CFD Analysis of Scale Effects on Conventional and Tip-Modified Propellers

Similar documents
Propeller Loads of Large Commercial Vessels at Crash Stop

Influence of Flow Transition on Open and Ducted Propeller Characteristics

On the advanced extrapolation method for a new type of podded propulsor via CFD simulations and model measurements

Numerical Analysis of Unsteady Open Water Characteristics of Surface Piercing Propeller

VERIFICATION AND VALIDATION OF RESISTANCE AND PROPULSION COMPUTATION

ENGR 4011 Resistance & Propulsion of Ships Assignment 4: 2017

A Study on Effects of Blade Pitch on the Hydrodynamic Performances of a Propeller by Using CFD

Numerical calculations of the hydrodynamic performance of the contra-rotating propeller (CRP) for high speed vehicle

Numerical and Experimental Characterization of a CP Propeller Unsteady Cavitation at Different Pitch Settings

Turbulent Boundary Layers & Turbulence Models. Lecture 09

RANSE CODE APPLICATION FOR DUCTED AND ENDPLATE PROPELLERS IN OPEN WATER

Active Control of Separated Cascade Flow

HYDRODYNAMIC PREDICTION OF THE DUCTED PROPELLER BY CFD SOLVER

Deliverable D.6.1. Application of CFD tools to the development of a novel propulsion concept

Potsdam Propeller Test Case (PPTC) Test Case Description

CFD Analysis of a Propeller Flow and Cavitation

Manhar Dhanak Florida Atlantic University Graduate Student: Zaqie Reza

Calculation of the Flow around the KVLCC2M Tanker

Hydro-Structure Analysis of Composite Marine Propeller under Pressure Hydrodynamic Loading

Pre-Swirl Stator and Propeller Design for Varying Operating Conditions

Reliability assessment of ship powering performance extrapolations using Monte Carlo methods

Explicit algebraic Reynolds stress models for internal flows

Full scale thruster performance and load determination based on numerical simulations

Masters in Mechanical Engineering Aerodynamics 1 st Semester 2015/16

CHAPTER 4 OPTIMIZATION OF COEFFICIENT OF LIFT, DRAG AND POWER - AN ITERATIVE APPROACH

Study on Extrapolation Method for Self-Propulsion Test with Pre-Swirl Device *

An Extrapolation Method Suitable for Scaling of Propellers of any Design

DETERMINING OPTIMUM GEOMETRY FOR A BOW THRUSTER PROPELLER

Mestrado Integrado em Engenharia Mecânica Aerodynamics 1 st Semester 2012/13

COMPUTATIONAL FLOW ANALYSIS THROUGH A DOUBLE-SUCTION IMPELLER OF A CENTRIFUGAL PUMP

ITTC Propeller Benchmark

DEVELOPMENT OF CFD MODEL FOR A SWIRL STABILIZED SPRAY COMBUSTOR

Propeller Analysis Using RANS/BEM Coupling Accounting for Blade Blockage

Numerical Investigation of the Impact of SES-Waterjet Interactions and Flow Non-uniformity on Pump Performance

Analysis of Crashback Forces Compared with Experimental Results

Yiran Su 1, Seungnam Kim 1, Weikang Du 1, Spyros A. Kinnas 2, Mikael Grekula 3, Jan Hallander 3, Da- Qing Li 3

ADVANCES IN FULL-SCALE WAKE-FIELD PREDICTIONS AND THE IMPLICATIONS FOR THE PROPELLER DESIGN

LARGE EDDY SIMULATION OF FLOW OVER NOZZLE GUIDE VANE OF A TRANSONIC HIGH PRESSURE TURBINE

ITTC Propeller Benchmark

ASSESSMENT OF RANS AND DES METHODS FOR THE AHMED BODY

FLOW CHARACTERISTICS IN A VOLUTE-TYPE CENTRIFUGAL PUMP USING LARGE EDDY SIMULATION

Numerical study of the effects of trailing-edge bluntness on highly turbulent hydro-foil flows

Parallel Computations of Unsteady Three-Dimensional Flows in a High Pressure Turbine

GPPS NUMERICAL PREDICTION OF UNSTEADY ENDWALL FLOW AND HEAT TRANSFER WITH ONCOMING WAKE

Research on Dynamic Stall and Aerodynamic Characteristics of Wind Turbine 3D Rotational Blade

Flow analysis in centrifugal compressor vaneless diffusers

A New Implementation of Vortex Lattice Method Applied to the Hydrodynamic Performance of the Propeller-Rudder

Fluid Dynamics Exercises and questions for the course

Masters in Mechanical Engineering. Problems of incompressible viscous flow. 2µ dx y(y h)+ U h y 0 < y < h,

Efficient Prediction of Crashback Performance of Controllable Pitch Propellers

Cavitation of a Propeller and Influence of a Wake Equalizing Duct

BOUNDARY LAYER FLOWS HINCHEY

Large Eddy Simulation of Crashback in Marine Propellers

Effects of the advance ratio on flow structures around marine propeller

A Numerical study of effect of Return Channel Vanes Shroud Wall Divergence Angle on the Cross-over System Performance in Centrifugal Compressors

Resistance and Propulsion

Offshore Hydromechanics Module 1

Study of the hydrodynamic flow around a 70m sailing boat for powering, wave pattern and propeller efficiency prediction

AN EXPLORATORY STUDY ON THE WORKING PRINCIPLES OF ENERGY SAVING DEVICES (ESDS) PIV, CFD INVESTIGATIONS AND ESD DESIGN GUIDELINES

Numerical Validation of Flow Through an S-shaped Diffuser

COMPARATIVE STUDY BETWEEN FLOWS AROUND SPHERE AND POD USING FINITE VOLUME METHOD

WALL ROUGHNESS EFFECTS ON SHOCK BOUNDARY LAYER INTERACTION FLOWS

Numerical Analysis of Azimuth Propulsor Performance in Seaways: Influence of Oblique Inflow and Free Surface

Lecture-4. Flow Past Immersed Bodies

COMPUTATIONAL SIMULATION OF THE FLOW PAST AN AIRFOIL FOR AN UNMANNED AERIAL VEHICLE

CHAPTER 7 NUMERICAL MODELLING OF A SPIRAL HEAT EXCHANGER USING CFD TECHNIQUE

Computational Analysis of Marine-Propeller Performance Using Transition-Sensitive Turbulence Modeling

Aerodynamics. High-Lift Devices

Numerical Prediction Of Torque On Guide Vanes In A Reversible Pump-Turbine

TURBULENT FLOW ACROSS A ROTATING CYLINDER WITH SURFACE ROUGHNESS

Wall treatments and wall functions

Analysis of Unsteady Propeller Blade Forces by RANS

STAR-CCM+: NACA0012 Flow and Aero-Acoustics Analysis James Ruiz Application Engineer January 26, 2011

Calculation of propeller s load noise using LES and BEM numerical acoustics coupling methods

Evaluation of Propeller Virtual Mass and Damping Coefficients by URANS- Method

Scaling of Tip Vortex Cavitation Inception for a Marine Open Propeller

Propeller geometry optimization for pressure pulses reduction: an analysis of the influence of the rake distribution

Validation 3. Laminar Flow Around a Circular Cylinder

ITTC Recommended Procedures and Guidelines Testing and Extrapolation Methods Propulsion, Performance Propulsion Test

Energy balance analysis of a propeller in open water

EFFECT OF FORCED ROTATING VANELESS DIFFUSERS ON CENTRIFUGAL COMPRESSOR STAGE PERFORMANCE

AN UNCERTAINTY ESTIMATION EXAMPLE FOR BACKWARD FACING STEP CFD SIMULATION. Abstract

Numerical Studies of Supersonic Jet Impingement on a Flat Plate

Study of the Losses in Fluid Machinery with the Help of Entropy

Confined water effects on the viscous flow around a tanker with propeller and rudder

Performance Analysis and Design of Vertical Axis Tidal Stream Turbine

On the Use of the γ Re θ Transition Model for the Prediction of the Propeller Performance at Model-Scale

Numerical analysis of propeller exciting force in oblique flow

CFD approach for design optimization and validation for axial flow hydraulic turbine

NUMERICAL SIMULATION OF STATIC INFLOW DISTORTION ON AN AXIAL FLOW FAN

Numerical Investigation of Laminar Flow over a Rotating Circular Cylinder

Actuator Surface Model for Wind Turbine Flow Computations

Chapter three. Two-dimensional Cascades. Laith Batarseh

Prediction of Propeller Performance Using Quasi-Continuous Method


The Calculations of Propeller Induced Velocity by RANS and Momentum Theory

Numerical Simulation of Rocket Engine Internal Flows

CFD RANS analysis of the rotational effects on the boundary layer of wind turbine blades

Keywords - Gas Turbine, Exhaust Diffuser, Annular Diffuser, CFD, Numerical Simulations.

Self-Excited Vibration in Hydraulic Ball Check Valve

Transcription:

Fifth International Symposium on Marine Propulsors smp 17, Espoo, Finland, June 2017 CFD Analysis of Scale Effects on Conventional and Tip-Modified Propellers Keun Woo Shin 1, Poul Andersen 2 1 Propeller & Aftship R&D Department, MAN Diesel & Turbo, Frederikshavn, Denmark 2 Department of Mechanical Engineering (MEK), Technical University of Denmark (DTU), Kgs.Lyngby, Denmark ABSTRACT Full-scale propeller performance is traditionally predicted by scaling model-scale test results, but the traditional scaling methods do not take into account hydrodynamic distinctions of tip-modified propellers in full-scale performance. An open-water CFD analysis is made on scale effects of tip-modified and conventional propellers, which are designed for the same operating condition with identical propeller diameter and expanded area ratio. While model-scale computations are made with a transition model, a fully turbulent flow is modeled in the full-scale computations. The investigation on the effects of the transition model shows that laminar and transitional flow modeling is crucial in model-scale computations. Grid-independent solutions at model and full scale are achieved by grid verification studies. The CFD analysis of scale effects shows that the efficiency gain of the tip-modified propeller is increased at full scale. The difference of scale effects between the tip-modified and conventional propellers is related to alterations of tip vortex and sectional pressure distributions by the bent tip and the higher spanwise loading at the tip region of the tip-modified propeller. Keywords Scale effect, Unconventional propeller, CFD, RANS, Transition model 1 INTRODUCTION The standard ITTC 1978 scaling method for open-water propeller characteristics adopted by most model test institutes is an extrapolation using Reynolds numberbased friction corrections of a representative blade section (ITTC 2008). RANS computations on conventional propellers with different skews and area ratios at model and full scales have shown that ITTC78 scaling method does not reflect larger scale effects for propellers with higher skews and larger area ratios (Krasilnikov et al 2009). Model tests and RANS computations on modelscale low- and high-skew conventional propellers with varying Reynolds number Rn have also shown larger scale effects for the high-skew propeller (Funeno 2002). Tip-modified propellers have been developed to improve propulsive efficiency and to increase lifting-surface span without shortening the clearance from the propeller tip to the hull surface. Nowadays research on tip-modified propellers is active with increased demand for reducing fuel consumption and greenhouse gas emissions in ship operations. Efficiency improvements of tip-modified propellers are made by reducing vortex shedding across the blade tip and redistributing the spanwise blade loading. As the intensity of tip vortex varies from model scale to full scale depending on Rn, scale effects on tipmodified propellers are more critical than on conventional propellers. Two tip-modified propellers and a conventional propeller have been designed for a set of common requirements. Open-water CFD investigations of scale effects on those three propellers by using two different RANS solvers have shown larger scale effects for the tip-modified propellers than for the conventional one (Brown et al 2014). A tip-modified propeller has been retrofitted on a tanker in lieu of an original conventional propeller. Selfpropulsion model tests and sea trials on both the propellers have shown slightly larger scale effects on the tip-modified propeller (Andersen et al 2005). Figure 1: Open-water efficiency η O,KT at the design thrust coefficient K T with respect to Rn 0.7 from CFD on tip-modified and conventional propellers (Shin & Andersen 2016) and from model tests on another conventional propeller (Bugalski et al 2013)

RANS computations have been made on tip-modified and conventional propellers designed for an identical operating condition with common geometrical parameters at a range of Rn 0.7 10 5 10 6 corresponding to those in self-propulsion and open-water propeller model tests (Shin & Andersen 2016), where Rn 0.7 = C 0.7 (0.7 π D N) 2 +V A 2 and C 0.7 is the chord length at 0.7R, D is the propeller diameter, N is the propeller rate of revolution, V A is the propeller advance speed and ν is the kinematic viscosity. Open-water model tests have been made on another conventional propeller with varying Rn 0.7 (Bugalski et al 2013). As Rn 0.7 is increased, open-water efficiency η O,KT at the design thrust coefficient K T rises up to a certain point and drops down for the two conventional propellers, as shown in Figure 1. As η O,KT for the tip-modified propeller constantly increases within the considered range of Rn 0.7, scale effects become relatively larger for the tipmodified propeller. In this research work, the CFD analysis of scale effects on tip-modified and conventional propellers is extended to a higher value of Rn 0.7 = 1.3 10 7 corresponding to full-scale propeller operations. Rn 0.7 is varied in a range of Rn 0.7 1.8 10 5 2.3 10 7 with keeping a constant Froude number, whereas Rn 0.7 is increased without conforming to Froude s law in Figure 1. The correlation of scale effects with viscous shear stress, non-viscous pressure and induced drag related to tip vortex is analyzed for both tipmodified and conventional propellers. RANS computations are conducted with and without a transition model to investigate the effects of laminar and transitional flows. CFD on model-scale propellers is often validated against model test results, but it is hard to find full-scale propeller performance measurements in open-water condition, which can be used for full-scale CFD validations. Therefore, grid independency of CFD results is highlighted for reliable full-scale CFD. Systematic grid verification studies have been made for full-scale hull simulations in bare-hull and self-propulsion conditions (Visonneau et al 2016). Grid independency has been studied by using three structured grids for model- and full-scale open-water propellers (Sanchez-Caja et al 2014). Unstructured grids have shown difficulty to obtain grid-independent results for full-scale propellers (Haimov et al 2011). Grid-independent solutions for model- and full-scale tip-modified propellers with respect to overall grid size and boundary layer flow resolution are attempted by using unstructured grids and improving grid similarity. Grid-independent results for the model-scale propeller are validated against open-water model test results. 2 PROPELLER MODELS Tip-modified and conventional propellers on a singlescrew 35,000 DWT bulk carrier are considered for CFD analysis of scale effects. Both propellers consisting of 4 ν (1) blades with identical diameter and expanded area ratio of D F = 5.9 m and A E /A O = 0.38, respectively, are designed for the same operating condition of V S = 14 kn and N = 91.3 rpm. The tip-modified propeller is a Kappel propeller characterized by a smoothly curved tip bending towards the suction side of the blade. The comparison of tip rake, spanwise load distribution and other geometrical parameters between two propellers are illustrated in Shin & Andersen (2016). A model propeller of D M = 0.25 m for the tip-modified propeller has been produced and evaluated in open-water and self-propulsion tests in Force Technology, Denmark. Figure 2: Experimental model (top) of tip-modified propeller and computational models of tip-modified (bottom left) and conventional (bottom right) propeller 3 CFD SETUP A steady incompressible RANS solver in the commercial CFD package StarCCM+ is adopted with curvaturecorrected k-ω SST turbulence model excluding a wall function for CFD simulations. Since laminar and transitional boundary layer flows are formed on a modelscale propeller, γ-re θ transition model is also adopted. The effects of laminar and transitional flows on modeland full-scale propellers are investigated by comparing CFD results with and without the transition model. A single blade is modeled with front hubcap and downstream shaft following the open-water test setup in a quarter-cylinder fluid domain with periodic boundary condition on the sides. The quarter-cylinder domain extends 3 D from the propeller plane to the inlet and 6 D to the outlet with a radius of 4 D. A moving reference frame for modeling propeller rotation is applied to an inner domain extending from the propeller plane 0.4 D upstream and 2.4 D downstream with a radius of 0.6 D. Computational grids are generated by unstructured trimmed hexahedral meshing with including prism layers on wall surfaces. The surface mesh size on the overall surface of the blade and hub is Δx M = 0.5 1.0 mm at model scale and Δx F = 2.0 4.0 mm at full scale. The grid

size increases twice after 10 layers. The maximum grid size is 7 Δx at the blade tip in Zone 1, 14 Δx around the rotating domain in Zone 2 and 35 Δx in Zone 3, as shown in Figure 3. The grid is refined to 0.5 Δx along the blade edges for resolving high curvature. Figure 3: Computational domain and grid on a longitudinal cross section 1. The cell height increases with 1.2 stretching rate in the prism layers. 4 GRID VERIFICATION A grid verification study is made with respect to Δx and Δh for model- and full-scale computations. Grid similarity is ensured by constant growth rates from the surface mesh to the outer volume mesh and zonal grid sizes defined as functions of Δx. The grid study is conducted only for the tip-modified propeller at the design advance ratio of J = 0.6. The full-scale grid study is made at the actual propeller speed of N = 1.6 rps in the design condition, whereas a higher speed of N = 20.0 rps than N = 7.8 rps scaled by Froude s law is applied to the model-scale grid study in the same way as in the open-water model test. The full-scale grid study is made without the transition model, because the boundary layer flow on a full-scale propeller is considered to be fully turbulent except for a small region at the leading edge. The model-scale grid study is conducted with and with the transition model. The model-scale grid study with the transition model is looked into first. Figure 4: K T and K Q as a function of normalized grid size for CFD on the tip-modified propeller at model and full scale The prismatic mesh consists of 9 layers with a total thickness of 0.2 mm at model scale and 0.14 mm at full scale. The first-cell height is Δh M = 9.6 μm at model scale and Δh F = 6.5 μm at full scale, which leads mostly to y+ Figure 5: K T and K Q as a function of normalized prism-layer thickness for CFD on the tip-modified propeller at model and full scale Δx M = 0.5 1.0 mm and Δh M = 9.6 μm are adopted for the final model-scale grid after the grid verification study. Figure 4 and 5 show the variations of K T and K Q at Δx/Δx M = 1.0 5.0 and Δh/Δh M = 0.5 4.0 for model-scale computations. Although a constant convergence rate is

Conventional Pressure side Suction side Tip-modified Pressure side Suction side not shown due to unstructured grids, both K T and K Q are converged with variances of less than 0.3% at 5 grid sizes of Δx/Δx M = 1.0 1.55. K T and K Q are also converged with respect to Δh/Δh M with variances of less than 0.8% at 5 prism-layer thicknesses of Δh/Δh M = 0.5 1.13, but the variances for Δh/Δh M are slightly larger than those for Δx/Δx M. The converged values of K T and K Q are close to K T = 0.206 and K Q = 0.303 from the model test with errors of 0.2% and 0.6%, respectively. Δx F = 2.0 4.0 mm and Δh F = 6.5 μm are adopted for the final full-scale grid after the grid verification study. While Δx F of the full-scale grid is 4 times larger than Δx M of the model-scale grid, Δx F /D F is about 6 times smaller than Δx M /D M, so the total number of computational cells is 17.8 million for the full-scale computation, which is about 3.5 times higher than 5.2 million for the model-scale computation. The variations of K T and K Q at Δx/Δx F = 1.0 9.0 and Δh/Δh F = 0.5 7.0 for full-scale computations are also shown in Figure 4 and 5. Both K T and K Q are converged with variances of less than 0.2% for 3 grid sizes of Δx/Δx F = 1.0 1.7. Although K T and K Q are seemingly converged with variances of less than 0.5% for 5 prism-layer thicknesses of Δh/Δh F = 1.0 1.8, K T and K Q continuously increase for smaller prism-layer thicknesses of Δh/Δh F 1.0, probably because turbulent boundary layer is unevenly resolved for thinner prism layers. The converged values of K T and K Q for full-scale computations are higher than those for model-scale computations by 9.9% and 6.7%, respectively. 3.5%. The K Q error is also increased to 1.0%, but the increase is not so much as that of the K T error. In Figure 6, the open-water curves from the computations on the final grid with and without the transition model are compared with that from the model test. While the deviations in K T, K Q and η O from the experimental result are within 2% at J = 0.2-0.6 for the computations with the transition model, the underestimations of K T and K Q become larger at low blade loadings of J 0.7. The underestimation of K T is more pronounced for the computations without the transition model at J 0.6, which leads to the underestimation of η O. In Figure 7, the surface-constrained streamlines and the skin friction coefficient C F on the blade surface of both propellers are compared between model-scale computations with and without the transition model on the final grid and a preliminary grid with a lower boundary layer resolution where y+ is about twice the value of the final grid, where C F = and τ W is the wall shear stress. τ W 0.5 ρ N 2 D 2 (2) Without With transition model transition model Final grid Preliminary grid Final grid Figure 6: Open-water curves from the model test and CFD with and without the transition model for the tip-modified propeller 5 TRANSITION MODEL When looking into the model-scale grid study without the transition model, K T and K Q are also converged for Δx/Δx M and Δh/Δh M in a similar way as those with the transition model, as shown in Figure 4 and 5, but K T is not converged for Δh/Δh M and it continuously increases for smaller prism-layer thicknesses of Δh/Δh M 1. The K T error of the computations without the transition model from the model test result is significantly increased to Figure 7: Surface-constrained streamlines and skin friction coefficient from the model-scale computations When comparing the streamlines from the final and preliminary grids, the turbulent flow separation indicated by radially outward deflections of tangential flows on the suction side near the trailing edge is more pronounced for the final grid with about twice higher boundary layer resolution than the preliminary grid. The tangential flow

With transition model Without transition model on the suction side before the flow separation is along rather constant radii for the preliminary grid like that from the computation without the transition model, whereas it is directed slightly upwards for the final grid. The difference of the pressure-side flow is relatively small between the final and preliminary grids because laminarturbulent transition is more gradual due to the favorable pressure gradient than on the suction side. Figure 9 can be related to the thrust increase, whereas the suction gradually falls down from about 0.7 C without the transition model. The K Q difference is up and down within 1.5%, because the torque increase from nonviscous pressure is on a similar level as the torque reduction from viscous shear stress. In Figure 10, vortex flows visualized by an iso-surface of Q-criterion Q = 30,000 show that the tip vortex extent is slightly larger for the computation with the transition model, which leads to larger induced drag from the tip vortex and so it can contribute to the larger torque from non-viscous pressure. Figure 9: Sectional pressure distribution at 0.8 R (C P = (P-P )/0.5 ρ N 2 D 2 ) Figure 8: Differences of K T and K Q between computations on the tip-modified propeller with and without the transition model (Δϕ = ϕ Transition / ϕ NoTransition 1) When excluding the transition model, the friction on the overall blade surface increases especially at the leading edge and the flow separation at the trailing edge of the suction side and at inner radii of the pressure side comparatively is reduced, compared to the computation result with the transition model. In Figure 8, K T and K Q are compared between the computations on the tipmodified propeller with and without the transition model in a range of J = 0.2 0.8. K T from the transition-modeled computation is larger than that from the fully-turbulent flow computation in the whole range of J and ΔK T increases at higher J, because the thrust increase from non-viscous pressure is larger than the thrust reduction from viscous shear stress and the difference is increased at higher J. When laminar and transitional flows are modeled, the high suction extended to 0.83 C at 0.8 R in Figure 10: Vortex flows visualized by iso-surfaces of Q- criterion Q = 30,000 with colors indicating the vorticity component around the axial direction on the tip-modified propeller with / without the transition model A full-scale computation with the transition model is made only for the final grid. The differences of K T and K Q are within 0.4% between the full-scale computations with and without the transition model. In Figure 11, the difference in the streamlines is not noticeable between the computations with and without the transition model, whereas C F is lower at the leading edge of both sides and higher at the mid-chord and the trailing edge of the suction side in the computation with the transition model. As C F at full scale in Figure 11 is lower than that at model scale in Figure 7, the portion of the torque from viscous shear stress to the total torque is 1.5 1.7% in the full-

Conventional Pressure side Suction side Tip-modified Pressure side Suction side scale computations, which is smaller than 3.3 6.4% in the model-scale computations. It is to be noted that the upper limit of C F in Figure 7 is 5 times higher than that in Figure 11. The difference in the torque from viscous shear stress between the full-scale computations with and without the transition model is within 0.2% of the total torque, which is also smaller than about 3% in the modelscale computations. In Figure 11, C F from the full-scale computation on a preliminary grid of Δx/Δx F = 2 is lower than that on the final grid. While streamlines are formed smoothly along constant radii for the final grid, it deflects radially outwards at inner radii and some are not smooth for the preliminary grid, which demonstrates that flow separation can occur numerically due to insufficient grid resolution. With Without transition model transition model Final grid Preliminary grid Final grid of K T is larger than that of K Q, η O at full scale is higher than that at model scale. Figure 12: Open-water curves from CFD for the tip-modified and conventional propellers at model and full scales Figure 11: Surface-constrained streamlines and skin friction coefficient from full-scale computations 6 SCALE EFFECT Computations are made on the tip-modified and conventional propellers at model and full scale by using the final grid from the grid verification study. Laminar to turbulent flow transitions are modeled by adopting the transition model in model-scale computations, whereas a fully turbulent flow is modeled in full-scale computations without the transition model. In Figure 12, K T and K Q at model scale are higher for the tip-modified propeller at the design condition and low loadings of J 0.6 and lower at high loadings of J 0.4. At full scale, K T and K Q for both propellers are increased over the whole considered range of J. Since the increase Figure 13: Differences of K T and K Q between model- and fullscale computations on tip-modified and conventional propellers (Δϕ = ϕ Full / ϕ Model 1) In Figure 13, the increases of K T and K Q from model to full scale are presented with portions from viscous shear stress and remaining pressure. The K T increase is larger for the tip-modified propeller at the design condition and

Conventional Model scale Tip-modified Conventional Model scale Tip-modified high loadings of J 0.6 and the difference of ΔK T between the two propellers become larger at lower values of J, while ΔK T is smaller for the tip-modified propeller at low loadings of J 0.7. ΔK Q is larger for the conventional propeller in the entire range of J and the difference of ΔK Q is increased at high values of J. As ΔK T is larger and ΔK Q is smaller for the tip-modified propeller at the design condition and high loadings of J 0.6, the difference of η O between the two propellers becomes larger at full scale. At the design condition of J = 0.6, η O is 1.6% higher for the tip-modified propeller at model scale, whereas the difference is increased to 2.7% at full scale. Both ΔK T and ΔK Q are larger for the conventional propeller at low loadings of J 0.7, but the difference of ΔK T is relatively larger than that of ΔK Q and so η O at full scale is higher for the conventional propeller at J 0.9. ΔK T and ΔK Q are mostly due to non-viscous pressure excluding viscous shear stress. The portions of K T and K Q from viscous shear stress are reduced at full scale, but the reduction is only within 0.3% for ΔK T. The reduction of K Q portion from viscous shear stress is 1.6 2.4%, which is relatively small compared to 5.6-10.5% increase of the portion from remaining pressure. pressure for the conventional propeller, whereas such fluctuations are weak and confined to a small region near the trailing edge for the tip-modified propeller. As the spanwise blade loading is more distributed towards the tip for the tip-modified propeller, the circulation at the maximum loading region of 0.6R 0.8R is 10.6 14.7% lowered compared to that for the conventional propeller. The comparison of the circulation distribution between the two propellers is found in Shin & Andersen (2016). The stronger flow separations related to the higher loading at 0.6R 0.8R can lead to the efficiency loss of the conventional propeller. Figure 14: Sectional pressure distributions at 0.6R (top) and 0.8R (bottom) for the tip-modified and conventional propellers at model and full scale The comparison of sectional pressure distributions between model and full scale in Figure 14 shows differences in the recovery pattern of the rooftop pressure on the suction side. While the rooftop pressure drops down with a high pressure gradient at x/c = 0.8 0.9 at model scale, it is recovered from x/c 0.7 with a rather moderate pressure gradient at full scale. At full scale, pressure fluctuations indicating flow separations occur at x/c = 0.9 0.95 in the recovering process of the rooftop Figure 15: Vortex flows visualized by iso-surfaces of Q-criteria Q = 30,000 at model scale and Q = 240 at full scale with colors indicating the vorticity component around the axial direction In Figure 15, the Q-criterion value for visualizing vortex flows at full scale is reduced more than 100 times from that at model scale so that the tip vortex extent normalized by the propeller diameter may be the same at both scales for the tip-modified propeller. The relative length of tip vortex is incomparable between model and full scale from iso-surfaces of an identical Q-criterion value mainly due to a higher propeller speed applied to the model-scale computations than that from Froude scaling. The tip vortex for the conventional propeller is slightly more

extended at full scale, which can bring additional efficiency loss on top of that at model scale. Model-scale computations are repeated on both the propellers at a lower Rn from Froude scaling than that in the open-water model test. Computations at 3 intermediate scales between model and full scale are made with including the transition model. In Figure 16, the open-water efficiencies η O,CTh and η O,KT interpolated at the design thrust loading coefficient C Th = 1.37 and the design thrust coefficient K T = 0.20, respectively, are presented with respect to Rn 0.7. Both η O,CTh and η O,KT are increased and the increasing rates are gradually reduced at higher Rn 0.7. The overall increasing rate of η O,CTh is lower than that of η O,KT and the increase of η O,CTh for each propeller is within 0.5% at Rn 0.7 2.6 10 6. Figure 16: Open-water efficiencies at C Th = 1.37 and K T = 0.20 with varying the scale factor from model to full scale The differences of η O,CTh and η O,KT between the two propellers are increased at higher Rn 0.7. While Δη O,CTh = 1.9% at Rn 0.7 = 1.8 10 5 is increased to 2.8% at Rn 0.7 = 2.3 10 7, Δη O,KT of 3.1 3.6% is larger than Δη O,CTh at the entire range of Rn 0.7, but the variation of Δη O,KT from model to full scale is smaller, where Δη = η Tip-modified / η Conventional 1. As the increase of the non-dimensionalized thrust at a certain value of J from model to full scale is larger for the tip-modified propeller, the difference of scale effects is more pronounced when considering η O,CTh. 7 CONCLUSION RANS computations have been made on tip-modified and conventional propellers at model and full scale. While a transition model is applied to the model-scale computations, a fully turbulent flow is modeled in the full-scale computations without the transition model. Grid-independent solutions at model and full scale have been achieved by grid verification studies. The comparison of computations with and with the transition model shows that the effects of laminar and transitional flows are crucial for model-scale propellers. The CFD analysis of scale effects shows that the increase of K T from model to full scale is larger for the tipmodified propeller. The increase of K Q is relatively smaller than that of K T and the K Q increase is larger for the conventional propeller, on the contrary. So the efficiency gain of the tip-modified propeller is increased at full scale. The scale effects are mostly due to nonviscous pressure excluding viscous shear stress. The larger scale effects of the tip-modified propeller in K T and η O are related to alterations of tip vortex and sectional pressure distributions by the bent tip and the higher spanwise loading at the tip region. REFERENCES Andersen, P., Friesch, J., Kappel, J. J., Lundegaard, L. & Patience, G. (2005). Development of a Marine Propeller with Nonplanar Lifting Surfaces. Marine Technology 42, No. 3. Brown, M., Sanchez-Caja, A., Adalid, J.G., Black, S., Sobrino, M. P., Duerr, P., Schroeder, S. & Saisto, I. (2014). Improving Propeller Efficiency Through Tip Loading. Proc. of 30th Symp. on Naval Hydrodynamics, Hobart, Tasmania, Australia. Bugalski, T., Streckwall, H. & Szantyr, J.A. (2013). Critical Review of Propeller Performance Scaling Methods Based on Model Experiments and Numerical Calculations. Polish Maritime Research Vol.20, 4(80), pp.71-79. Funeno, I. (2002). On Viscous Flow around Marine Propellers Hub Vortex and Scale Effect. Journal of Kansai Society of Naval Architects No. 238. Haimov, H., Vicario, J. & Del-Corral, J. (2011). RANSE Code Application for Ducted and Endplate Propellers in Open Water. Proc. of 2 nd Intl. Symp. on Marine Propulsors SMP 11, Hamburg, Germany ITTC. (2008) ITTC Recommended Procedures and Guidelines: Performance, Propulsion 1978 ITTC Performance Prediction Method. 25 th ITTC Krasilnikov, V., Sun, J. & Halse, K. H. (2009). CFD Investigation in Scale Effect on Propellers with Different Magnitude of Skew in Turbulent Flow. Proc. of 1 st Intl. Symp. on Marine Propulsors SMP 09, Tronheim, Norway Sanchez-Caja, A., Gonzalez-Adalid, J., Perez-Sobrino, M. & Sipila, T. (2014). Scale Effects on Tip Loaded Propeller Performance Using a RANSE Solver. Ocean Engineering 88, pp.607-617. Shin, K.W. & Andersen, P. (2016). CFD Study on Effective Wake of Conventional and Tip-modified Propellers. Proc. of 31 st Symp. on Naval Hydrodynamics, Monterey, CA, USA Visonneau, G.B., Deng, G.B., Guilmineau, E., Queutey, P. & Wackers, J. (2016). Local and Global Assessment of the Flow around the Japan Bulk Carrier with and without Energy Saving Devices at Model and Full Scale. Proc. of 31 st Symp. on Naval Hydrodynamics, Monterey, CA, USA

DISCUSSION Questions from Ramon Quereda Thank you very much for your nice presentation. 1) Many authors involved in the open-water extrapolation do not take into account correction in lift when they consider the transitional region at model scale, because they neglect it influence in K T and K Q. What is your opinion in reference with unconventional propellers? 2) Referring to the influence of Rn on the efficiency of the model propeller in open-water tests, I consider that the open-water test should be carried out at Rn = 5 10 5 or higher if possible, whereas ITTC recommendation is 2 10 5. What do you think about it? Author s closure 1) In Figure 13, the increase of K T from model to full scale is larger than that of K Q. It indicates that the scaling method needs to take into account the lift coefficient as well as the drag coefficient. Since the critical difference of unconventional propellers from conventional ones is the spanwise load distribution, a series of chordwise sections along the span need to be considered in the scaling method. The strip method of Streckwall et al (2013) is an example of the scaling method considering several chordwise sections, but it takes into account only the friction drag. 3) Since the open-water propeller flow in the design condition is mostly steady, the difference between the steady and unsteady simulations is insignificant. The computational effort is reduced by adopting the steady simulation. Questions from Papoulias Dimitrios Figure 14 illustrates some strange behavior for the transitional area (x/c 0.8) and differences between the scaled replica model and the real size propeller. Author s closure The steep pressure gradient on the suction side at x/c 0.8 in the model-scale computation indicates the flow separation from a sharp transition from laminar to turbulent flow. As the flow separation is delayed by the more efficient mixing in the fully turbulent boundary layer at full scale, the suction is recovered more gradually than at model scale. REFERENCES Streckwall, H., Greitsch, L., Miller, J., Scharf, M. & Bugalski, T. (2013). Development of a Strip Method Proposed as New Standard for Propeller Performance Scaling. Ship Technology Research Vol.60, pp.58-69 2) Bugalski et al (2013) and Shin & Andersen (2016) show that the open-water efficiency is dependent on Rn at Rn = 10 5-4 10 5 and the dependency is weakened at Rn = 5 10 5-6 10 5. The experiment and computational results imply the necessity for carrying out the open-water test at Rn = 5 10 5 or higher. Questions from Zheng Chaosheng 1) Why is the shape of Zone 3 in Figure 3 chosen to refine the grid? 2) Why is the quarter-cylinder domain with a single blade used in the simulation instead of the full-cylinder domain with all the four blades? 3) What is the difference from the unsteady simulation using the sliding mesh? Author s closure 1) The grid refinement in the conical-shape region downstream from the propeller is for resolving the flow perturbed by the propeller wake. The flow perturbations propagate in a conical shape by the main axial flow with diffusion. 2) The quarter-cylinder domain is adopted for reducing the computational effort. As the sides with the periodic boundary condition are mapped with less than a 0.01% error, it has equivalent accuracy to the computation with the full-cylinder domain and the whole propeller model.