
Reynolds number Rnc and the transition Reynolds number momentum conservation for a continuous flow. In the
Rnt using empirical methods. following method it is assumed that the Reynolds stress
evolving from time averaging is modelled using the eddy
Most of these empirical methods are based on the viscosity hypothesis. Assuming incompressible flow this
assumption that transition onset is linked to the Reynolds yields, see FERZINGER (1991):
number based on the momentum thickness of the boundary
∂ui ∂ (u j ui )
layer. Empirical equations for the transition onset
momentum thickness Reynolds number are given by many 1 ∂( p + 2 / 3 ρ k )
+ =−
researchers. They take into account parameters describing ∂t ∂x j ρ ∂xi
details of the flow as pressure gradient and freestream (1)
⎛ ∂u ⎞
turbulence intensity. +
∂ ⎜ (ν + ν T )( ∂ui + j ) ⎟ + g i
∂x j ⎜ ∂x j ∂xi ⎟⎠

The general problem of implementing one of
these methods within RANSE codes is that RANSE ∂ui
simulations use local information only for solving =0 (2)
∂xi
governing equations. Derivation of integral values, which
are needed for determining the momentum thickness, is
troublesome and should be avoided. Turbulence Model

Menters method to circumvent these problems is The turbulence model used in the presented
based on a couple of ideas that allow him to calculate approach calculates the turbulent viscosity νT from the
transition from local variables only. His approach covers turbulent kinetic energy k and the specific turbulent
many of the different aspects of transition:A new empirical dissipation ω:
formulation for the transition onset momentum thickness
Reynolds number in the freestream is given. a1k
νT = (3)
max (a1ω, ΩF2 )
• A transport equation is used to diffuse the
freestream transition onset momentum thickness
The turbulent kinetic energy k and the specific
Reynolds number into the boundary layer.
turbulent dissipation rate ω are calculated using the γ-ReΘ-
• The momentum thickness Reynolds number is
Turbulence model shown below.
correlated with the vorticity Reynolds number
which can be calculated from local variables
• ∂ (ρk ) ∂ (ρui k ) ~ ~ ∂ ⎛ ∂k ⎞
An intermittency function is calculated from a + = Pk − Dk + ⎜⎜ (μ + σ k μt ) ⎟⎟ (4)
transport equation, the source term of this ∂t ∂xi ∂xi ⎝ ∂xi ⎠
∂ (ρω ) ∂ (ρuiω )
+
P
= α k − Dω + Cd ω ( ~
∂ ρ Re Θ,t ) + ∂ (ρu R~e ) = P
i Θ ,t
+
∂ ⎛
⎜ σ Θ,t (μ + μ t )
~
∂ Re Θ,t ⎞

∂t ∂xi νt ∂t ∂xi
Θ ,t
∂xi ⎜ ∂xi ⎟⎠
(5) ⎝
∂ ⎛ ∂ω ⎞
+ ⎜⎜ (μ + σ k μt ) ⎟ ( 11 )
∂xi ⎝ ∂xi ⎟⎠
~
Here the source term PΘ,t ensures that Re Θ,t
The γ-ReΘ-Turbulence model is equivalent to the matches the local value of ReΘ,t outside the boundary layer,
standard SST-Model besides having modified production the latter one obtained from ( 8 ). The source term Pθ,t
~ ~
and destruction terms ( Pk , Dk ) in the turbulent kinetic contains a blending function, which turns off the source
~ ~ term in the boundary layer.
energy equation. Pk and Dk are calculated from
production and destruction of turbulence in fully turbulent The transition onset momentum thickness
regimes, Pk and Dk, and an intermittency function γ which ~
Reynolds number in the boundary layer, Re Θ,t , is related to
controls production and destruction of turbulent kinetic
energy depending on the transition onset momentum the critical momentum thickness Reynolds number Reθ , c ,
thickness Reynolds number: which is the Reynolds number where first instabilities in
the laminar boundary layer occur. Unfortunately this
~
Pk = γ Eff Pk (6) relationship is proprietary and a hidden procedure within
Menters method:
~
Dk = min(max(γ Eff ,0.1),1.0) Dk (7) ~
(
Reθ , c = f Re Θ,t ) ( 12 )

The critical momentum thickness Reynolds


Transition-Model
number Reθ , c is now forwarded to the intermittency
The transition model assumes that the transition
onset momentum thickness Reynolds number outside the transport equation, which controls the production and
boundary layer can be calculated by an empirical formula. destruction of turbulence in the turbulence model:
Menter suggests a new empirical correlation for ReΘ,t
which relates it to the freestream turbulence intensity Tu ∂ (ργ ) ∂ (ρu i γ ) ∂ ⎛⎛
⎜ ⎜ μ + μt
⎞ ∂γ



+ = Pγ 1 − E γ 1 + Pγ 2 − E γ 2 + ( 13 )
and two empirical correlation factors λ0 and K. The ∂t ∂x i ∂x i ⎜⎜ σf ⎟ ∂x i ⎟
⎝⎝ ⎠ ⎠
correlation is defined as:
In this transport equation, Pγ1 and Eγ1 depict the
Re Θ,t = 803.73[Tu + 0.6067] F (λ0 , K )
−1.027
(8) transition sources necessary to start the production of
turbulence whereas Pγ2 and Eγ2 are the destruction or
⎛ Θ 2 ⎞ dU relaminarization sources.
λ0 = ⎜⎜ ⎟⎟ (9)
⎝ ν ⎠ ds Pγ1 is controlled by an onset function which
depends on the ratio of the local momentum thickness
⎛ ν ⎞ dU Reynolds number Re Θ and the critical momentum
K =⎜ 2 ⎟ ( 10 )
⎝ U ⎠ ds thickness Reynolds number Reθ , c . To avoid the
calculation of the non-local momentum thickness Reynolds
The function for λ0 shows strong similarities to a pressure number Menter uses a correlation between momentum
gradient method proposed by Thwaites, whereas the thickness Reynolds number Reθ and the local vorticity
second parameter K is the so called acceleration parameter
Reynolds number Reν:
similar to the method proposed by Mayle, for details see
WHITE (1991). The acceleration parameter K is used to
Reν ,max
characterize the accelerations at the beginning of the Re = ( 14 )
transition process. Function F(l0,K) is used to decide if a Θ 2.193
polynomial function depending on λ0 or K is used to
determine ReΘ,t. Reν is based on local values only and can
therefore easily determined at each computational node:
A local value of the transition onset momentum
thickness Reynolds number is calculated from a transport ρy 2
Reν = Ω ( 15 )
equation, assuming that the freestream value of Reθ ,t , μ
calculated from ( 8 ), diffuses into the boundary layer:
APPLICATIONS
In the following the method of Menter is validated
by comparing it with respective experimental and
numerical results from other sources. It is then used to
investigate the flow around the appendage set of an ORC
GP 42 level racer. While the first investigation assumes
planar flow, the appendage flow simulation is fully 3D.
The original empirical methods to predict transition from
the momentum thickness Reynolds number are 2D by
nature. However due to the substitution of integral values
with local variables fully 3D flow can be investigated with
Menters method. In addition, since the local transition
onset momentum thickness Reynolds number in the
~
boundary layer, Re Θ,t , is calculated from diffusion of the
freestream transition onset momentum thickness Reynolds Fig. 2: Grid around NACA 642-015 profile
number, in-homogeneity of turbulence intensity in the
freestream can be taken into account. This allows to take Fig 3 shows the drag coefficient cD over angle of
account of interaction of appendage elements like bulb and attack AoA for RANSE calculation (both laminar-turbulent
blade or even blade and rudder. and fully turbulent), XFOIL calculations and experimental
data. Drag coefficient cD is calculated by dividing the drag
2D TEST CASE: NACA 642-015 PROFILE force obtained from the computation with dynamic
A profile of the well researched NACA 6-series pressure and the planform area.
was chosen for analysis to verify the capability of the γ-
ReΘ-Turbulence model to predict transition. The results of D
cD = ( 16 )
the profile analysis have been compared with experimental 0.5 ⋅ ρ ⋅ U ∞2 ⋅ ARe f
data collected from Abbott/Doenhoff, ABBOTT (1959)
and numerical results generated by the well known profile
code XFOIL of DRELA (2001). 2.0E-02
CD [-]

1.8E-02
The simulation has been performed using a
hexahedral mesh with C-grid topology and approx. 180000 1.6E-02
computational nodes. As an important pre-conditions to
achieve accurate results using the γ-ReΘ-Turbulence model 1.4E-02

the dimensionless wall distance has to be restricted to Y+ ≤ 1.2E-02


1. Thus the wall nearest node of the grid has to be placed at
a distance of approximately 0.003 mm from the surface to 1.0E-02
achieve correct results for a Reynolds number of 3*106.
8.0E-03
Consequently great care has to be taken in the design of
the profile surface, which has to be very smooth to allow 6.0E-03
proper gridding of the cells in the vicinity of it.
4.0E-03
Fig. 2 shows the hexahedral mesh, the inset CD XFOIL CD CFX
2.0E-03
showing grid resolution at the profile leading edge. CD EXP CD CFX Turb

0.0E+00
A sweep of angle of attacks from 0° to 10° have 0 2 4 AoA [°] 6 8 10
been tested at constant inflow speed. Turbulence intensity
in the freestream has been set to 0.85%, which corresponds Fig. 3: Drag coefficient cD for NACA 642-015
to an Ncrit-Value of 3.012 in the XFoil settings. The exact
Turbulence level at which the experiments were conducted Generally the RANSE results show rather good
in the wind tunnel is unknown, however Abbott/Doenhoff compliance with both numerical and experimental
mentioned it to be in the order of a few hundredths of one comparative data, in particular close to the non-lifting
percent. condition AoA=0°. This is a good indicator for proper
laminar to turbulent transition prediction. Here the results
from the new transition model almost coincides with the
experimental data at angles from 0 to 2 degrees. Compared
to the experimental results the beam of the drag bucket is
narrower for the RANSE results, however comparing the
RANSE results with those from XFOIL, the trend is
inversed. Additional experimental results are needed here
to evaluate the merit of RANSE compared to XFOIL
results.

For wider angles of attack, the curves for the laminar-


turbulent RANSE calculations show a different slope than
the comparative data and quickly approach the results from
the fully turbulent RANSE calculation. In this region the
need for further research in order to determine possible
flaws in the calculation setup and to calibrate the model is
obvious. It is also worth considering that the quality of the
experimental and numerical comparative data itself is
unknown.

Fig. 4 compares transition points on top and bottom of the


surface resulting from numerical calculations with CFX
and XFoil. It can be nicely seen that the curve slopes are
very similar. Given the fact that XFOIL predicts narrower
drag bucket than the RANSE calculations, the horizontal Fig. 5: Wall Shear Stress on NACA 642-015 profile:
shift of the top transition points are very plausible.
a) fully turbulent flow b) laminar-turbulent flow

1 Here the wall shear stress progression shows the


0.9 Xtr/c Pressure Side XFOIL typical characteristic of laminar to turbulent transition: a
Xtr/c Suction Side XFOIL sharp increase close to the stagnation point followed by a
Xtr/c Pressure Side CFX
0.8
Xtr/c Suction Side CFX distinct shear stress decrease over almost 60% of the chord
0.7 length. The transition to turbulent flow can be identified at
the location where shear stress starts to increase again
0.6
Xtr/c [-] indicating turbulent flow in the backward region.
0.5

0.4
BULB LENGTH VARIATION
0.3 As a practical application a length variation of a
0.2 keel bulb configuration for an ORC GP 42 level class racer
will be presented here. When trying to predict the optimum
0.1 length for a ballast bulb of given buoyancy the following
0 problem occurs:
0 2 4 AoA 6[°] 8 10
Bulbs optimised using fully turbulent CFD codes
Fig. 4: Transition Points on upper and lower profile are known to become very long and slender, whilst bulbs
surface as predicted by CFX and XFOIL which are optimised using methods capable to take into
account laminar turbulent transition (like wind tunnel
In general the agreement between the different testing) tend to be too bulky. The reason for this is, that for
methods compared here seems to be reasonable. This holds fully turbulent flow a prolongation of the bulb increases
particularly for a situation, where design input from an frictional resistance only slightly as can be derived from
optimisation has to be generated. Comparing variants of Fig. 5. For a bulb with a laminar region at the forward part
topologically similar geometries like ballast bulbs of any prolongation increases only the region where turbulent
different volume distribution or length will give quite flow prevails, resulting in an over-proportional increase of
reasonable trends as long as the simulation set-up is not resistance. Of course the bulb resistance in fully turbulent
changed. flow is generally higher than in laminar turbulent transition
flow.
Fig. 5 shows porcupine-plots representing the
wall shear stress on the profile. The upper plot depicts a For comparison the bulb length investigation is
profile in fully turbulent flow, visible by a smoothly carried out for laminar turbulent transition flow as well as
decreasing distribution of wall shear stress in flow for fully turbulent flow. This allows to estimate the error
direction. In the lower plot the result for a laminar- margin, giving a measure of the merits bound to the use of
turbulent transition simulation is shown. the presented method.
Geometry The bulbs come completely without chines and
No complete design of a GP 42 racer has been have an asymmetric profile to keep the centre of gravity
available at the YRU Kiel. Consequently the blade and the low. This results in a rather flat bottom part.
benchmark bulb have been developed according to the
ORC GP42 level class rules with dimensions likely to be
used in reality. From estimation of blade dimensions and Simulation Setup
weight a benchmark bulb with the following characteristic The simulation has been set up in a boxed
properties has been derived: environment similar to those encountered in wind tunnels,
with the following differences: The box walls are treated as
CSYS 18 Benchmark frictionless free slip walls and the appendage configuration
Length L [m] : 2.400 is tested in full scale, thus eliminating any problems with
Wetted Surface S [m²] : 2.215 Reynolds similarity. The boundary conditions applied are a
Vertical Center of Gravity VCB [m] : 0.154 constant inflow speed of 5.144 m/s on the inlet, which is
Long. Center of Gravity LCB [m] : -0.323
corresponding to a speed of 10 knots. On the outlet von
Volume Vol [m³] : 0.170
Neumann condition (derivation of flow forces
Height H [m] : 0.326
Width B [m] : 0.482
perpendicular to the boundary is zero) is applied. The bulb
B/H Ratio SQR [m] : 1.480 is investigated in downwind condition only, the free stream
Maxium Draft MD [m] : 2.300 turbulence intensity has been set to 0.85% as during the
Tab. 1: Geometric properties of the benchmark bulb profile investigation.

The bulb variants are derived from the benchmark All geometries have been simulated using two
geometry by affine distortion of the bulb length, height and different gridding schemes: A hexa-tetra grid uses
width. During the study the bulb volume and its squish hexahedral cells in the vicinity of the keel to accurately
ratio (B/H-ratio) as well as the maximum draft remains resolve the boundary layer, while a tetrahedral grid
constant. The following variants have been developed: topology has been used for the far field in order to restrict
the necessary number of grid cells. In a tetra-prism grid
L [m] S [m²] VCB [m] H [m] B [m] Vol [m³] the boundary layer flow is resolved using flat prisms while
CSYS 18_V1 1.900 1.988 0.173 0.367 0.543 0.170 the rest of the flow domain is discretized using tetrahedral
CSYS 18_V2 2.150 2.104 0.164 0.345 0.510 0.170 cells. The first method consists of an inner box around
CSYS 18_V3 2.650 2.323 0.147 0.310 0.459 0.170 blade and bulb, which is meshed with a multi-block
CSYS 18_V4 2.900 2.427 0.140 0.297 0.439 0.170 structured hexahedral mesh. The far field (outer box) is
CSYS 18_V5 3.150 2.527 0.135 0.285 0.422 0.170 meshed with a relatively coarse tetra mesh to save
Tab. 2: Geometric properties of the bulb variants computational time. The hexahedral and the tetrahedral
grida are finally connected with a node-matching interface
Fig. 6 shows bulb variants in profile and sectional via pyramid cells.
view. One can see that the variants derived from the rather
moderate benchmark become more and more radical with
increasing respective decreasing length.

Fig. 7: Bulb and blade (blue) with surrounding hexa-tetra


hybrid grid

Fig. 6: Overview of the bulb variants The second meshing approach is a fine
unstructured tetrahedral grid with prism layers at the
investigated surfaces for a refined resolution in the
boundary layer. This method has the advantage that it can It has to be mentioned that the meshing effort for
be easily adapted to any geometry, but comes with the laminar-turbulent calculations is huge, since the grid
disadvantages that the task of resolving the correct generator has to capture the intersections between bulb and
boundary layer height is not as easily performed as with blade with the same high resolution as with the profile
the hexahedral mesh and the computational effort is bigger. investigation. Therefore it is even more important than for
the 2 D test case to invest great care in developing smooth
laminar-turbulent fully turbulent and clean surfaces.
Hexa-Tetra- Tetra-Prism Hexa-Tetra- Tetra-Prism
Elements Hybrid Grid Grid Hybrid Grid Grid
Hexahedral 2784428 - 1512944 -
Tetra 828109 992155 1421511 992155 Results Hexa-Tetra Grid
Prism - 4350902 - 869978 The computational results achieved with the hexa-
Pyramids 50656 158 41152 66 tetra grid suffer from a most unexpected drawback: a
Σ Elements 3663193 5343215 2975607 1862199
Σ Nodes 2909960 2358636 1745740 618495
review of the form factors shows that the pressure drag of
Tab. 3: Grid parameters the bulb is heavily underestimated. The form factor is
defined as follows:
Both grid variants have been developed with
approximately the same grid resolution, see Tab. 3. The (1 + k ) = cD
( 17 )
tetra-hexa grid has a computational load which is around c F , Hughes
half a million nodes greater than with the tetra-prism grid.
This mainly owed to the fact that the tetra-prism grids where cD is defined as in formula 14, but normalized using
allows for more aimed node distribution and can therefore the wetted surface instead of the planform area. The
save some nodes, whilst hexahedral grids always come coefficient cF,Hughes resembles the friction line as introduced
with some constraints that cause the grid size to increase. by Hughes.
The grids for the fully turbulent simulations are 0.063
generally around 40% (hybrid) – 70% (tetra-prism) smaller c F , Hughes = ( 18 )
than the laminar-turbulent grids. This is due to high (log(Re) − 2)2
resolution needed in the boundary layer for the γ-ReΘ-
Transition model to work correctly. The differences A length variation is usually an interplay between
between the prism elements in fully turbulent and laminar the viscous drag, mainly depending on the wetted surface,
unstructured grids depict this very well, see Tab. 3. and the pressure drag, depending on the form of the
Naturally this also enlarges the differences in node investigated body. A massive pressure underestimation
numbers between the two grid approaches developed for leads to a very blunt bulb, this being the body with the
fully turbulent flow. least wetted surface. Obviously these results are useless as
they do not resemble reality.

1.20
Drag / DragBench
F/FBench, Vcg/VcgB[-]

1.15 VCGBulb/VCGBench
Viscous Drag / VDragBench
Pressure Drag / PDragBench
1.10

1.05

1.00

0.95

0.90

0.85
Fig. 8: Unstructured tetra grid with prism layer on bulb 1.8 2 2.2 L [m] 2.4 2.6 2.8 3
and blade surfaces
Fig. 9: Results of fully turbulent length variation with
hexa-tetra-hybrid grid
Fig. 9 and Fig. 10 show the results for the fully did not appear here and the ratio of pressure and viscous
turbulent and the laminar- turbulent test run with the drag looks very realistic.
hexahedral grid. All results have been modified in a way
that they depict the difference to the benchmark geometry Fig. 11 shows results for fully turbulent
in percent. Besides total drag the curves for the resistance calculations using the tetra-prism grid. Here an optimum of
components viscous and pressure drag, as well as the total drag was acquired for a bulb length of 3.15m, which
development of the bulb vertical centre of gravity are also is +0.75m longer than the benchmark bulb. The optimum
given to show the different trends. for the fully turbulent calculation is rather flat, however it
fits well with theory and common knowledge that this kind
1.15 of simulation does tend to produce long and needle-like
bulbs.
F/FBench, Vcg/VcgB [-]

1.10 1.20
Drag / DragBench

F/FBench, Vcg/VvgB[-]
VCGBulb/VCGBench
1.05 1.15
Viscous Drag / VDragBench
Pressure Drag / PDragBench

1.00 1.10

0.95 1.05

0.90 Drag / DragBench


1.00
VCGBulb/VCGBench
Viscous Drag / VDragBench
Pressure Drag / PDragBench
0.85 0.95
1.8 2 2.2 2.4 L [m] 2.6 2.8 3

Fig. 10: Results of laminar turbulent length variation with


0.90
hexa-tetra-hybrid grid

In case of the fully turbulent length variation an


0.85
optimum was found, at a length of 2.15m, this being 0.25m
1.8 2.0 2.2 2.4 L [m] 2.6 2.8 3.0 3.2
smaller than the benchmark bulb. This resulted in a bulb
which was bulkier than one would expect for a fully Fig. 11: Length Variation with respect to Benchmark
turbulent calculation. The calculations taking laminar- Geometry CSYS18_BM (Fully Turbulent Mode)
turbulent transition into account did not lead to an
optimum at all. Here the result was that any shortening of The results of the simulations conducted with the
length leads to decrease of resistance. new laminar-turbulent-transition model are shown in Fig.
12. As to be expected, the pressure drag increases with
A thorough investigation of the simulation lead to decreasing bulb length due to the bulkier shape of the
the conclusion that the problems have to be grid related. body. Simultaneously the viscous drag decreases because
Whether the problems arose from the mixing of tetra and the wetted surface of the body gets smaller. The optimum
hexa grid, from the blocking structure or from the node distribution of pressure and viscous drag is reached for a
distribution could not yet be resolved. Currently we bulb length of 2.15m. This corresponds to a delta of -
assume the pressure underestimation to be related to 0.25m to the benchmark. The optimum bulb length for a
distortion of the hexa grid caused by the use of a bulb in laminar turbulent flow is about 1 m shorter than for
unconventional blocking strategy. a bulb in fully turbulent flow.

Results Unstructured Tetra-Prism Grid


The results received from the simulations with the
second meshing approach, using the tetra-prism grid, are
by far more encouraging. The problems of massive
pressure underestimation which occurred with the first grid
1.25 flow on the bulb top to transit earlier to turbulence thus
Drag / DragBench overriding the natural transition process.
VCGBulb/VCGBench
F/FBench,Vcg/VcgB [-]

1.20 Viscous Drag / VDragBench Fig. 14 and Fig. 15 depict wall shear stress for the
Pressure Drag / PDragBench benchmark bulb for fully turbulent flow and for laminar-
1.15 turbulent transitional flow. Fig. 14 shows shear stress
increasing from zero at the nose to maximum somewhere
behind mid section. Afterwards the wall shear stress
1.10
decreases towards the end.

1.05

1.00

0.95

0.90

0.85
1.8 2 2.2 L [m] 2.4 2.6 2.8 3

Fig. 12: Influence of Length Variation with respect to


Benchmark Geometry CSYS18_BM (Laminar-Turbulent-
Transition Model)

Fig. 13 shows longitudinal location of transition


onset for the bulbs. The transition locations are measured
Fig. 14: Contour plot of wall shear stress on benchmark
at bottom, top and side of the bulb and normalized with the
bulb (fully turbulent)
bulb length. An increase of bulb length decreases the
longitudinal position of the transition location quite
smoothly towards the bulb nose. The sole exception here is Fig. 15 shows a similar contour plot for laminar-
bulb variant V3 which shows a unexpected peak on the turbulent transition flow around benchmark bulb. Shear
bulb bottom. stress is characterized by a sudden increase of wall shear
stress on the nose, followed by a swift decrease to a very
low level.
1.0
Xtr/c Top
Xtr/c [-]

Xtr/c Bottom
0.8
Xtr/c Side

0.6

0.4

0.2

0.0
1.8 2.1 L [m] 2.4 2.7 3.0

Fig. 13: Transition Points on bulb calculated with the γ-


ReΘ-Transition model
By comparing the results for the bulb top, bottom
and side, one can see that the changes are most distinctive
for the bulb top. This seems to be plausible since the top of Fig. 15: Contour plot of wall shear stress on benchmark
the bulb geometry is intersecting with the blade. Corner bulb (laminar- turbulent)
vertices as well as transition on the blade may cause the
This low level of shear stress is maintained until
the critical Reynolds number Rec is reached. From here the
wall shear stress increases to almost the same value as with
the fully turbulent calculation. It then follows the same
pattern as described for the fully turbulent calculation and
decreases towards the beaver-tail end of the bulb.

Fig. 16, Fig. 17 and Fig. 18 show turbulent kinetic


energy on the surface of the benchmark bulb, the V2
variant (very short bulb) and the V4 variant (very long
bulb) respectively. The images unveil the development of
laminar and turbulent zones for the three selected bulb
lengths. Here blue colour regions indicate laminar flow,
whilst regions of turbulent flow are coloured red. The
bulbs shown here are the benchmark bulb, the optimum-
length bulb for laminar turbulent transition flow (V2) and
the optimum-length bulb for fully turbulent flow (V4).
Fig. 17: Turbulent kinetic energy on the V2 bulb

Fig. 18 shows the optimum-length bulb for fully


turbulent flow. One can see that the flow instabilities on
the upper part of the bulb surface are not strong enough to
force a transition on the blade surface. Instead the laminar
flow on the blade seems stable enough help in stabilizing
the laminar regions at the bulb regions near to the blade
junction. The comparatively weak instabilities on the bulb
top are probably owed to the position of maximum
thickness, which moves further aft for the longer bulbs.

Fig. 16: Contour plot of turbulent kinetic energy on the


benchmark bulb indicating laminar (blue) turbulent (red)
flow

Flow at the bulb-blade junction of the benchmark


bulb transits relatively early to turbulent. By looking at the
transition course at blade and bulb, it seems that
instabilities of the flow on the bulb force the flow on the
blade to transit to turbulence forward of the natural
transition location.
Fig. 18: Turbulent kinetic energy on the V4 bulb
The optimum bulb for laminar-turbulent transition
flow (Fig.17) possesses several differences compared to Fig.19 shows turbulent kinetic energy at bulb
the benchmark bulb. Noticeable is the rather late transition bottoms for the entire range of investigated bulb lengths.
of laminar to turbulent flow on the entire bulb surface. The By comparison the different lengths of laminar and
laminar-turbulent transition at the upper part of the bulb turbulent regions can be detected easily.
surface is probably triggered by the flow instabilities on
the blade. However, late transition at the bulb side may be It can be generally stated that the flow on the top
a consequence of higher curvature and thus negative and bottom side of the bulb transits earlier to turbulent
pressure gradients compared to the benchmark. flow than at the bulb sides. Furthermore, the contour of the
transit regions varies from a more or less even sectional
distribution around the shortest bulb (V1) towards a wedge
shape turbulent peak at bottom of the longest bulb (V4.)

This phenomena is surely owed to the geometry


of the bulb which was designed rather flat with a B/H ratio
of 1.48 in order to keep the centre of gravity low.
Additionally the bulb bottoms are flatter than the bulb tops.
This also enhances hydrostatic stability, but is not suited to
stabilize the laminar flow.

Fig. 20: Boundary layer profile at transition location

Conclusion
In this paper a new laminar-turbulent transition
model has been tested for its ability to predict laminar-to-
turbulent transition flow around yacht appendages. The
method consists of an empirical approach for the transition
onset momentum thickness Reynolds number and
correlations of integral values and local variables to be
solvable within a RANSE code.

The method has first been verified by comparing


results of a 2D-profile investigation with respective
investigations using a boundary layer method (XFOIL) as
well as experimental data. The drag coefficients from the
different data sources have been compared and the
agreement between them was found to be generally
sufficient.
Fig. 19: Contour plot of kinetic energy on bulb bottom
The second test case, a length variation of a
The most obvious evidence for this theory is the typical ORC42 ballast bulb, successfully extends the
smallest bulb (Variant V1), of which the bottom and top calculation to a fully 3D environment. The length
side show highest curvature. This introduces a laminar variation shows that the assumption of fully turbulent flow
region of almost equal length around the entire body. does not only lead to a significant overestimation of the
flow forces, but – of even more importance –also may lead
A cut through the blade profile shown in the Fig. to wrong determination of optimal geometric properties of
the ballast bulb.
20 demonstrates that the γ-ReΘ-Transition model is capable
of catching flow phenomena otherwise only seen in costly
experimental investigation. One can see that due to an It could be shown that the optimum low-drag
increase of pressure, which results in a loss of kinetic bulb length is significantly longer assuming fully turbulent
energy (due to friction), the laminar boundary layer is no flow than the optimum bulb length determined taking
longer able to maintain attached flow at the wall and laminar regions on the bulb surface into account.
consequently separates. This triggers the boundary layer to
transit to turbulence, thus gaining additional momentum It has to be pointed out that this paper investigates
which diffuses back into the now turbulent boundary layer optimum bulb lengths with respect to resistance
allowing it to reattach to the wall. minimization only. To maximize boat speed the impact of
bulb length on hydrostatic stability has to be taken into
account.
REFERENCES
Abott, I.H., v. Doenhoff, A.E. (1959): Theory of Wing
Sections, Dover Publications Inc., New York, 1959

Drela, M. (2001): XFOIL 6.9 User Primer,


http://web.mit.edu/drela/Public/web/xfoil/,
MIT Cambridge, 2001

Ferziger, J.H., Peric, M. (2002): Computational Methods


for Fluid Dynamics, Springer, New York 2002

Mayle, R.E. (1991): The Role of Laminar-Turbulent


Transition in Gas Turbine Engines, ASME Journal of
Turbo machinery, Vol. 113, 1991

Menter, F.R., Langtry, R.B., Likki, S.R., Suzen, Y.B.,


Huang, P.G., Völker, S. (2004): A Correlation Based
Transition Model Using Local Variables Part I – Model
Formulation, Proceedings of ASME Turbo Expo 2004,
Vienna/AT, June 2004

Menter, F.R., Langtry, R.B., Likki, S.R., Suzen, Y.B.,


Huang, P.G., Völker, S. (2004): A Correlation Based
Transition Model Using Local Variables Part II – Test
Cases and Industrial Applications, Proceedings of ASME
Turbo Expo 2004, Vienna/AT, June 2004

White, F. M. (1991): Viscous Fluid Flow, Mc-Hill Book


Co, New York, 1991

Potrebbero piacerti anche