Anda di halaman 1dari 13

1 INTRODUCTION

Nowadays panel methods represent a standard for propeller flow analysis and, especially at de-
sign advance coefficients, the results obtained from this approach can be considered enough ac-
curate to be employed for propeller design and optimization. Unfortunately off-design condi-
tions are a more challenging task for a potential flow approach, that needs special treatments or
viscous corrections, such as the vortex wake alignment (Gaggero and Brizzolara, 2007), proper
definition of the geometry at tip (Lee C.S, 2004) and ultimately a strong coupling with a bound-
ary layer solver in order to allow for separated flow. Hence, with the increase of computational
power, a RANSE solver, though still requiring higher computational times (in the range of one
order of magnitude), could be a valid approach to solve for these uncertainties and to have a
complete description of the steady flow around the propeller. But only in the recent years these
methods have been applied for complex geometries, thus it is quite difficult to find extensive
validation studies for these kind of engineering problems. After an initial series of studies per-
formed by different authors on DTMB 4119 propeller for the 22
nd
ITTC conference (Gindroz
B., 1999), other presented works known to the authors are those of Rhee and Joshi (2005) on a
more modern highly skewed propeller which shows interesting local flow characteristics as pre-
dicted by the CFD code and demonstrates a certain inaccuracies in the prediction of the torque.
Other interesting articles regards the studies made by several institutes, such as that of Qin
(2006) and Becchi et al. (2005) in the framework of the leading edge project, unfortunately on
unpublished propeller geometries. Interesting recent application is that of Sileo et al. (2007)
which shows CFD results obtained for a small ROV propeller in the complete four quadrant
range of operation.
A systematic comparison between RANS and Panel Methods for
Propeller Analysis
S. Brizzolara, D. Villa & S. Gaggero
Marine CFD Group, University of Genova Department of Naval Architecture and Marine Engineering,
Genova, Italy
ABSTRACT: An extensive validation study about the quality of the results of two different
CFD methods currently used for marine propeller design and analysis is going on at the Marine
CFD Group of the University of Genova. The intention being that to find the optimal character-
istics of mesh generation and problem definition for each code, as well as their relative limits of
application in solving the isolated propeller problem. The paper presents a series of results ob-
tained in the case of five standard test propellers, all simulated into a uniform onset flow (axis-
symmetric problem) over a wide range of advance coefficient. The hydrodynamic results are
compared not only in terms of global characteristics, such as pressure and frictional component
of thrust and torque, but also in terms of local variables, such as pressure distributions on the
blade sections and limiting streamlines, in many cases compared also with experimental data.
The main theoretical and numerical details of the panel method and RANSE solver used are also
resumed.
More recently, novel CFD studies on the old DTMB 4119 propeller (Kulczyk et al., 2007)
highlight the high risk of results variability obtainable with RANSE codes on propellers and the
effects of a good mesh quality and of the turbulence model on the quality of the results. The re-
cent frontier of the CFD research for propellers seems to be devoted to the application of the
methods for the prediction of cavitation, even in difficult (Bulten, 2006) also in unsteady work-
ing conditions (Kawamura et al., 2006).
The present paper presents a systematic application of CFD codes (a in house developed
panel method and a commercial RANSE solver) on a broad range of propeller geometries, used
in the past for validation of lifting surface codes at MIT (Kerwin, 1978 and Greeley, 1982) for
which a wide database of experimental results is available: the 4119 and the 4381-4 highly
skewed propeller family. The intent is to give a picture of what is today achievable with two
state of the art methods for propeller analysis, and the relative difference between the two meth-
ods, both applied with the best experience and with definite intellectual independence by the
same research group. Emphasis is given to the comparison of pressure distribution, velocities
induced in the wake and characteristics strictly related to viscous effects, such as frictional
components of thrust and torque, friction coefficients on the blade and streamline paths on the
blade surface.
In the next two paragraphs a short overview of the two methods is given together with the geo-
metric and physical particulars of the CFD models created to obtain the presented results. Then
prior to the presentation of the numerical results, a paragraph is dedicated to the description of
the geometries of the simulated propellers to enhance the validation of codes on the same cases
among the scientific community.
2 PANEL METHOD AND NUMERICAL MODEL
With the assumption of an inviscid, irrotational and incompressible fluid, the flow around an ar-
bitrary body can be written in terms of a scalar function, the perturbation potential , in a way
such that the total velocity on each point of the flow field can be written, in a propeller fixed
reference coordinate system, as the sum of the undisturbed inflow plus the gradient of the per-
turbation potential:
inf
( , , ) ( , , ) ( , , )
t
x y z x y z x y z = + + q U r

(1)
The perturbation potential also satisfies a partial differential equation: mixing the continu-
ity equation with equation (1) a Laplaces equation arises and by applying adequate boundary
conditions and Greens second theorem it is possible to write the differential problem as a
Fredholm equation of second kind for the potential at each point p laying on the body surface.

( )
1 1 1
2 ( ) ( ) ( )
( ) ( ) ( ) B B W
q
p q q
S S S
q pq q pq q pq
t
t t dS dS t dS
r t r t r t




= +





n n n

(2)
Equation (2) expresses the potential on the propeller blade as a superposition of the potential
induced by a continuous distribution of sources on the blade and hub surfaces and a continuous
distribution of dipoles on the blade, hub and wake surfaces that can be calculated, directly, via
kinematic boundary condition, or, indirectly, inverting equation (2). Via the non-permeability
conditions on the body surface the source strength is known and equal to:
inf
( , , , )
p
q q q q
q
x y z t

=

U n
n

(3)
while the dipoles strength, unknown, is equal to the perturbation potential on the solid
boundaries (blade and hub) and to the potential jump in the wake via the Kutta condition (fol-
lowing Morino approach or a pressure iterative approach as in Gaggero & Brizzolara (2007) and
Lee (1988)). This surface must satisfy a kinematic boundary condition and a dynamic boundary
condition: the kinematic boundary condition requires that the wake surface is a stream surface
of the flow, while the dynamic boundary condition requires that the wake surface is a force free
surface, so the pressure jump between the upper and lower wake surface must be zero. After
discretization of body surfaces and singularity intensities as constant distribution over each
quadrilateral patch, equation (2) can be written as a linear system of algebraic equations, whose
solution is the dipoles intensity over each panel, equal to the perturbation potential. By numeri-
cal differentiation it is possible, by applying equation (1), to calculate the total velocity on each
point of the blade and to compute the local pressure coefficient using Bernoullis theorem:
( )
inf
2
2
2
1
t
P
q
C =

+

U r

(4)
In order to calculate forces on the propeller the pressure has to be integrated over the blades
surface and a viscous correction needs to be adopted to include frictional effect not taken into
account with the potential approach. In particular the viscous correction employed for each i
section along the radius is based on a frictional coefficient dependents by the local sectional ra-
tio between maximum thickness and chord and a chord based Reynolds number:
( )
1 ( )
D Nchord i
t
C k a i R
c


= +



(5)
where k, a and are adequate coefficients. After the convergence study, the simulations with
panel methods were made using a mesh with about 20 panels strips along the radius and about
50 panels along blade profiles, for a total number of about 1200 for one propeller with relative
hub portion.
3 RANSE SOLVER AND NUMERICAL MODEL
Aside the panel method, developed in house, a state of the art RANS solver has been applied to
the same propellers to capture viscous phenomena and to highlight main deficiencies of the po-
tential flow approach. The software suite (CD-Adapco, 2008) has the capability to solve the tur-
bulent viscous flow around a moving body in stationary or non stationary conditions.
The solver is applied to the following group of equations which express the mass and mo-
mentum balance with a Eulerian approach and Reynolds time-Average approach with the par-
ticular boundary conditions valid for the devised CFD model. The RANS equations can be ex-
pressed, in our case, for an incompressible flow:

+ + + =
=
M
P S T U U
U
Re
0



(6)
where U is the average velocity vector field, P is the average pressure field, is the dynamic
viscosity, T
Re
is the tensor of Reynolds stresses and S
M
is the vector of momentum sources.
The component of T
Re
are computed in agreement with the k- turbulence model selected for
this application:
( )
( )
( )
( )

= +

= +

=
k
C D D
k
C grad div div
t
D D k grad div k div
t
k
k D D k
x
U
x
U
ij ij t
t
ij ij
k
t
ij ij t ij ij
i
j
j
i
t ij
2
2 1
Re
2
2
3
2
2
3
2

U
U

(7)
in which

is the turbulent viscosity, k is the turbulent kinetic energy and is the dissipation
term of turbulent kinetic energy.
The RANS solver is based on a Finite-Volume method to discretise the physical domain. The
equation for an incompressible single-phase fluid has been used in the simulation, in combina-
tion with the implicit solver to find the stationary field of all hydrodynamic unknown quantities.
The software uses a SIMPLE method to conjugate pressure field and velocity field, and a AMG
(Algebraic Multi-Grid) solver to accelerate the convergence of the solution. The calculation was
distributed on 6 processors, obtaining a quasi-linear optimal speed-up: this has permitted to
reach reasonable computational time of about 2 hours for each advance coefficient. The stan-
dard k- turbulence model was selected to close the hydrodynamic problem together with a two
layer wall function.
The peculiar unstructured polyhedral mesh (with trimmed type of cells) was created by using
the internal mesher, paying special attention to keep the y+ value on the blade well below a
maximum value of 100 (y+=20 on the major part of the surface), in order to resolve with a suf-
ficient accuracy the flow in the turbulent boundary layer region. The moving reference system
for the solution of the RANSE is used to impose the blade speed of rotation when immersed into
the onset flow.
In order to solve the viscous flow problem in the particular case of the propeller it is neces-
sary to define a set of proper boundary conditions. A constant velocity-inlet condition was set at
a distance of about one diameter upstream of the propeller disc: distant enough from the blade
so that the suction generated by the propeller effect does not affect the constant pressure profile
at the inlet section.
To optimize the computational time, exploitation is made of the aximsymmetric nature of the
flow: so only a single blade is modelled together with its support domain (a prismatic sector)
and a periodic boundary at the interface between two adjacent sectors is used. Hence the geo-
metric domain considered for the problem becomes like that of figure 1, for example.
Moreover around the blade, far enough from it (4 / 5 diameters) a pressure-outlet condition is
imposed, assuming the atmospheric pressure on this boundary; in the cells wherein the com-
puted velocity is predicted inflow, a pressure equal to the stagnation theory is imposed. A longi-
tudinal section of the generated mesh is given in figure 1 (right): it consists in an un-structured
mesh most part of which is carthesian, i.e. cubic cells with progressive refinement coming
closer to the body (in order to improve the quality of the solution avoiding a large number of
cells) The rest of the mesh elements, in a closer proximity of the body, is made of polyhedral
elements (trimmed cubes) in the region of transition between the Cartesian mesh. Near the blade
surface, where the boundary layer need to be accurately solved, a three layers of prismatic cells
extruded from the surface has been create (i.e. the blade and hub surface meshes have been ex-
truded in a direction orthogonal to the surface). Figure 1 presents a longitudinal section of a
typical mesh used for the calculations in this study. The number and type of refinement of the
mesh have been standardized after convergence studies on several propellers. In the average a
total number of about 8105 cells has been used for the simulations presented.


Figure 1. Domain and trimmed mesh used for the propeller CFD model.
4 PROPELLERS ADOPTED FOR VALIDATION
The validation study presented in the following section is focused on five propellers, developed
at David Taylor Model Basin: the standard DTMB 4119, a three blades propeller still widely
used for numerical code validation (Gindroz, 1999; Kulczyk J., 2007 ), and the DTMB 438x
family (DTMB 4381, 4382, 4383 and 4384), obtained systematically from DTMB 4381, a five
blades propeller, keeping the same chord and thickness distributions while progressively in-
creasing the skew distribution and correcting the pitch and maximum camber with the a lifting
surface method, to obtain always the optimum circulation along the radius. All the propellers, as
reported in table 1, are known via their standard design tables and are all designed with NACA
66 modified profile and a=0.8 camber line.

Table 1. Propellers Geometry
r/R c/D rk/D sk[] P/D tmax/c fmax/c
Prop 4119
0.20 0.3200 0.00 0.00 1.1050 0.2055 0.0143
0.30 0.3635 0.00 0.00 1.1022 0.1553 0.0232
0.40 0.4048 0. 00 0.00 1.0983 0.1180 0.0230
0.50 0.4392 0.00 0.00 1.0932 0.0902 0.0218
0.60 0.4610 0.00 0.00 1.0879 0.0696 0.0207
0.70 0.4622 0. 00 0.00 1.0839 0.0542 0.0200
0.80 0.4347 0.00 0.00 1.0811 0.0421 0.0197
0.90 0.3613 0.00 0.0000 1.0785 0.0332 0.0182
0.95 0.2775 0. 00 0.00 1.0770 0.0323 0.0163
0.98 0.2045 0.00 0.00 1.0761 0.0321 0.0145
1.00 0.0800 0.00 0.00 1.0750 0.0316 0.0118


r/R c/D rk/D tmax/c sk [] P/D fmax/c
Prop. 4381 4382 4383 4384 4381 4382 4383 4384 4381 4382 4383 4384
0.20 0.1740 0.0000 0.2494 0.0000 0.0000 0.0000 0.0000 1.3320 1.4550 1.5660 1.6750 0.0351 0.0430 0.0402 0.0545
0.25 0.2020 0.0000 0.1960 0.0000 2.3280 4.6470 6.9610 1.3380 1.4440 1.5390 1.6290 0.0369 0.0395 0.0408 0.0506
0.30 0.2290 0.0000 0.1563 0.0000 4.6550 9.2930 13.9210 1.3450 1.4330 1.5120 1.5840 0.0368 0.0370 0.0407 0.0479
0.40 0.2750 0.0000 0.1069 0.0000 9.3630 18.8160 28.4260 1.3580 1.4120 1.4590 1.4960 0.0348 0.0344 0.0385 0.0453
0.50 0.3120 0.0000 0.0769 0.0000 13.9480 27.9910 42.1520 1.3360 1.3610 1.3860 1.4060 0.0307 0.0305 0.0342 0.0401
0.60 0.3370 0.0000 0.0567 0.0000 18.3780 36.7700 55.1990 1.2800 1.2850 1.2960 1.3050 0.0245 0.0247 0.0281 0.0334
0.70 0.3470 0.0000 0.0421 0.0000 22.7470 45.4530 68.0980 1.2100 1.2000 1.1980 1.1990 0.0191 0.0199 0.0230 0.0278
0.80 0.3340 0.0000 0.0314 0.0000 27.1450 54.2450 81.2830 1.1370 1.1120 1.0960 1.0860 0.0148 0.0161 0.0189 0.0232
0.90 0.2800 0.0000 0.0239 0.0000 31.5750 63.1020 94.6240 1.0660 1.0270 0.9960 0.9730 0.0123 0.0134 0.0159 0.0193
0.95 0.2100 0.0000 0.0229 0.0000 33.7880 67.5310 101.3000 1.0310 0.9850 0.9450 0.9160 0.0128 0.0140 0.0168 0.0201
1.00 0.0010 0.0000 0.0160 0.0000 36.0000 72.0000 108.0000 0.9950 0.9420 0.8950 0.8590 0.0123 0.0134 0.0159 0.0193


5 COMPARISON OF RESULTS
Each propeller has been simulated with the two numerical methods in a wide range of advance
coefficient around the design operational point (J = 0.833 for DTMB 4119 propeller, J = 0.889
for all the four other).
To cross-check the quality of the solution of the two methods, the pressure distribution pre-
dicted on the blade with the two codes has been compared. Figure 2 to 6 present the diagrams of
the pressure coefficients computed at the design advance coefficient for the five propellers (re-
spectively 4119, and 4381 to 4384) on three different blade sections along radius (r/R=0.3,
r/R=0.7 and r/R=0.9). In case of the 4119 also the experimental values are reported, as deduced
by Jessup (1989), with an indirect method based on the measurement of the velocity in the prox-
imity of the blade.
In general a good agreement between the experimental values and computed ones (either with
the Panel code and the RANS solver) can be noted, with the exception of a limited region close
to the trailing edge, where the ability of the RANS solver to consider the effect of the boundary
layer displacement thickness on the pressure distribution is necessary to come closer to the real
flow. The blade profiles are predicted to work at almost the same effective hydrodynamic angle
by the two methods (about the first two third of the pressure diagram, especially the part closer
to the leading edge, are practically coincident) with an appreciable difference only in a region
close to the hub (r/R=0.3), where the CFD, where the effect of the boundary layer is more pro-
nounced (due to the profiles higher thickness)and the different prediction of cross-flow can also
play a role.
The series of graphs of figures 2 to 6 is completed with the comparison of the computed
thrust and torque curves predicted by the two codes in a wide range of advance ratio against ex-
perimental values measured at cavitation tunnel of David Taylor Model Basin and that of M.I.T.
(where available). In general, again, computed values fit very well to the experimental ones,
also for what regards the torque which not always confirmed by other researchers using other
methods. While both the panel method and the RANS solver are able to predict exactly the
thrust and torque in proximity of the design advance coefficients, the RANS solver shows better
accuracies al lower advance coefficients values, i.e. for highly loaded conditions, quite off-
design. In this operating region, the viscous effects and possibly flow separation at the leading
edge start to play a relative importance also on the global values and the current version of the
panel method (already improved with respect to standard panel method as outlined in the previ-
ous sections, but still not coupled with a thin boundary layer solver) is inherently incapable to
take into account these phenomena.


Figure 2. Pressure distribution and Thrust and Torque Coefficients Propeller DTMB 4119


Figure 3. Pressure distributions (r/R=0.3, 0.7, 0.9 respectively) and Thrust and Torque Coefficients Pro-
peller DTMB 4381


Figure 4. C
P
on blade profiles (at r/R=0.3, 0.7, 0.9 respectively), K
T
, K
Q
: Propeller DTMB 4382


Figure 5. C
P
on blade profiles (at r/R=0.3, 0.7, 0.9 respectively), K
T
, K
Q
: Propeller DTMB 4383


Figure 6. C
P
on blade profiles (at r/R=0.3, 0.7, 0.9 respectively), K
T
, K
Q
: Propeller DTMB 4382

For the 4119 propeller, experimental measurements of velocities in the wake flow field of the
propeller were performed by Jessup (1989) and they are compared with those predicted by the
two codes. The comparison of figure 7 relates to a circular arc in a transverse section about one
third of diameter aft of the propeller, at a radius r/R=0.7 and with an angular development equal
to 2/z (z is the blade number) in order to span from the vortical wake surface shed by one blade
to that of the consecutive one. Radial, tangential and axial components of the velocity field are
compared. In general a sufficiently good agreement is found on the trends between the experi-
mental value and the numerical obtained by both codes (with a better conformity of the RANS
model), except for the region close to the core of the helical vortex line (at the examined radial
position), where the experiments show a much sharper reduction in the axial induced component
and an increase of the tangential component, consistent with the vortex theory. It is believed that
the RANSE model was not sufficiently refined in the region of the wake to capture this local-
ized viscous phenomena, hence dissipating the vortical energy too much; while it is strange that
the panel method, based on potential flow so on a conservative formulation, shows results more
similar to the (viscous) RANS model than to the experiments, in this case.

Table 2. Comparison of the prediction of thrust and torque, due to pressure and tangential components,
with Panel Method and RANS solver.
DTMB 4119 J = 0.833 Visc. Corr. No Visc. Pan Press+Shear Press. Only RANS
K
T
0.1463 0.1508 0.0045 0.1384 0.1436 0.0052

10K
Q
0.2831 0.2518 0.0313 0.2819 0.2488 0.0331
DTMB 4383 J = 0.889 Visc. Corr. No Visc. Pan Press+Shear Press. Only RANS
K
T
0.2218 0.2289 0.0071 0.2082 0.2149 0.0067

10K
Q
0.4670 0.4260 0.0410 0.4444 0.3995 0.0449
DTMB 4383 J = 0.400 Visc. Corr. No Visc. Pan Press+Shear Press. Only RANS
K
T
0.4577 0.4629 0.0052 0.4608 0.4650 0.0042

10K
Q
0.6900 0.6531 0.0369 0.7686 0.7351 0.0335


40 60 80 100 120 140
0.2
0
0.2
0.4
0.6
0.8
1
1.2
1.4

V
X
/
V
,

V
T
V
,

V
R
/
V

V
X
/V exp.
V
T
/V exp.
V
R
/V exp.
V
X
/V Rans
V
T
/V Rans
V
R
/V Rans
V
X
/V Pan.
V
R
/V Pan
V
T
/V Pan

Figure 7. Comparison between axial, radial and tangential induced velocities computed and measured
downstream, x/D = 0.328, r/R = 0.7

The predictions of induced velocities on the blade surface is better rendered by plotting the
streamlines predicted, as in the example of figures 8 and 9 for the face an back side of DTMB-
4119 and figures 10 and 11 for the DTMB 4383, both simulated at their respective design ad-
vance ratios. Main difference between the two numerical methods are found in the region close
to the hub mainly at the trailing edge of the blade, where the RANS method shows large radial
outward velocity components, that the panel method shows in much minor entity only on the
back of 4119, while on the face it shows the opposite trend. Similar picture in the case of the
4383 DTMB propeller, with a more extended blade area in which the RANS code predict these
strong outward cross flow velocities. These effects could be due to boundary layer abrupt thick-
ening at the trailing edge of the blade which force the attached flow to curve and avoid this low
energy region. This phenomenon, capture by the RANS method, is clarified and confirmed by
the behaviour of the friction coefficient on the blade, as represented in figures 12 and 13, which
shows in the same region an almost null tangential force on the blade.
In terms of global results, though, the difference between the two methods is almost disap-
pear. Table 2, in fact, shows the thrust and torque coefficients computed by the two methods,
with and without the friction forces: for the panel method this correspond to the viscous correc-
tion of equation (5), and for RANS the calculation of the forces have been done with the only
pressure components or including the shear stresses on the blade and hub surface. It is clear that,
while there is a small discrepancy between the two methods, in particular regarding thrust, the
empirical formula used in the panel method to allow for viscous effects is substantially able to
capture the global result of the above mentioned phenomena. In fact, the relative differences
(viscous computation minus potential computation and pressure computation minus pressure
plus shear stress computation) are almost equal between the two methods. This finding is true
also in off design conditions the viscous correction applied to the panel method seems to predict
well the relative difference between pressure and shear stress calculations, while a substantial
difference remains (especially in the torque computation) with respect to RANS calculation.
This difference, then, has to be imputed to the viscous-pressure effects, evidently not adequately
modelled by the simple formula (5) and not to shear stress components.


Figure 8. Streamlines on DTMB 4119 (Back). Panel method left, RANS right


Figure 9: Streamlines on DTMB 4119 (Face). Panel method left, RANS right




Figure 10: Streamlines on DTMB 4383 (Back). Panel method left, RANS right



Figure 11: Streamlines on DTMB 4383 (Face). Panel method left, RANS right




Figure 12: Local friction coefficient C
f
on DTMB 4119, at J=0.833. Face (left), Back (right)


Figure 13: Local friction coefficient C
f
on DTMB 4383 at J=0.889. Face (left), Back (right)
6 CONCLUSION AND FURTHER PERSPECTIVES
A systematic application of state of a state of the art panel method (developed by the same ma-
rine CFD group at DINAV) and a RANSE solver has been performed on a series of propellers
for which the experimental measurements obtained in cavitation tunnel are available.
From the analysis of the local and global parameters the following conclusions can be drawn:
both methods are able to predict with a sufficient accuracy for design purposes (within 2-3%)
the thrust and torque curves in a wide range of advance coefficients. In fact, pressure distribu-
tions predicted by both methods along the blade at design advance coefficient are very similar
(except for a limited region at the trailing edge) and in also congruent with the experimental
measurements.
The RANSE solver, as expected, demonstrates a better accuracy at low advance ratios, i.e. is
better suited for representing highly loaded propeller, or in off-design conditions, where impor-
tant viscous effects boundary layer thickness (and eventually separation) at the trailing edge and
at blade root and the generation of a strong secondary cross flows along the blade, start to condi-
tion heavily the potential flow solution. These effects are currently not adequately represented
by the actual too simplified empirical viscous corrections that are applied to the potential flow
method and confirm the possibility of improving the panel methods through a strong coupling
with a boundary element method, allowing for limited flow separated regions. This can improve
also the prediction of radial components of the velocities especially on the back of the blade sur-
face, close to the trailing edge, at root. In this respect, systematic applications of RANS code as
those presented in this paper can be useful for studying deriving alternative more general em-
pirical corrections, for instance based on the deviation of the operating advance ratio from the
design value. An semi-empirical treatment for modeling these dead flow regions on the blade
could be sufficient as a first approximation and is currently under investigation.
For what regards the predictions of induced velocities in the wake, both methods agree rather
well with experiments, but both are not able to adequately capture the increase of tangential ve-
locities and decrease of axial velocities close to the position of the helical vortex line shed from
the blade in the wake. However, the main hydrodynamic effects that may be necessary to study
the hydrodynamic action of a propeller on body immersed in its wake (like a rudder or another
propeller in the case of contra-rotating system) are sufficiently well captured by both methods
for engineering purposes.
Future activities will be directed to the coupling of a thin boundary layer theory to the panel
method and to the application of such investigations as those presented in this study to the case
of sheet cavitating propeller. In fact, both the in house panel method and the RANS solver are
able to consider (with different theoretical models) these phenomena.

7 REFERENCES
Becchi P., Pittaluga C. (2005), Comparison between RANSE calculation and Panel Method Results for
the Hydrodynamic Analysis of Marine Propeller , RINA Marine CFD, Royal Institution of Naval Ar-
chitects
Bulten N., Oprea I. (2006), Evaluation of McCormicks rule for propeller tip vortex cavitation inception
based on CFD results, Proceedings of CAV 2006 International Symposium on Propeller Cavitation,
Wageningen (NL)
Gaggero S., Brizzolara S. (2007), Exact Modelling of Trailing Vorticity in Panel Method for Marine
Propeller, Proceedings of ICMRT 2007, Ischia (IT)
Gindroz B. (1999), Final report and recommendations to the 22
nd
ITTC Propulsion Committee, Venice
(IT).
Greeley D. S., Kerwin J.E. (1982), Numerical Methods for Propeller Design and Analysis in Steady
Flow SNAME Transaction, Vol. 90, pp. 415-453
Jessup S. (1989), An Experimental Investigation of Viscous Effects of Propeller Blade Flow, PhD The-
sis, The Catholic University of America
Kawamura T., Takekoshi Y., Minowa T., Maeda M., Yamaguchi H., Fujii A., Kimura K., Taketani T.
(2006), Simulation of Unsteady Cavitating Flow around Marine Propeller using a RANS CFD code,
Proceedings of CAV 2006 International Symposium on Propeller Cavitation, Wageningen (NL)
Kerwin J.E., Lee C.S. (1978), Prediction of Steady and Unsteady Marine Propeller Performances by
Numerical Lifting Surface Theory, SNAME Transaction, Vol. 88, pp. 218-253
Kulczyk J., Skraburski L., Zawislak M. (2007), Analysis of screw propeller 4119 using the Fluent sys-
tem, Archives of Civil and Mechanical Engineering, Vol. VII 2007 No. 4
Lee C.S., Kim G.D., Kerwin J.E. (2004), A B-Spline Based High Order Panel Method for the Analysis
of Steady Flow around Marine Propeller, 25
th
Symposium on Naval Hydrodynamics, St. Johns,
Newfoundland.
Lee J.T. (1988), A Potential Based Panel Method for the Analysis of Marine Propeller in Steady Flow,
PhD. Thesis, M.I.T. Department of Ocean Engineering
Rhee S.H., Joshi S. (2005), Computational Validation for Flow around Marine Propeller using Unstruc-
ture Mesh Based Navier-Stokes Solver, JSME International Journal Series B, Vol. 48 (2005), pp.562-
570.
Sanchez-Caja A. (1998), DTRC Propeller 4119 Calculations at VTT, Proceedings of the 22nd ITTC
Propulsion Committee Propeller RANS/Panel Method Workshop, Grenoble, France.
Sileo L., Bonfiglioli A., Magi V., (2007), RANSEs Simulation of the Flow past a Marine Propeller under
Design and Off-design Conditions, Proceedings of ICMRT 2007, Ischia (IT)
CD-Adapco (2008), StarCCM+, v. 2.10 Users manual.

Anda mungkin juga menyukai