Improving Propeller Efficiency Through Tip Loading: NOVEMBER 2014
Improving Propeller Efficiency Through Tip Loading: NOVEMBER 2014
Improving Propeller Efficiency Through Tip Loading: NOVEMBER 2014
discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/272021083
CITATION READS
1 212
8 AUTHORS, INCLUDING:
1
30th Symposium on Naval Hydrodynamics
Hobart, Tasmania, Australia, 2-7 November 2014
2
30th Symposium on Naval Hydrodynamics
Hobart, Tasmania, Australia, 2-7 November 2014
3
30th Symposium on Naval Hydrodynamics
Hobart, Tasmania, Australia, 2-7 November 2014
4
30th Symposium on Naval Hydrodynamics
Hobart, Tasmania, Australia, 2-7 November 2014
PROPELLER DESCRIPTIONS (see Figure 9). This design was developed by Bailar
Marine Consulting using NSWCCD design
The NSWCCD tip loaded propeller (TLP) model as processes, including advanced blade sections (Bailar
fabricated by CEHIPAR using bronze is shown in et al. 1991). This propeller was not manufactured, but
Figure 7. It was designed for a spanwise circulation the open water and cavitation performance were
distribution with a maximum at 0.80R. The TLP analyzed using potential flow tools and the viscous
includes a blended, pressure side winglet (positive flow RANS solver, NavyFOAM (Schroeder et al.,
hydrorake) designed using NSWCCD blade shaping 2009).
methods modified for use on TLPs. The blade
sections include a NACA 66 thickness distribution
and a custom chordwise loading distribution with
more trailing edge loading than the more traditional
NACA a=0.8.
Figure 7: NSWCCD tip loaded propeller (TLP). Figure 10 shows spanwise chord and
thickness distributions while Figure 11 shows
The Sistemar contracted-loaded tip propeller spanwise pitch and camber distributions for all three
(CLT) is shown in Figure 8. The CLT was designed propellers. The NSWCCD-designed propellers
using Sistemar’s in-house methods and includes a include a large amount of camber along the entire
pressure side end-plate (positive hydrorake). span, while the CLT propeller has less camber but a
higher pitch especially near the tip. In Figures 10 and
11, “span” is defined as the ratio of the local radial
distance from the hub over the radial distance
between the hub and the tip.
5
30th Symposium on Naval Hydrodynamics
Hobart, Tasmania, Australia, 2-7 November 2014
Figure 11: Spanwise pitch and camber distributions. Figure 12: Blade surface mesh for TLP NavyFOAM full-
scale Reynolds number analyses.
NavyFOAM is an unstructured, finite-
volume based flow solver with RANS and LES FINFLO is also a finite-volume based flow
capabilities. The version used for this project was a solver with RANS capabilities. The implicit solution
single-phase RANS solver in a rotating reference is based on approximately factorized time-integration
frame. One blade passage was gridded and periodic with local time-stepping. In the present
boundary conditions were employed to include the incompressible case the code uses an upwind-biased
effects of the other four blades of the propeller. The approximation for convective fluxes, while either
domain was discretized with a block-structured, Roe's flux-difference splitting or Van Leer's flux-
body-fitted mesh using ANSYS ICEMCFD software. vector splitting is applied for compressible flows. In
Semi-circular trailing edges of 0.2mm radius (at the incompressible case, a Rhie and Chow (1983)
model scale) were enforced on all three propellers. type method has been implemented in the code in the
For both model-scale and full-scale Reynolds simplified form presented by Johansson and
numbers, wall functions were used to simulate the Davidson (1995). The pressure is central-differenced
boundary layer, and a y+ at the wall of 30 was taken and a damping term is added via a convective
as a target. For this study, the H. R. Wilcox k-Omega velocity. A multigrid method is used for the
turbulence model was used. A more complete acceleration of convergence. Solutions on the coarse
6
30th Symposium on Naval Hydrodynamics
Hobart, Tasmania, Australia, 2-7 November 2014
grid levels are used as a starting point for the MODEL HARDWARE AND
calculation in order to accelerate convergence. A
viscous solution is extended to the wall, i.e. no wall EXPERIMENTS
functions were used and the wall y+ was Models of the TLP and the CLT propellers were
approximately 1. Calculations were made using constructed of bronze at 1/26th scale. The diameter of
Chien’s k-epsilon and the SST k-omega turbulence the model-scale propellers was 250mm. The models
models by VTT, however, only the SST k-omega were constructed by the model manufacturing facility
turbulence model results are discussed here. The at the Canal de Experiencias Hidrodinámicas de El
solution of the RANS equations can be obtained by Pardo (CEHIPAR). After fabrication, the two models
either the pseudo-compressibility or pressure were measured using a portable coordinate measuring
correction method. The computations presented in machine in the CEHIPAR manufacturing facility.
this report have been made with the pressure The measurements indicated that the models were
correction method. A more complete description of within the tolerance specified by ITTC
the FINFLO numerical method including recommendations (ITTC procedures 7.5-01-01-01 &
discretization of the governing equations and solution 7.5-01-02-02) and were suitable for hydrodynamic
algorithm can be found in Sánchez-Caja, et al. (1999 testing.
and 2012).
The open water tests were conducted in the
The grids were generated for the FINFLO calm water towing tank at CEHIPAR. The rotational
analyses using an in-house, VTT developed program speed of the models was set at 15 revolutions per
with the help of templates. An O-C topology was second (RPS), and the speed of the carriage was
used around the blades and endplates with the aim of varied in order to obtain thrust and torque values over
reducing the number of cells for a given accuracy. a range of advance coefficients. The Reynolds
This allowed for control of the grid orthogonality number for the open water experiments at the design
over the propeller surfaces. A zero-thickness trailing advance coefficient was approximately 4.32x105.
edge was enforced for both propellers by the gridding
scheme. Figure 13 demonstrates the distribution of The cavitation inception and pressure pulse
cells on the blade surfaces for the CLT propeller. The measurements were conducted in the cavitation
grids around the propeller consisted of 2.4 million tunnel at CEHIPAR. The nominal rotational speed of
cells per blade distributed in 19 blocks for both the the models was again 15 RPS and resulted in similar
full-scale and model-scale calculations. Reynolds numbers as the open water experiments.
Tunnel corrections for measured thrust and torque are
not performed in this case as the model disk area to
tunnel area ratio was 0.061. A thrust identity method
was used to set conditions in the tunnel.
7
30th Symposium on Naval Hydrodynamics
Hobart, Tasmania, Australia, 2-7 November 2014
KT
ETAO
Figure 14: Open water thrust, torque and efficiency of TLP and CLT propellers at model scale.
The nomenclature for thrust, torque and higher-than-design advance coefficients. The
efficiency follows the standard ITTC definition and increased torque in the experiment could be due to
are provided here: the formation of a separated leading edge vortex,
present in the experiments but not fully captured in
𝑇 either NSWCCD’s or VTT’s RANS simulations. The
𝐾𝑇 =
𝜌𝑛2 𝐷4 FINFLO analysis does show a small area of
separation near the leading edge, between r/R values
𝑄 of approximately 0.7 to 0.9, at an advance coefficient
𝐾𝑄 =
𝜌𝑛2 𝐷 5 of 1.108. This separation is indicated by sharp bends
in the surface streamlines in Figure 15. It is possible
𝐽𝐴 ∗ 𝐾𝑇 that the extent of this separated region is under-
𝐸𝑇𝐴𝑜 =
2𝜋 ∗ 𝐾𝑄 predicted in the RANS, causing the slight under-
prediction in the torque at greater than design
advance coefficients.
The data shown in Figure 14 demonstrates The baseline propeller, while not tested in
good agreement between the experiments and both open water, was analyzed using the NavyFOAM flow
RANS solvers, with the NavyFOAM analysis slightly solver. The open water thrust, torque and efficiency
closer to the experimental data for the torque of the values calculated using the NavyFOAM solver for
TLP and CLT propellers. the three designs at the model-scale Reynolds number
are shown in Figure 16.
While the correlation between the RANS
computations and the experiment is generally good, At the design advance coefficient, the TLP
both RANS simulations under-predict the torque of has slightly greater efficiency than the baseline,
the TLP propeller around an advance coefficient of which in turn has slightly greater efficiency than the
1.1, resulting in a slight over-prediction of model- CLT. However, each propeller produces a somewhat
scale efficiency compared to the experiment at different thrust at the design advance coefficient, so
8
30th Symposium on Naval Hydrodynamics
Hobart, Tasmania, Australia, 2-7 November 2014
10KQ
KT
In Pérez Sobrino, et al. (2005) it is asserted
that the ITTC’78 scaling procedure must be enhanced
to capture the scaling effects of CLT propellers,
because the ITTC’78 procedure uses a similarity
assumption that reduces the spanwise quantities to a
Figure 16: NavyFOAM open water results for the TLP single value of 0.75R. For CLT propellers this
(filled squares), CLT (triangles) and Baseline (hollow assumption is not valid due to the unique tip
squares) at full-scale Reynolds number. geometry and special radial load distribution with
non-zero load at the tip that leads to rather different
Additional numerical simulations using FINFLO with flow characteristics compared to conventional
the k-epsilon model were performed by VTT in order propellers at model scale, so corrections to the
to check the influence of partially laminar flow on the ITTC’78 procedure were proposed.
results. Further discussion on this topic can be found
in Sánchez-Caja, et al. (2014). The present work aims to further this
research by applying the scaling procedure of Pérez
Sobrino to a new CLT design as well as the
9
30th Symposium on Naval Hydrodynamics
Hobart, Tasmania, Australia, 2-7 November 2014
NSWCCD-developed TLP design, and to compare percent. The NavyFOAM-calculated thrust and
the results of the scaling procedure with multiple torque, however, are approximately 5% higher than
RANS assessments. Hsin, et al. (2010) found that the scaled experimental data.
RANS assessments showed an additional scaling
effect for propellers with end-plates.
∆CF = - 0.00169
κ = 0.0094
End plate form factor = 0.5; 1+k= 1.5.
ETAO
The same correlation coefficients were
applied for the TLP and CLT propellers. Figure 18
shows the open water curves from the experiments
corrected by the Sistemar procedure, as well as the
full-scale computational results from FINFLO and
NavyFOAM for the TLP. Figure 19 shows the open Figure 19: Full-scale open water predictions for CLT from
water curves from the experiment corrected by the Sistemar scaling procedure (lines), FINFLO full-scale
Sistemar procedure, as well as the full-scale analysis (hollow points) and NavyFOAM (solid points).
computational results from FINFLO and NavyFOAM
for the CLT.
Figure 20 shows the scaling corrections as a
percent change of KT, KQ and ETAo (at the design
advance coefficient of 0.923) as predicted by the
10KQ various methods for the TLP.
KT
ETAO
10
30th Symposium on Naval Hydrodynamics
Hobart, Tasmania, Australia, 2-7 November 2014
The integrated pressure and shear stress for thrust and show velocity in the chord-wise direction normalized by
torque predicted by NavyFOAM were examined inflow velocity.
separately to better understand the differences
between the model-scale and full-scale computations.
If only changes in the viscous shear stresses are
considered, the model-to-full-scale corrections
predicted by NavyFOAM are reduced greatly. These
values are shown in Figure 20, labeled “NavyFOAM
(Cf only)”. Clearly, there must be a Reynolds number
effect on the blade pressure distributions not captured
by the friction-coefficient-based ITTC’78 method,
and possibly captured to a lesser extent by the
FINFLO RANS solver and Sistemar method.
11
30th Symposium on Naval Hydrodynamics
Hobart, Tasmania, Australia, 2-7 November 2014
viscous scaling predicted by the RANS is greater a greater extent than predicted by only a change in
than the ITTC’78 method due to the fully-turbulent the friction coefficient.
boundary layer assumption in the NavyFOAM solver.
In contrast to the TLP, for the CLT propeller the
overall scaling corrections predicted by the two
RANS methods are much closer to each other than
for the TLP. This result opens the question: if the
difference between the NavyFOAM and the FINFLO
results for the TLP was due to a difference in trailing
edge geometry (that is also present in the CLT
analyses), why do the two solvers correlate better for
the CLT? In an attempt to answer this question, the
NavyFOAM CLT solution was interrogated using the
same 0.7R constant-radius slice procedure as the
TLP. The model-scale and full-scale flows are shown
in Figure 24 and Figure 25, respectively.
12
30th Symposium on Naval Hydrodynamics
Hobart, Tasmania, Australia, 2-7 November 2014
13
30th Symposium on Naval Hydrodynamics
Hobart, Tasmania, Australia, 2-7 November 2014
versus the thrust loading coefficient (CT) is shown in The cavitation inception data demonstrates that the
Figure 31. TLP had no suction side cavitation, but a tip vortex at
the design point. The CLT propeller had much earlier
suction side cavitation inception due to its greater
pitch and higher sectional angle of attack.
VA= 8.0m/s
Figure 30: Effect of USN and Sistemar TE bevels on open Figure 32: TLP cavitation inception curves.
water coefficients.
VA= 8.0m/s
14
30th Symposium on Naval Hydrodynamics
Hobart, Tasmania, Australia, 2-7 November 2014
CONCLUSIONS
In this paper, the authors have presented a
comparative design and performance of tip loaded
propellers, TLP and CLT, with a particular focus on
Reynolds number scaling effects on open water
performance. The computational results showed good
agreement with experimental data, and the
differences between the results have been discussed.
From this study the authors have concluded that:
Figure 35: Cavitation pattern on CLT propeller at design
point (JA=0.94, sigma=4.092). - There is definite potential for improving
efficiency and cavitation inception through
the implementation of tip loaded propellers,
15
30th Symposium on Naval Hydrodynamics
Hobart, Tasmania, Australia, 2-7 November 2014
as demonstrated by the higher efficiency and ONR Code 331, under the administration of Dr. Ki-
cavitation inception of the TLP compared Han Kim.
with the baseline propeller. The CLT
propeller matched the baseline propeller for
efficiency at the design thrust coefficient. REFERENCES
However, its efficiency was reduced due to a Adalid, J.G., and Gennaro, G., “Latest experiences
separated leading edge vortex caused by the with Contracted and Loaded Tip (CLT) propellers,”
large sectional angle-of-attack at the design Sustainable Maritime Transportation and
condition. It is expected that a redesign of Exploitation of Sea Resources, 1st ed., Vol. 1, Taylor
the CLT propeller with reduced angle-of- & Francis Group, London, 2011, pp. 47-53.
attack loading would have a higher
efficiency. Andersen, P., Friesch, J., Kappel, J.J., Lundegaard,
- The scale effects have been investigated and L., and Patience, G., “Development of a Marine
demonstrated to come from multiple sources Propeller With Nonplanar Lifting Surfaces,” Marine
(viscous and pressure effects as well as the Technology, Vol. 42, issue 3, 2005, pp. 144-158.
presence of separation). The computational
results show a large sensitivity to trailing Bailar, J.W., Jessup, S.D., and Shen, Y.T.,
edge details (i.e. sharp vs. rounded) in “Improvement of Surface Ship Propeller Cavitation
shaping the flow at model-scale Reynolds Performance Using Advanced Blade Sections,”
numbers. The computational results also Proceedings of the 23rd American Towing Tank
demonstrate the loss of lift due to camber Conference, vol. 1, 1991, pp. 185-193.
and trailing edge loading at model scale. The
TLP and the CLT have been identified as Hsin, C.-Y., Chang, K.-K., Chi, R.-C., and Chen, P.-
having an increased scaling effect due to F., “Design and Analysis of the End-Plate Effect
added wetted surface area in the outer span Propellers,” Proceedings 28th Symposium on Naval
region. Hydrodynamics, 2010.
In the effort to further improve and innovate Hsin, C.-Y., Kerwin J.E., and Kinnas, S.A., "A Panel
in the field of propeller design, the concepts Method for the Analysis of the Flow around Highly
demonstrated here should continue to be studied and Skewed Propellers," Proceedings of the SNAME
developed. The design-by-analysis approach Propellers '91 Symposium, Society of Naval
employed during the detail design of the TLP Architects and Marine Engineers, vol. 1, 1991.
propeller generated a successful design by current
measures, but perhaps a superior design could be Johansson, P., and Davidson, L., “Modified
achieved with better design methods and tools Collocated SIMPLEC Algorithm Applied to
specific to the case of the tip loaded propeller. Buoyancy-Affected Turbulent Flow Using a Multi-
Furthermore, while the current study was limited to Grid Procedure,” Numerical Heat Transfer, Part B,
open water design and performance, it is Vol. 28, pp. 39-57, 1995.
recommended that the performance of tip loaded
propellers in the ship wake (i.e. non-uniform inflow) Kerwin, J.E., Michael, T.J., and Neely, S.K.,
be investigated to assess the effects of more realistic “Improved Algorithm for the Design/Analysis of
flow fields on the efficiency and cavitation Multi-Component Complex Propulsors,” Proceedings
performance. of the Propellers/Shafting Symposium, Society of
Naval Architects and Marine Engineers, vol. 1, 2006.
16
30th Symposium on Naval Hydrodynamics
Hobart, Tasmania, Australia, 2-7 November 2014
17