Sunteți pe pagina 1din 16

Fourth International Symposium on Marine Propulsors

smp15, Austin, Texas, USA, June 2015

Cavitating Flow Calculations for the E779A Propeller in Open Water and Behind
Conditions: Code Comparison and Solution Validation

Guilherme Vaz1 , David Hally2 , Tobias Huuva3 , Norbert Bulten4 , Pol Muller5 , Paolo Becchi6 , Jose L. R. Herrer7 ,
Stewart Whitworth8 , Romain Mace9 , Andrei Korsstrom10
1 2 3
MARIN, The Netherlands. Defence Research and Development Canada. Caterpillar, Sweden.
4 5 6 7
Wartsila, The Netherlands. DCNS Research, France. CETENA, Italy. Navantia, Spain.
8 9 10
Lloyds Register, UK. DGA Techniques hydrodynamiques, France. ABB, Finland.

ABSTRACT However, even though wetted flow CFD calculations for propul-
sors are becoming straightforward exercises, the same is not true
As part of the Cooperative Research Ships SHARCS project, cal- for cavitating flow simulations which are inherently unsteady and
culations of the E779A propeller in open water and in a cavitation more expensive, and are done using cavitation models which are
tunnel behind wake generating plates have been performed by ten semi-empirical and not yet mature enough. Also, for accurate
different institutions using eight different flow codes. Both full predictions of pressure fluctuations on ship hulls caused by the
RANS and RANS-BEM coupled approaches have been used to dynamic behaviour of the cavity, the interaction between tur-
predict wetted and cavitating flows. Propeller performance char- bulence modelling, cavitation modelling and numerical details
acteristics, pressure distributions, limiting-streamlines and cav- needs to be completely controlled. This is not an easy task even
itation volumes have been analyzed. Cavitation patterns have when considering only the first harmonic of the propeller blade-
been compared with photographs. In addition, pressure fluctu- passage frequency.
ations at the cavitation tunnel walls and at some hydrophones
have been computed and compared against available experimen- In SMP-2009 [1], results obtained for the well-known E779A
tal data. For loading and cavitation extents, there is good agree- benchmark test-case [2] with five different CFD codes were
ment among the different calculations. Compared with measure- published. Both open water flow and the flow behind a pre-
ments, the predicted cavity extents are good but the propeller scribed non-uniform velocity field (trying to model the effect
thrust for the behind condition was uniformly under-predicted. of a wake stimulator) were considered. The results were com-
The overall agreement is an improvement over earlier studies us- pared against BEM results and experimental data. Bensow and
ing the same data set. The pressure fluctuations predicted by Bark [3] present implicit LES results for this case. In both these
half the participants were in reasonable agreement with measure- works, only performance characteristics and qualitative compar-
ments, but the remaining calculations predicted pressure levels a isons of cavity extents were made for a simplified geometrical
factor of four or five too high. setup and for few flow conditions.
In SMP-2011 a workshop was organized for two common test-
Keywords cases: the twisted Delft foil and the Potsdam Propeller Test Case
CFD, BEM, URANS, RANS+RANS, RANS+BEM, Propellers, (PPTC). For the propeller case [4], the emphasis was only on
E779A, Open water, Behind condition, Cavitation, Pressure the comparison of open water cavitating flow performance pre-
fluctuations, ANSYS
R
CFX R
, ANSYS Fluent R
, Excalibur, dictions; the PPTC propeller had little cavitation mostly at the
TM
R
FINE /Marine, OpenFOAM , PROCAL, ReFRESCO, root. The ITTC-2011 committee [5] gave a detailed assessment
Star CCM+ R of the status of numerical tools for simulation of cavitation and
associated pressure fluctuations. They found that: Promising
1 Introduction advances in propulsion simulation by both in-house and com-
mercial CFD software are made during the last three years. It
At present, viscous flow Reynolds-Averaged Navier-Stokes seems necessary to model the true geometry of propeller and
(RANS) codes for marine propellers, thrusters, complex propul- make the computational mesh sufficiently fine for both hull and
sors, and even complete ships with propulsor arrangements, are propeller, in order that the propulsion factors are predicted more
available within the maritime industry. During the design of accurately. Meanwhile, the body-force approach remains to be
propulsors, potential flow tools such as Boundary Element Meth- an alternative that is efficient and easy-to-use for engineering
ods (BEM) are still the work-horse codes; for analysis, viscous purposes. There is, however, a lack of benchmark data for the
flow tools using RANS, URANS or even hybrid DES/SAS ap- validation of propulsion prediction. Further R&D work for nu-
proaches are taking over. This is due to wide availability of com- merical propulsion simulations are proposed as follow: 1) Study
mercial and open-source CFD solvers, as well as cheaper and of numerical uncertainties arising from mesh resolution, turbu-
more powerful hardware which, allied with parallelization ac- lence modelling and numerical discretization schemes; 2) Full
celeration techniques, make calculations feasible now that were scale propulsion prediction; 3) Prediction of cavitation and fluc-
impossible in the past. tuating pressure for propeller operating behind the hull.


c Cooperative Research Ships and Her Majesty the Queen in Right of Canada, as represented by the Minister of National Defence, 2015.
In this context, within the Cooperative Research Ships (CRS) 2 Test-Case
consortium (www.crships.org), the SHARCS (Ship Hydrody-
The test-case consists of the flow around the INSEAN E779A
namic Advanced RANS Cavitation Simulations) working-group
propeller inside a cavitation tunnel [2]: Fig. 1. A compre-
is currently investigating the performance of state-of-the-art CFD
hensive series of experimental data addressing the propeller in
tools for predicting pressure fluctuations on ships due to cavitat-
uniform, as well as non-homogeneous, flow was gathered at
ing propellers. Three cases with increasing complexity are being
INSEAN over the last decade [2, 14, 15]. The E779A pro-
considered: an open water propeller, a propeller in the behind
peller has also previously been subject of calculations by the
condition and a fully appended and propelled ship. In this paper
EU VIRTUE project [1,16], the EU STREAMLINE project [17]
results from the first two cases are presented. The E779A pro-
and the CRS PROCAL project [18]. In the current work, the
peller was again used as a test case because it is one of the only
cavitation patterns, cavitation dynamics and associated pressure
high-quality publicly available test cases which considers both
fluctuations are the major focus. Nevertheless, wetted flow cal-
open water and behind conditions with a large amount of avail-
culations are also considered and whenever relevant compared
able experimental data. Due to the advances in the capabilities of
against experimental data. In order to minimize the differences
the codes and hardware available, fewer concessions in terms of
between the calculations, the geometry, physical and numerical
computational grids, time-steps and geometrical simplifications
conditions were the same for all participants. The following sec-
have been made relative to the SMP-2009 study. To date, only
tion describes the set-up in detail.
model-scale conditions have been considered in CRS SHARCS
work.
Ten different institutions have actively participated in this
work: ABB-Finland, Caterpillar-Sweden, CETENA-Italy,
DCNS Research-France, DGA Techniques hydrodynamiques-
France (DGAH), Defence Research and Development Canada
(DRDC), Lloyds Register-UK (LR), MARIN-Netherlands,
Navantia-Spain and Wartsila-Netherlands. Both potential flow
(BEM) and viscous flow (RANS) approaches have been consid-
ered, including the coupling of both in a so-called RANS-BEM
coupled approach. For the potential flow calculations only one
tool (or tool combination) has been used: PROCAL [6] and Ex-
calibur [7]. These are tools developed within CRS and avail-
able to all participants. For the viscous flow approach, open-
source, open-usage and commercial tools have been used (in al-
phabetical order): 1) ANSYS CFX [8]; 2) ANSYS Fluent [9];
3) FINE/Marine [10]; 4) OpenFOAM [11]; 5) ReFRESCO [12]; Figure 1: E779A propeller inside the cavitation tunnel.
6) STAR-CCM+ [13]. Due to the number of different partici-
pants and codes participating, this study is representative of the 2.1 Geometries
current ability of RANS and BEM tools to tackle cavitating flow INSEAN E779A is a four-bladed, fixed pitch, right-handed
simulations on a propeller and associated pressure fluctuations. model propeller, originally designed in 1959. Its diameter, D,
This paper presents results for the E779A propeller in open wa- is 227.27 mm. Details of its geometry can be found in [2, 14, 15].
ter and in a cavitation tunnel positioned behind wake generating During the VIRTUE project the actual geometry of the propeller
plates. For open water wetted flow, propeller performance char- blade was measured and an IGES file created. The IGES ge-
acteristics have been quantitatively analyzed. For cavitating flow, ometry was cleaned and smoothed and a new solid geometrical
pressure distributions, limiting-streamlines, the cavity patterns, description was prepared and used in the current work.
cavity volume and sensitivity to the volume fraction chosen to 2.2 Open Water Calculations
represent the cavity have been studied. For the behind condition,
attention was paid to the nominal velocity field behind the wake The cavitation tunnel of Fig. 1 has a test section with square
generating plates. Cavitating flow predictions for the propeller cross-section of width 0.6 m and length 2.6 m. To simplify the
operating behind the stationary plates were then made. The same computational modelling of the propeller in open water condi-
quantities as for the open water case have been analyzed. In ad- tions, an idealized tunnel was used having a circular cross section
dition, pressure fluctuations at the tunnel walls and at some hy- equal in area to the actual tunnel, as was done in [1, 16].
drophones have been computed and compared against available Fig. 2 illustrates the prescribed computational domain; the do-
experimental data. main diameter was 2.942 D. A common definition of boundary
The paper is organized as follows. The E779A propeller test- conditions was used: prescribed velocity and turbulence intensity
cases are described in Section 2. Section 3 gives details of the at the inlet, uniform pressure at the outlet (although not strictly
numerical approaches, numerical settings and grids used. The compatible with swirling flow), and slip at the tunnel walls. No-
results for open water conditions are presented in Section 4 fol- slip conditions were required on shaft, fairing, hub, propeller and
lowed by the results for the behind condition in Section 5. Fi- cap surfaces. While the propeller, hub and cap rotated with rate
nally, Section 6 presents conclusions made from the study. n, the shaft and fairing were stationary. We emphasize that this
test case corresponds to the cavitation-tunnel set-up and not the of the original geometry were not described in any of the avail-
towing-tank set-up, also available in [2]. In the experiments, no able references [2,14,15]: some decisions were made so the flow
roughness was used at the leading edges to stimulate transition domain and tunnel were the same for all participants. The tunnel,
to turbulence. The following physical parameters were used: wake generator and shaft/hub were generated so that the center
T = 20 C, = 998 kg/m3 , = 1.008 103 N s/m2 , of the propeller plane is at (0, 0, 0). The cross section of the
= 1.01 106 m2 /s. tunnel has a rectangular section with sides of 0.6 m and corner
pvap = 2.337 103 P a, vap = 0.017kg/m3 , radius of 0.1 m. The length of the tunnel was chosen to be 2.2 m.
vap = 1.02 105 N s/m3 . The tunnel was longitudinally positioned so that the distance be-
Uniform velocity field V = Vin at the inlet. tween the propeller plane and the outlet was 4D, as was done
Inlet and background turbulence intensity of 2%. for the open water case. The wake generator was modelled as
Propeller rotation rate n = 36 rps, with the different advance described in [15]. The length of the spacer where it attaches to
coefficients, J = Vin /nD, being obtained by changing the the tunnel roof was unknown and therefore was assumed to be
inflow velocity Vin . 0.76 m in the longitudinal direction. The dimensions of the shaft
were also unknown and therefore simplified and estimated. The
For wetted flow, an open water diagram was calculated. The vertical plate upstream of the wake generator which attaches the
advance coefficients J = 0.71 and J = 0.83 were obligatory shaft to the tunnel roof, clearly visible in the photo of Fig. 1, was
conditions since those were also used for the cavitating flow omitted. The larger radius of the shaft upstream of the wake gen-
cases. For cavitating flow three conditions were consid- erator was estimated as 0.08 m and it was extended all the way
ered: 1) (J = 0.71, n = 0.63); 2) (J = 0.71, n = 1.763); 3) to the inlet.
(J = 0.83, n = 1.029), where n = (p pref )/ 12 n2 D2 . In
order to keep the conditions trim-independent relative to the A common definition of boundary conditions was also required
moving blade, gravity was not included. The reference pres- for this case: they were the same as for the open water calcula-
sure used in the definition of the cavitation number was the out- tions except that the tunnel walls and wake stimulator were no-
let pressure. The Reynolds number based on chord length at slip walls. Fig. 3 illustrates the chosen domain and boundary
r/R = 0.7 is 5 105 when J = 0.71. Therefore the flow cannot conditions.
be considered fully turbulent; rather, it is critical or even sub-
critical for the lower radii sections where large parts of laminar,
transitional and turbulent flow can coexist.

Figure 3: Domain and boundary conditions for the calculations


in the behind condition.

The physical parameters were the same as for the open water
test case except that the rotation rate was n = 30.5 rps corre-
sponding to (open water) J = 0.897. The propeller operated in
a non-uniform wake caused by the boundary layers of the plates
and other appendages. The interaction between the stationary
and rotating parts of the computational domain was modelled ac-
cording to the CFD codes and/or users best practices.
For this case, only one loading condition, J = 0.897, was
computed. For wetted flow conditions, both a nominal-wake
and a total-wake case were considered: i.e. a case without the
propeller and with the operating propeller, respectively. The
nominal velocity field in a plane 0.26D upstream of the pro-
peller disk was compared among all participants and compared
with existing time-averaged experimental data [15]. With the
propeller included, the emphasis lies on the propeller perfor-
mance, cavitation dynamics and pressure fluctuations. Two cav-
Figure 2: Domain and boundary conditions for the open water itation numbers were considered, = 2.5 and = 5.5, where
calculations. = (p pref )/ 21 Vin
2
. For the cavitation dynamics, the cal-
culations and experiments were compared at different propeller
2.3 Calculations in the Behind Condition blade positions. Four pressure sensors, P1 to P4, were placed
For the calculations in the behind condition, the cavitation tun- on the tunnel walls directly to right, left, above and below the
nel and some of its components were modelled. Several details centre of the propeller disk. Four hydrophones, H1 to H4, were
placed in a vertical line on the centreplane one radius aft of the incident pressures (one has therefore a RANS-BEM-BEM cou-
propeller disk and 80 mm, 100 mm, 120 mm and 200 mm be- pled approach). Van Wijngaarden [7] provides details of Excal-
low its centre, respectively. Fig. 4 shows their locations in the ibur and hydro-acoustic coupling.
tunnel and relative to the propeller. Pressure fluctuations were
calculated, analyzed and compared against experiments for these 3.2 CFD Approach
eight locations. 3.2.1 Open water calculations: The propeller rotation
was modelled in several ways: 1) using a body-fixed reference
system, on which calculations can be steady; 2) using an earth-
fixed reference system where the complete domain is rotating
and calculations are unsteady; 3) using an earth-fixed reference
system where only a sub-domain involving the blades is rotating,
and sliding-grids or interfaces are needed; 4) using a body or
earth-fixed reference system and considering only a 2/Z angu-
lar sector of the computational domain shown in Fig. 2 together
Figure 4: Locations of pressure taps and hydrophones. with cyclic/periodic boundary conditions. While for the wetted
flow case steady RANS calculations could be performed, for the
3 Numerical Formulation cavitating flow case all calculations were unsteady except those
of CETENA and two of the three performed by Navantia.
3.1 BEM Approach
All cavitating flow calculations were restarted from the wetted
As mentioned above, both BEM potential flow and CFD vis- flow converged solution. To obtain the correct cavitation number
cous flow tools, and their combination, have been used in the , some participants decreased from a high value to the lower
current work, the emphasis of this paper being however on the desired value in steps in order to prevent large sudden variations
CFD tools. of the cavitation pattern and possibly divergence of the calcula-
tion. In addition, some participants modified the reference pres-
BEM open water calculations, whether for wetted or for cavitat-
sure pref to obtain the desired cavitation numbers while others
ing flow, are usually performed in an infinite domain (no tunnel
modify the vapour pressure, pvap .
walls) assuming that the flow is steady. The propeller then op-
erates on an undisturbed uniform velocity field V = Vin . These Four different cavitation models were used: the Kunz model [25],
calculations are straightforward and computationally very effi- the Zwart model [26], the Singhal model [27] and the Sauer
cient, with computational times of the order of minutes, max- model [28]. There are several differences in the numerical and
imum hours, on a typical workstation. Vaz and Bosschers [19] physical behaviour of these models but, in general, and based on
provide details of the set-up of a BEM calculation using the same previous experience of the authors, one can state that the Kunz
cases that are studied here. For the behind condition, a potential- model has the numerical advantage (and physical disadvantage)
flow-only approach implies that the propeller is operating in the of having the condensation source/sink term independent of the
effective wake (see Carlton [20] for more details), which takes pressure p and the Sauer model of being the most realistic, and
into account the velocity deficit due to the plates wake and has therefore the most widely used. Huuva [29] provides a detailed
to be determined a priori using some other method. The calcu- description and a comparison of these models as applied to pro-
lations are then unsteady and, even though still computationally pellers.
efficient, the computation times are usually of the order of many
hours, maximum a day, also on a workstation. All partners used eddy-viscosity turbulence models, variants of
the k  or k models, commonly used for non-cavitating
In the current work, for open water conditions the same approach flow simulations. Caterpillar however used the Reboud cavitat-
is used. However, for the behind condition the effective wake ing flow correction [30], a damping of the eddy-viscosity in the
is not calculated a priori but during the computational process, mixture region. All partners except MARIN used wall-functions
using RANS-BEM coupling [18, 21, 22, 23]. The procedure used for near-wall turbulence model boundary conditions and there-
here performs a (cheap) steady RANS calculation for the domain fore had high y + values.
containing the tunnel, plates and shaft, together with an (also
cheap) unsteady BEM calculation (wetted and cavitating flow) 3.2.2 Calculations in the Behind Condition: All calcu-
for the propeller blades. The coupling is done iteratively, where lations in the behind condition were unsteady and used sliding-
in the RANS domain the propeller is modelled via body-forces, grids/interfaces in an earth-fixed reference frame. In some cases,
and in the BEM domain the interaction is done via the effective- the wetted flow calculations were initially done using a quasi-
wake. The detailed procedure employed here by DRDC is ex- steady frozen-rotor approach, followed by the fully-unsteady
plained by Hally [24]. In order to compute the pressure fluctua- wetted flow sliding-interfaces approach. MARIN and LR solved
tions caused by the cavitating rotating propeller, use of the acous- the turbulence-model equations up to the wall instead of using a
tic potential flow BEM tool Excalibur is also made. In this case, wall-function approximation. For the rest of the set-up, most nu-
the hydrodynamic propeller noise/pressure sources computed by merical choices made for the open water calculations were main-
the hydrodynamic BEM are passed to the acoustic BEM code tained here. For the calculation of the pressure fluctuations, no
which solves the scattering effect of the solid boundaries on the compressibility effects were considered in the CFD approach and
the pressure values coming directly from the URANS calcula-
tions were monitored in time at the required points without any
additional post-processing.
3.3 Codes, Numerical Settings and Grids
Tabs. 1 and 2 summarize the most relevant numerical choices and
settings used for all cavitating flow calculations performed. For
more details on the tools, the reader may consult the respective
web sites cited. One may observe that the tools, grid generators, (a) ABB+PROCAL. (b) Caterpillar+OpenFOAM.
approaches, numerical discretization choices and resolution, are
very heterogenous and representative of the current state-of-the-
art for application of CFD to hydrodynamic problems.
Fig. 5 illustrates the grids used for the open water calculations.
Fig. 6 shows a slice by the y = 0 plane through each grid used for
the behind condition. For the partners that performed both cases,
the blade surface grid of the open water calculation presented in
Fig. 5 was kept similar for the calculations in the behind condi-
(c) CETENA+ANSYS CFX. (d) DCNS+FINE/Marine.
tion (though the MARIN grid was finer close the leading-edge).
However, in the rotating sub-grid, most partners refined the grids
at the propeller slip-stream but not all the way to the tunnel walls
or to the locations of the hydrophones.
4 Results: Open Water Calculations
4.1 Wetted Flow
All results now presented are based on the previously de-
scribed numerical settings and best-practice guidelines of each (e) MARIN+ReFRESCO. (f) Navantia+STAR-CCM+.
partner and numerical codes. The iterative, spatial and time-
discretization convergence of each calculation is therefore con-
sidered adequate. Some partners have performed extra studies to
study these effects but those results are not shown here.
Fig. 7 compares the results of six partners for the thrust, torque
and efficiency in an open water diagram. The experimental data
[2, 14, 15] were measured in a towing-tank at a slightly different
Reynolds number and are therefore used here simply for qualita-
(g) Wartsila+STAR-CCM+.
tive comparison. The difference between all numerical results
is less than 5% for KT and KQ , and 10% for . Compared Figure 5: Grids for the open water calculations.
with earlier studies (e.g. the EU VIRTUE project [1]), the spread
between the results is lower in the current work. When com- different levels of grid resolution in those zones, different turbu-
pared with the experimental data, the averaged differences are lence models and boundary conditions and intrinsic inviscid flow
even lower than 5%. The differences in KT and KQ are larger assumptions of the BEM code for the trailing edge.
at higher propeller loadings (lower J). The potential flow results 4.2 Cavitating Flow
for the open water diagram have similar accuracies as the viscous
flow ones, although there is a small but consistent over-prediction In this section we present first several results for the nominal
of KQ . condition with J = 0.71 and n = 1.763, then the comparison
between all numerical results and the experimental data for all
Fig. 8 shows the distribution of the pressure coefficient Cpn on three conditions.
the propeller blades and Fig. 9 shows Cpn versus x/c for two
different radial sections: r/R = 0.7 and 0.9. The pressure An interesting issue that has been discussed in the past [1] is how
distributions differ mainly for contours between Cpn = 2.0 and to define the cavity surface. As an iso-surface of the vapour vol-
3.0. There are also point-to-point oscillations in some cases due ume fraction? For which value? And are the cavity extents sensi-
to interpolation errors by the visualization package at hanging tive to this value? Or based on the iso-surface where Cp = ?
nodes: they are not in the original solution. This is corroborated Figs. 10 and 11 show the pressure distribution on the blade and
by Fig. 9 which does not show any oscillations. Also, Fig. 9 at some radial sections. One can see a clear Cpn = n contour
shows that all codes (including the potential flow code) agree in Fig. 10 and line in Fig. 11 but with some differences between
reasonably well, with some differences at the leading-edge, suc- all partners, especially when close to the cavity detachment and
tion peak and trailing-edge. Close to the propeller tip these dif- re-attachment. It is also known that there must be a pressure
ferences are larger. These small discrepancies are caused by the lower than the vapour pressure in order for cavitation to start and
Settings/Partner ABB Caterpillar CETENA DCNS MARIN Navantia Wartsila
Code PROCAL v2.2 OpenFOAM v2.1 ANSYS CFX v15 FINE/Marine v3.1.1 ReFRESCO v1.3.1 Star CCM+ v8.04 Star CCM+ v7.04
Calc.Type Steady Unsteady+MVG+Cyclic-bc Steady+Cyclic-bc Unsteady+SI Steady/Unsteady+AFM Steady/Unsteady+MVG Unsteady+MVG
Grid Default (radial) Hybrid (hexahedrals+prisms) Hybrid (tetrahedrals+prisms) Trimmed Hexahedrals Structured Hexahedrals Polyhedrals Trimmed Hexahedrals
Provise [31] ANSATM [32] ANSYS [33] HEXPRESSTM [34] GridProTM [35] Star CCM+ [13] Star CCM+ [13]
MCells (200 40) panels 1.3 7.3 10.0 9.5 6.6 8.6
Time-step 0.02 2.0 0.5 1.3 0.4
(Max/Avg y + ) (171, 75) (15, 7) (60, 20) (1, 0.1) (125, 62) (331, 150)
Discretization Default 2nd order for mom. 2nd order for mom. 2nd order for mom. 2nd order for mom. 2nd order for mom. 2nd order for mom.
1st order for turb. 1st order for turb. 1st order for turb. 1st order for turb. 2nd order for turb. 2nd order for turb.
1st order cav. 1st order cav. 1st order cav. 1st order cav. 1st order cav. 1st order cav.
2nd order in time 1st order in time 1st order in time 2nd order in time 1st order in time 1st order in time
Turb.Model RNG k- + Reboud cor. Std. k- k- SST 1994 k- SST 1994 Real. k- Std./Real. k-
Cav.Model Default Kunz Model Zwart Model Sauer Model Sauer Model Sauer Model Sauer Model
Cc = 10000 Fc = 0.03 n0 = 108 m3 n0 = 108 m3 n0 = 1012 m3 n0 = 1012 m3
Ce = 400 Fe = 300 R = 105 m R = 106 m R = 106 m
Iterative Conv. tol = 0.1 L2 < 106 all residuals
Other Detach.=LE Only 1 blade Only 1 blade

Table 1: Settings for the open water cavitating flow calculations.

Settings/Partner Caterpillar DGA DRDC LR MARIN Wartsila


Code OpenFOAM v2.1 ANSYS Fluent v13 ANSYS CFX v15+PROCAL v2.217 Star CCM+ v8.06 ReFRESCO v2.0.0 Star CCM+ v7.04
+Excalibur v2.6.1.3
Calc.Type Unsteady+SI Unsteady+SI Steady(RANS)+Unsteady(PROCAL) Unsteady+SI Unsteady+SI Unsteady+SI
+Freq.(Excalibur)
Grid Polyhedral Structured Hexahedrals Structured Hexahedrals + BEM Unstructured Polyhedral Structured + Unstructured Hexahedrals Trimmed Hexahedrals
ANSA [32] ICEM-CFD [36] Pointwise [37] + Provise [31] Star CCM+ [13] GridPro [35] + HEXPRESS [34] Star CCM+ [13]
MCells 9.2 19.3 9.2+(156800 + 253)panels 22.1 8.1 14.5
Time-step 0.2 0.25 2.5 0.125 0.25 0.1
(Max/Avg y + ) Propeller: (85, 21) Propeller: (33, 11) Propeller: Propeller:(31, 0.6) Propeller:(1.4, 0.5) All: (265, 100)
Tunnel: (245, 83) Tunnel: (311, 31) Tunnel: (300, 100) Tunnel: (85, 1.7) Tunnel:(370, 60)
Discretization 2nd order for mom. 2nd order for mom. BEM: Default 2nd order for mom. 2nd order for mom. 2nd order for mom.
1st order for turb. 1st order for turb. RANS: 2nd order for mom. 2nd order for turb. 1st order for turb. 2nd order for turb.
2nd order cav. 1st order cav. 1st order for turb. 2nd order cav. 1st order cav. 1st order cav.
1st order in time 1st order in time 2nd order in time 2nd order in time 1st order in time
Turb.Model RNG k- + Reboud cor. RNG k- k- SST 1994 DES k- SST k- SST 2003 Real. k-
Cav.Model Kunz Model Singhal Model PROCAL: Default Sauer Model Sauer Model Sauer Model
Cc = 104 Non-cond.mass frac. = 107 n0 = 1012 m3 n0 = 108 m3 n0 = 1012 m3
Ce = 500 R = 106 m R = 105 m R = 106 m
Iterative Conv. 30 iterations per time-step 5 iterations per time-step L2 < 5 105 all residuals 20 iterations per time-step
or 60 iterations per time-step
Other Initialized with FR PROCAL:Detach.=LE, Prescribed wake Initialized with FR

Table 2: Settings for the in-behind cavitating flow calculations.


(a) Caterpillar+OpenFOAM. (b) DGAH+ANSYS Fluent.
(a) ABB. (b) Caterpillar.

(c) DRDC+ANSYS CFX (d) LR+STAR-CCM+.


(c) DCNS. (d) MARIN. (e) Wartsila.

Figure 8: Open water wetted flow Cpn distribution. J = 0.71.


6
4
r/R = 0.7
2

Cpn
0
(e) MARIN+ReFRESCO. (f) Wartsila+STAR-CCM+.
2
4
Figure 6: Grids for the calculations in the behind condition.
6
0 0.2 0.4 0.6 0.8 1
0.9 Fractional Chord Length

10KQ 16
0.8
12
0.7 r/R = 0.9
8
Cpn

0.6 KT 4
0.5 0

0.4 4
0 0.2 0.4 0.6 0.8 1
0.3 Fractional Chord Length
ABB Caterpillar
0.2 CETENA DCNS
MARIN Navantia Wrtsil
0.1
0 Figure 9: Open water wetted flow Cpn vs x/c on radial sections
0 0.2 0.4 0.6 0.8 1 1.2 r/R = 0.7 and 0.9. J = 0.71.
J
50% causes only small qualitative differences in the cavity ex-
Experiment tents. Notice that for a potential flow approach, as in PROCAL,
ABB Caterpillar CETENA this is not an issue since well-defined dynamic and kinematic
MARIN Navantia Wrtsil boundary conditions are used to define the cavity surface (see [6]
for more details). Based on these results we chose a vapour vol-
Figure 7: Open water wetted flow KT , KQ and .
ume iso-surface of 0.10 (10% of vapour in a grid cell) to define
the cavity surface.
that this under-pressure is very much dependent on the cavitation
model. Nevertheless, the low pressures within the cavity seen in Figs. 1315 show the cavity extents for all participants, for the
the Caterpillar and DCNS distributions in Fig. 11 are considered three flow conditions, together with the available experimental
outliers. data. Several interesting trends are worth emphasizing:
Fig. 12 shows the influence of the value of the vapour volume For the nominal condition with J = 0.71 and n = 1.763,
iso-surface on the cavity extents for one calculation; however, it the comparison between all numerical results, and against the
is representative of all the calculations and independent of the experimental data, can be considered good. Nevertheless,
cavitation model used. One can see that the lower the value the some calculations under-predict the extents while others over-
larger the cavity. However, varying the value between 10% and predict them. Also, the viscous flow approaches detect cavita-
(a) ABB. (b) Caterpillar.

(a) v = 0.1. (b) v = 0.5. (c) v = 0.9.

Figure 12: Influence of v iso-surface on cavitation extent.


J = 0.71, n = 1.763.
tion does not show any cavitation and the viscous flow ap-
proaches (all of them!) consistently show a super-cavitating
(c) CETENA. (d) DCNS. sheet. It should be noted, however, that the potential flow cal-
culation predicted pressure levels below the vapour pressure at
the lower radii, but because only sheet cavitation is modelled
and because the detachment point was constrained to be at the
leading edge, the algorithm did not let cavitation occur.
When J = 0.83 and n = 1.029, a very small low pressure re-
gion is visible for all calculations. Depending on the cavitation
model and its settings, some of the RANS calculations predict
(e) MARIN. (f) Navantia. (g) Wartsila. a small tongue-shaped mid-chord sheet cavity. The potential
flow cavitation model does not permit mid-chord cavitation.
Figure 10: Open water Cpn distribution for cavitating flow.
J = 0.71, n = 1.763. In general, the MARIN, Navantia and Wartsila results look
very similar to each other, even though the codes, grids and
4
numerical settings are very different. The common numerical
r/R = 0.7 feature between these three approaches is the Sauer cavitation
2
model.
Cpn

0
The pressure distributions predicted by both potential and vis-
2
cous flow approaches are considered to be accurate and similar
4 for the seven different calculations. All of them predict regions
0 0.2 0.4 0.6 0.8 1 of pressure lower than vapour pressure that are larger than the
Fractional Chord Length
cavity extents present in the experiments. But it is also known
8 from the literature (for example see Kuiper [38] or Franc [39])
6 r/R = 0.9 that cavitation does not occur in water with low nuclei content,
4 even with high under-pressures, and also not when the flow is
Cpn

2 laminar. For that reason, leading edge roughness is commonly


0 used at some model-basin facilities to promote cavitation. The
2
differences between the numerical results and the experiments
are thought to be explained by this lack of nuclei, roughness
0 0.2 0.4 0.6 0.8 1
Fractional Chord Length and associated turbulent flow.
ABB Caterpillar
CETENA DCNS In general, the results of all viscous flow calculations are very
MARIN Navantia Wrtsil much alike, even when using completely different codes, nu-
merical settings and turbulence/cavitation models. This is re-
Figure 11: Open water cavitating flow Cpn vs x/c for radial assuring and shows the maturity of CFD for cavitating flows.
sections r/R = 0.7, 0.9. J = 0.71, n = 1.763.
5 Results: Behind Condition
tion at the blade lower radii which is not present in the experi-
ments nor in the potential flow code predictions. 5.1 Nominal Wake Flow
Fig. 16 presents the computed axial velocity distribution normal-
When J = 0.71 and n = 0.630, the difference between
ized by the inlet velocity, Vx /Vin , at a plane 0.26D upstream
the potential flow and viscous flow approaches and the ex-
of the propeller disk. In Fig. 16 all results have been obtained
periments are very large. While in the experiments one can
using a steady RANS approach. The differences between them
very clearly see bubble cavitation, the potential flow calcula-
(a) ABB+PROCAL. (b) Caterpillar+OpenFOAM. (c) CETENA+ANSYS CFX. (d) DCNS+FINE/Marine.

(e) MARIN+ReFRESCO. (f) Navantia+STAR-CCM+. (g) Wartsila+STAR-CCM+. (h) Experiments.


Figure 13: Open water cavity extents. J = 0.71, n = 1.763.

(a) ABB+PROCAL. (b) Caterpillar+OpenFOAM. (c) CETENA+ANSYS CFX. (d) DCNS+FINE/Marine.

(e) MARIN+ReFRESCO. (f) Navantia+STAR-CCM+. (g) Wartsila+STAR-CCM+. (h) Experiments.


Figure 14: Open water cavity extents. J = 0.71, n = 0.630.

(a) ABB+PROCAL. (b) Caterpillar+OpenFOAM. (c) CETENA+ANSYS CFX. (d) DCNS+FINE/Marine.

(e) MARIN+ReFRESCO. (f) Navantia+STAR-CCM+. (g) Wartsila+STAR-CCM+. (h) Experiments.


Figure 15: Open water cavity extents. J = 0.83, n = 1.029.
are clear, with better similarity between MARIN and LR, and
DGAH and Wartsila. Nevertheless, the wake is very sharp with
large gradients in the circumferential direction, much larger than
would normally be experienced behind a ship. In some cases,
a strong asymmetry is also visible as well as large low-velocity
areas which correspond to the wakes of the rods connecting the
planes. (a) Caterpillar. (b) DGAH. (c) DRDC.
The large variation in the steady wakes is due to unsteady von
Karman vortex shedding from the rods; they cannot be captured
correctly with a steady approach. Time-averaged wakes from un-
steady calculations performed by some partners compared much
better with the experimental data (see Fig. 17) but the experiment
still shows a region of velocity deficit below the shaft that is not
found in the calculations.
(d) LR. (e) MARIN. (f) Wartsila.
5.2 Wetted Flow
Tab. 3 lists the predicted averaged loads, Kt and Kq , for the pro- Figure 16: Nominal axial velocity at x/D = 0.26 for the be-
peller rotating behind the wake generating plates. The measured hind condition. J = 0.897.
loads are Kt = 0.175 and 10Kq = 0.334 [15] so the values are
very low for all partners. The spread in computed averaged loads
is around 11% for Kt , about twice as large as in the open water
case; the reason for the relatively poor agreement is unclear. For
Kq , the spread is 5%, about the same as for the open water case.

Partner-Code Avg. Kt Avg. 10Kq


Caterpillar+OpenFOAM 0.154 0.303
DGAH+ANSYS Fluent 0.147 0.298
DRDC+ANSYS CFX 0.145 0.291
(a) MARIN+ReFRESCO. (b) Experimental data.
+PROCAL
LR+STAR-CCM+ 0.140 0.289 Figure 17: Axial velocity at x = 0.26D for the behind condi-
MARIN+ReFRESCO 0.151 0.300 tion. (left) Time-averaged unsteady calculations; (right) Time-
averaged experimental LDV data. J = 0.897.
Wartsila+STAR-CCM+ 0.139 0.302
0.18 0.34
0.17
Table 3: Mean wetted flow loads. 0.32
0.16
Fig. 18 shows the variation of thrust and torque for one blade pas- 0.15 0.30
sage; one can see that the minimum thrust/torque is achieved at 0.14
0.28 10KQ
a blade angle near = 30 and the maximum at 80 , for all vis- 0.13
KT
cous flow calculations. The potential flow results present some 0.12 0.26
oscillations deviating from the normal sinusoidal-like behaviour 0 30 60 90 0 30 60 90
Angle (degrees) Angle (degrees)
of thrust and torque temporal distribution. Caterpillar DGAH DRDC
LR MARIN Wrtsil
Fig. 19 shows Cpn on radial sections r/R = 0.7 and 0.9 for
one blade angle, = 0 . For this angle the pressure distribu- Figure 18: Wetted flow propeller loads for the behind condition.
tions at two high-radii sections are very similar for all calcula- J = 0.897.
tions. The three major visible differences are: 1) the oscillations
on the potential flow results, which are due to the wake velocity partners and computed when available). Notice that this volume
content and the lack of diffusion in a potential flow code; 2) dif- comprises the complete domain and not only one blade or only
ferent values of the minimum pressure, a quantity very sensitive the propeller; if cavitation appears on the plates or inside the hub
to the leading-edge grid quality and resolution; 3) a hump in the vortex this volume considers those locations too.
pressure distribution for section r/R = 0.90 which is due to a
leading-edge vortex, only captured by the calculations not using Fig. 20 shows that the cavity volume variation is qualitatively
wall-functions. the same for all viscous flow approaches, the results of the po-
tential flow approach being different in terms of amplitude and
5.3 Cavitating Flow phase. The results of the viscous flow approach present sharper
Fig. 20 shows the variation with time of the cavity volume. The temporal variations, and the maximum volume does not occur at
error bars in this figure represent the standard deviation com- 0 but closer to 15 . Two viscous flow results show high cavity
puted from the several steady-state revolutions (different for all volume outside the range of the wake peak. It is known from
10 0.17 0.18
8 Wetted Wetted
0.16 = 5.5 0.17 = 5.5
6 = 2.5 = 2.5
r/R = 0.7 0.15 0.16
4
Cpn

2 0.14 0.15
0 0.13 0.14
2 0.12 0.13
4
0.11 0.12
6 0 30 60 90 0 30 60 90
0 0.2 0.4 0.6 0.8 1
Angle (degrees) Angle (degrees)
Fractional Chord Length
25 (a) LR+STAR-CCM+. (b) MARIN+ReFRESCO.
20 Figure 22: Kt for wetted and cavitating flow in the behind con-
15 dition. J = 0.897.
Cpn

10 r/R = 0.9
The effect of cavitation on the loads can be seen in Fig. 22 which
5
compares the loads calculated by MARIN and LR for wetted
0
flow and for = 2.5 and = 5.5; the results of DGAH and
0 0.2 0.4 0.6 0.8 1 Wartsila were qualitatively similar. For the = 2.5 condition
Fractional Chord Length the effect is considerable between 10 and 40 , with an increase
Caterpillar DGAH DRDC
LR MARIN Wrtsil
in loading occurring during the collapse of the cavity. A small
decrease of performance is seen while the cavity is growing, i.e.
Figure 19: Cpn vs x/c on radial sections r/R = 0.7 and 0.9 for between 70 and 10 (or 20 and 10 in Fig. 20).
wetted flow in the behind condition. = 0 , J = 0.897. Figs. 23 and 24 show the calculated pressure distributions for
= 2.5 when the blade is at = 0 . The results from all calcu-
6 lations are similar, with the differences between viscous and po-
DGAH
Cavity Volume (cm )
3

5 DRDC tential flow approaches being visible especially close to the cav-
LR
4 MARIN ity re-attachment area. Differences at the beginning of the cavity
3 Wrtsil are also visible with some results showing pressures lower than
2 vapour pressure. While for a potential flow approach p = pvap
1 is an imposed dynamic boundary condition, for a viscous flow
0 cavitation modelling approach an under-pressure is needed to
40 20 0 20 40
Angle (degrees) feed the evaporation process. However, the amount of the under-
pressure and the size of the area where it occurs depend on the
Figure 20: Cavity volume for the behind condition. J = 0.897, cavitation model, its constants and the grid resolution.
= 2.5.

the classical theory of cavitation and propellers [39, 20], that the
second time-derivative of the cavity volume is the main contri-
bution to the cavitating flow pressure fluctuations and therefore
the results presented in Fig. 20 are important for understanding
and explaining the results obtained for the pressure fluctuations
presented below.
(a) Caterpillar. (b) DGAH. (c) DRDC.
Fig. 21 shows the variation of KT and KQ over one blade pas-
sage when = 2.5. The differences between the numerical re-
sults have increased relative to the open water flow. Non-physical
high-frequency oscillations appear in the Wartsila calculations.
0.17 0.34
0.16 0.32
0.15 0.30
0.14 0.28 (d) LR. (e) MARIN. (f) Wartsila.
0.13 0.26 10KQ Figure 23: Cpn distribution for cavitating flow in the behind con-
KT
0.12 0.24 dition. = 0 , J = 0.897, = 2.5.
0.11 0.22
0 30 60 90 0 30 60 90 Figs. 2527 and Figs. 2830 show the results obtained for the
Angle (degrees) Angle (degrees) cavity extents for = 2.5 and = 5.5, respectively. Only
Caterpillar DGAH DRDC
LR MARIN Wrtsil angles = 20 , 0 and +20 were chosen for illustrative pur-
poses. For the viscous flow results, the cavity is represented by
Figure 21: Cavitating flow propeller loads for the behind condi- an iso-surface of vapour volume equal to 10%. The following
tion. J = 0.897, = 2.5. observations can be made:
6
shaft rate frequency provided in the experimental data set [15].
4
r/R = 0.7 Similarly, the DRDC time history at P1 was obtained from four
2
harmonics of blade rate frequency calculated using Excalibur;
Cpn

0 DRDC is not included in the results at H1 as Excalibur cannot


2 provide realistic predictions of the pressure in the wake. Since
4 neither the experimental nor the DRDC data sets include an ab-
6 solute pressure reference, the mean pressure was subtracted from
0 0.2 0.4 0.6 0.8 1
Fractional Chord Length all data sets to aid in comparison. It should be noted that, for the
4 experiments, the rotation angle of the blades at time t = 0 is not
3 known. Therefore, the experimental curve may be shifted right
2 or left by an arbitrary amount.
1
Cpn

0 The pressure fluctuations predicted by Caterpillar, LR and


1
Wartsila are four to five times higher than the experimental val-
2
3 r/R = 0.9 ues and, as can be seen by comparing the upper portions of Figs.
4 33 and 34, the time histories of the pressure are insensitive to the
0 0.2 0.4 0.6 0.8 1
sensor location, the results for all pressure taps and hydrophones
Fractional Chord Length
Caterpillar DGAH DRDC being almost the same. The LR and Wartsila predictions are
LR MARIN Wrtsil very similar; each used STAR-CCM+ with the Sauer cavitation
model. The Caterpillar prediction is similar except that it lacks
the pronounced peaks near 3, 11, 20 and 28 msecs; they used
Figure 24: Cpn vs x/c on radial sections r/R = 0.7 and 0.9 for
OpenFOAM with the Kunz cavitation model.
cavitating flow in the behind condition. = 0 , J = 0.897, =
2.5. DGAH, DRDC and MARIN predict more realistic pressure fluc-
tuations when compared with the reconstructed experimental sig-
In general, the agreement between the viscous flow results and nal. For the P sensors, the amplitude tends to be under-predicted,
the experiments is good, for both cavitation numbers, and is perhaps due to the lower loading of the propeller relative to the
much better than was found for open water conditions. The ex- experiment. At some of the P sensors, DGAH and MARIN pre-
perimental results now show a cavity that extends to low radial dict the complete signal quite well, as is the case for MARIN
sections in a smooth distribution likely due to the non-uniform in Fig. 34. At the H sensors, there was generally worse agree-
wake that disturbs the flow enough to promote cavitation in- ment between the numerical and experimental values which is
ception and growth (see Kuiper [38]). understandable given the rapid coarsening of the grids in the ax-
For = 20 , the potential flow results show an isolated ial direction. Nevertheless, the values obtained by MARIN and
patch of cavitation at the lower radii similar to the open wa- DGAH were of the same order of magnitude as the experimental
ter potential flow results at J = 0.71 and n = 0.63 (see values. With respect to potential versus viscous flow method re-
Fig. 14a). This deserves further investigation (Vaz and Boss- sults, the results by DRDC are comparable with the more expen-
chers [19] showed cavitation at lower radial sections than sive URANS approaches; however, Excalibur is not capable of
shown here). This is probably due to the modelling used predicting the pressure inside or very close to the propeller wake,
for detachment-point, but RANS-BEM coupling for cavitating so the pressures at H1, H2 and H3 were not correctly computed.
flow conditions could also play a role. Even though the RANS-BEM coupling approach does not suffer
Apart from one calculation, all the results seem to under- from numerical diffusion, the pressure fluctuations predicted by
predict the cavity extents. This is explained by the under- this method also under-predicted the experimental results.
prediction of the propeller loads previously discussed.
LR predicts a cavitating hub vortex; these were the calcula- All organizations predicted P2 to have the highest levels for the
tions using the finest grid of all partners. LR, MARIN and first harmonic among the P sensors; the experiment shows that
Wartsila, all using the Sauer cavitation model, predict very P3 has the highest level with P2 second. All organizations ex-
similar cavity extents. A small cavitating tip-vortex is seen in cept Caterpillar correctly predicted that P3 would have higher
some of the calculations. At 20 , only Caterpillar predicts the levels than P1 (these are the sensors on the left and right walls of
mid-chord cavitation seen in the experiments. the tunnel). All organizations except DRDC correctly predicted
that P4 would have the lowest amplitude for the first harmonic
5.4 Pressure Fluctuations among the P sensors, though in the DRDC results the P1 and P4
levels only differ by 2 P a. In the experiments, the amplitudes
In this section only the results for the lowest cavitation number of the third and fourth harmonics at P2 exceed the first and sec-
= 2.5 will be presented. Figs. 31 and 32 show histograms of ond harmonics. This behaviour was not predicted by any of the
the first four harmonics of blade passage frequency at the sen- organizations.
sors H1, inside the propeller wake, and P2, on the wall to the
left of the propeller. The results from the other sensors are qual- Two major poorly-understood weaknesses of these calculations
itatively similar. Figs. 33 and 34 show the time histories of the are the high-frequency oscillations seen in the results of some
pressure for the same sensors. For these plots, the time histories partners even for non-cavitating loads, and the very high pressure
of the experimental data were obtained from the 48 harmonics of fluctuations predicted by some partners. To try to shed some
(a) Caterpillar. (b) DGAH. (c) DRDC. (d) LR. (e) MARIN. (f) Wartsila. (g) Experiments.
Figure 25: Cavity extents for the behind condition. = 20 . J = 0.897, = 2.5.

(a) Caterpillar. (b) DGAH. (c) DRDC. (d) LR. (e) MARIN. (f) Wartsila. (g) Experiments.
Figure 26: Cavity extents for the behind condition. = 0 , J = 0.897, = 2.5.

(a) Caterpillar. (b) DGAH. (c) DRDC. (d) LR. (e) MARIN. (f) Wartsila. (g) Experiments.
Figure 27: Cavity extents for the behind condition. = +20 , J = 0.897, = 2.5.

(a) Caterpillar. (b) DGAH. (c) DRDC. (d) LR. (e) MARIN. (f) Wartsila. (g) Experiments.
Figure 28: Cavity extents for the behind condition. = 20 , J = 0.897, = 5.5.

(a) Caterpillar. (b) DGAH. (c) DRDC. (d) LR. (e) MARIN. (f) Wartsila. (g) Experiments.
Figure 29: Cavity extents for the behind condition. = 0 , J = 0.897, = 5.5.

(a) Caterpillar. (b) DGAH. (c) DRDC. (d) LR. (e) MARIN. (f) Wartsila. (g) Experiments.
Figure 30: Cavity extents for the behind condition. = +20 , J = 0.897, = 5.5.
3 Experiment For the two partners using STAR-CCM+, Fig. 35 shows Cpn
Amplitude (kPa)

2.5 Caterpillar
DGAH close to the sliding-interfaces at one blade angle. Clearly some
2 LR
1.5 MARIN
pressure perturbations exist at those areas. The results of DGAH,
1
Wrtsil MARIN and Caterpillar (not presented here) showed no abnor-
0.5 mal pressure jumps between the interfaces, and obviously the re-
0 sults by DRDC do not use interfaces. Therefore, these anomalies
1 2 3 4 could explain some high-frequency peaks but not the very high
Harmonic
pressure fluctuations.
Figure 31: Amplitudes of the first four harmonics of pressure at
sensor H1. J = 0.897, = 2.5.
3.5 Experiment
3
Amplitude (kPa)

Caterpillar
2.5 DGAH
2 DRDC
LR
1.5 MARIN
1 Wrtsil
0.5
0
1 2 3 4
Harmonic

Figure 32: Amplitudes of the first four harmonics of pressure at


sensor P1. J = 0.897, = 2.5.
15 Caterpillar
Pressure (kPa)

10 LR
Wrtsil
5 Experiment
0
5
0 5 10 15 20 25 30
Time (msecs)
2 Figure 35: Pressure perturbations close to/at the sliding inter-
DGAH
Pressure (kPa)

1.5
1 MARIN faces in cavitating flow: (top) LR; (bottom) Wartsila. = 0 ,
Experiment
0.5 J = 0.897, = 2.5.
0
0.5
1 Fig. 36 shows the results obtained for the wetted flow case at the
1.5
0 5 10 15 20 25 30 P1 sensor (LR did not perform these calculations). The predicted
Time (msecs)
pressures are now of the same order of magnitude but high-
Figure 33: Pressure time history at sensor H1. J = 0.897, frequency oscillations are present in the Caterpillar and Wartsila
= 2.5. results. Even for the smoother DGAH and MARIN results there
are some minor numerical (non-physical) oscillations. However,
15 Caterpillar notice that the scale of the ordinate is much smaller with ampli-
Pressure (kPa)

10 LR
Wrtsil tudes of the order of 100 to 200 Pa. The equivalent variation in
5 Experiment Cpn is Cpn < 0.01. It is possible that the iterative conver-
0 gence per time-step is not low enough to permit such a level of
5 accuracy.
0 5 10 15 20 25 30
Time (msecs) 0.3 Caterpillar
Pressure (kPa)

2 0.2 DGAH
DGAH
Pressure (kPa)

0.1 DRDC
DRDC
1 MARIN 0 MARIN
Experiment 0.1 Wrtsil
0 0.2
0.3
1 0 5 10 15 20 25 30
0 5 10 15 20 25 30 Time (msecs)
Time (msecs)
Figure 36: Pressure time history at sensor P1 for wetted flow.
Figure 34: Pressure time history at sensor P1. J = 0.897, J = 0.897.
= 2.5.
We conclude that: 1) the origin of the outlying pressure fluc-
light on these issues, two extra studies were performed: 1) a de- tuations by some of the partners is likely due to the cavitation
tailed analysis of the pressure distribution close to the interfaces modelling (not only the cavitation model itself but its implemen-
between the non-rotating and the rotating grids; 2) a comparison tation and inter-connection with the other equations solved); 2)
between all numerical predictions for the pressure fluctuations the higher-frequency perturbations are probably due to numeri-
for the wetted flow case. cal errors at the interfaces and/or lack of iterative convergence,
with or without the sliding-interfaces. This clearly deserves fur- the first harmonic of the blade-passage frequency. The pres-
ther investigation, both by the users and by the developers of the sures at the sensors at the wall were better predicted than at
viscous flow tools used. the hydrophones in the flow. The RANS-BEM approach was
located in the group with fair agreement with the experiments.
6 Conclusions Although further verification is needed, it is conjectured that
the high levels of the pressure fluctuations for the cavitating
In this paper, calculations for the E779A propeller in open water
flow case are due to the cavitation model implementation, and
and in a cavitation tunnel behind wake generating plates have
that the high-frequency oscillations are due to numerical ac-
been presented. Ten different institutions performed calcula-
curacy at the interfaces and/or enough iterative convergence to
tions and eight different codes were used including full RANS
achieve accurate results for very low pressure levels.
and RANS-BEM coupled approaches. The results are therefore
representative of the state-of-the-art for cavitating simulations Compared with previous comparative studies, the current results
on propellers and associated pressure fluctuations using incom- present both a better agreement between all viscous and poten-
pressible flow (potential and viscous) numerical approaches. tial flow results and with the experimental data. When comparing
the viscous versus the potential flow approaches, one has to bal-
From the open water results one can draw the following major
ance the pros and cons of both approaches (accuracy versus cost).
conclusions:
From the results obtained here, and ones available in the litera-
Very different grids were used, both structured and unstruc- ture, a RANS-BEM approach seems the best option to use when
tured, although unstructured grids were more common. All first tackling a new cavitating flow problem in the behind con-
except two partners used large y + values and wall-function dition; a full RANS calculation can be used once more detailed
boundary conditions. Both viscous flow RANS and potential results are needed. We must emphasize however that, while a
flow BEM approaches were employed. full RANS approach is inherently expensive, a RANS-BEM ap-
Despite the variety of methods used, very similar loads were proach is very sensitive to numerical settings and is not as robust
obtained. Differences between all results are about 5% for nor as accurate as a full RANS approach. Future work in the CRS
KT and KQ . Difference in flow fields are more visible, but SHARCS working group will deal with similar comparative nu-
not large. merical studies for a real appended ship plus cavitating propeller
There were large differences in cavitation extent which seem configuration, associated pressure fluctuations, together with val-
due to the different cavitation models and associated constants idation against available experimental data.
rather than to any other numerical settings. The results ob-
tained by different codes using the same cavitation model are REFERENCES
very similar. Differences in cavity size may be due to the lack
of turbulence, seeding/nuclei issues and the interaction with [1] Salvatore, F., Streckwall, H., and van Terwisga, T., 2009.
cavitation detachment (a known problem with E779A pro- Propeller Cavitation Modelling by CFD - Results from
peller). the VIRTUE 2008 Rome Workshop. In First Interna-
tional Symposium on Marine Propulsors SMP09, Trond-
From the results for the behind condition one can state: heim, Norway, June.
CPU times varied between the calculations as is to be ex- [2] Pereira, F., Salvatore, F., and Di Felice, F., 2004. Mea-
pected given the different methods employed. Full RANS cal- surement and Modelling of Propeller Cavitation in Uniform
culations with sliding-interfaces typically take weeks, while Infow. J. of Fluids Engineering, 126, July, pp. 671679.
RANS-BEM calculations typically take one or two days. [3] Bensow, R.E., and Bark, G., 2010. Implicit LES Predic-
For wetted flow, the nominal wake predicted using steady tions of the Cavitating Flow on a Propeller. J. of Fluids
RANS differed among the partners. However the unsteady Engineering, 132, pp. 110.
time-averaged nominal wake was close to the (time-averaged) [4] Klasson, O., and Huuva, T., 2011. Potsdam Propeller Test
experimental LDV data. Nevertheless, the average thrust value Case (PPTC). Proc. of the SMP2011. Hamburg, Germany,
numerically predicted was far from the experiments for all June.
codes/partners. Given the good agreement with the LDV data [5] ITTC Propulsion Committee , 2011. Final Report and
and the good prediction of open water performance character- Recommendations to the 26th ITTC. Proc. of the 26th
istics, this is difficult to explain and deserves further attention. ITTC.
For cavitating flow, the loads and pressure distributions on the
[6] Vaz, G., 2005. Modelling of Sheet Cavitation of Hydro-
blade and their variation with time look qualitatively similar.
foils and Marin Propellers using Boundary Element Meth-
Once again, predictions by different partners using the same
ods. PhD thesis, IST,Instituto Superior Tecnico, Portugal,
cavitation model are very similar. In general, the cavity ex-
June.
tents agreed well with the experimental data though all codes
under-predicted them slightly and the potential flow code had [7] van Wijngaarden, E., 2011. Prediction of propeller-
difficulty modelling the cavitation at lower radii. induced hull-pressure fluctuations. PhD thesis, Delft Uni-
Predictions of pressure fluctuations fell into two groups: one versity of Technology, Delft, the Netherlands, November.
with high-frequency oscillations and levels over-predicted by [8] www.ansys.com/Products/Simulation+Technology/Fluid+
four to five times; the other with a generally fair agreement for Dynamics/Fluid+Dynamics+Products/ANSYS+CFX.
[9] www.ansys.com/Products/Simulation+Technology/Fluid+ Von Karman Institute for Fluid Dynamics.
Dynamics/Fluid+Dynamics+Products/ANSYS+Fluent. [27] Singhal, A. K., Athavale, M. M., Li, H., and Jiang, Y., 2002.
[10] www.numeca.com/en/products/finetmmarine. Mathematical Basis and Validation of the Full Cavitation
[11] www.openfoam.com. Model. J. of Fluids Engineering, 124, pp. 617624.
[12] www.refresco.org. [28] Sauer, J., 2000. Instationar kavitierende stromungen -
ein neues modell, basierend auf front capturing (vof) und
[13] www.cd-adapco.com/products/star-ccm-plus.
blasenddynamik. PhD thesis, University of Karlsruhe,
[14] Pereira, F., Salvatore, F., Di Felice, F., and Soave, M., 2004. Germany. (in German).
Experimental Investigation of a Cavitating Propeller in
[29] Huuva, T., 2008. Large Eddy Simulation of Cavitating and
Non-Uniform Inflow. In Proc. of the Twenty-Fifth ONR
Non-cavitating flow. PhD thesis, Chalmers University of
Symposium on Naval Hydrodynamics, St. Johns, New-
Technology, February.
foundland, Canada.
[30] Coutier-Delgosha, O. and Astolfi, J. A., 2003. Nu-
[15] Salvatore, F., 2007. The INSEAN E779A Propeller Ex-
merical Prediction of the Cavitating Flow on a Two-
perimental Dataset. Tech. Rep. D4.1.3, INSEAN, VIRTUE
Dimensional Symmetrical Hydrofoil with a Single Fluid
WP4, May.
Model. Cav2003 Proc., Osaka, Japan, Nov.
[16] Rijpkema, D., and Vaz, G., 2011. Viscous Flow Com-
[31] Bosschers, J., 2009. PROCAL v2.0 Users Guide. Tech.
putations on Propulsors: Verification, Validation and Scale
Rep. RP 20834-6-RD, MARIN, June.
Effects. In Proc. of RINA-CFD2011, London, UK, March.
[32] www.beta-cae.gr/ansa.htm.
[17] Vaz, G., Windt, J., Wackers, J., Manzke, M., and Eskils-
son, C., 2012. Application and Verification of Grid Adap- [33] www.ansys.com/Products/Workflow+Technology/
tation. Tech. Rep. STREAMLINE-MARIN-DEL-D323-1, ANSYS+Workbench+Platform/ANSYS+Meshing.
MARIN, September. [34] www.numeca.com/en/products/automeshtm/hexpresstm.
[18] Bosschers, J., Vaz, G., Starke, A., and van Wijngaarden, E., [35] www.gridpro.com.
2008. Computational Analysis of Propeller Sheet Cavita- [36] www.ansys.com/Products/Other+Products/ANSYS+
tion and Propeller-Ship Interaction. In MARINE CFD2008 ICEM+CFD.
Proc. , Southampton, UK, March.
[37] www.pointwise.com.
[19] Vaz, G., and Bosschers, G., 2006. Modelling Three Di-
[38] Kuiper, G., 1981. Cavitation inception on ship propeller
mensional Sheet Cavitation on Marine Propellers using a
models. PhD thesis, Netherlands Ship Model Basin.
Boundary Element Method. In Cav2006 Proc. , Wagenin-
gen, The Netherlands, September. [39] Franc, J.-P., and Michel, J.-M., 2004. Fundamentals of
Cavitation. Kluwer Academic Publishers, Dordrecht, The
[20] Carlton, J. S., 2007. Marine Propellers and Propulsion.
Netherlands.
Butterworth-Heinemann Ltd.
[21] Mishra, B., 2005. Prediction of Performance of Podded
Propulsors via Coupling of a Vortex-Lattice Method with
an Euler or a RANS Solver. PhD thesis, The University of
Texas at Austin, August.
[22] Starke, A. R., and Bosschers, J., 2012. Analysis of
scale effects in ship powering performance using a hybrid
rans-bem approach. 29th Symp. Naval Hydrodynamics,
Gothenburg, Sweden.
[23] Rijpkema, D., Starke, A. R., and Bosschers, J., 2013. Nu-
merical Simulation of Propeller-Hull Interaction and De-
termination of the Effective Wake Field using a Hybrid
RANS-BEM Approach. In Proc. of the 3rd International
Symposium on Marine Propulsors, SMP2013, Tasmania,
Australia, May.
[24] Hally, D., 2015. Propeller Analysis using RANS/BEM
Coupling Allowing for Blade Blockage. Proc. of the
SMP2015. Austin, Texas, USA., July.
[25] Kunz, R., Boger, D., Stinebring, D., Chyczewski, T., Lin-
dau, J., and Gibeling, H., 2000. A Preconditioned Navier-
Stokes Method for Two-phase Flows with Application to
Cavitation. Computer & Fluids, 29, pp. 849875.
[26] Zwart, P. J., 2005. Numerical moddeling of free surface
and cavitating flows. In Industrial Two-phase Flow CFD,

S-ar putea să vă placă și