Energy Conversion and Management: X 17 (2023) 100349 Contents lists available at ScienceDirect Energy Conversion and Management: X journal homepage: www.sciencedirect.com/journal/energy-conversion-and-management-x Numerical analysis of propellers for electric boats using computational fluid dynamics modelling Oliver Lovibond a, Anas F.A. Elbarghthi a, b, Vaclav Dvorak b, Chuang Wen a, * a b Faculty of Environment, Science and Economy, University of Exeter, Exeter EX4 4QF, UK Department of Applied Mechanics, Faculty of Mechanical Engineering, Technical University of Liberec, Studentská 1402/2, 46117 Liberec, Czech Republic A R T I C L E I N F O A B S T R A C T Keywords: Wageningen B-series propeller Electric Boat Flow behaviour Multiple Reference Frame Computational Fluid Dynamics In the maritime industry, propellers are the most commonly used form of propulsion and are core to the optimum performance of a ship. Generally, the performance characteristics of a marine propeller are determined and analysed by experiments like open water and self-propulsion scale model tests which are costly and timeconsuming at the initial design stage. In this study, the computational fluid dynamics (CFD) simulations were performed to evaluate propeller performance. Three Wageningen B-series propellers with varying Expanded Area Ratios (EAR) were modelled with respect to the design constraints, such as ship speed and rotational velocity. The performance of the hydrodynamic coefficients, thrust, torque and open water efficiency are then analysed using the CFD modelling. These characteristics are then validated against experimental data obtained from the Netherlands Ship Model Basin open water test in Wageningen and used to investigate the flow behaviour. The analysis considers the Multiple Reference Frame (MRF) model. This study provided a well-founded framework for applying CFD in the analysis and selection of Wageningen B-series propellers, as well as investigated the relationship between the EAR, flow behaviour, thrust coefficient, and torque coefficient for electric boats. The results show that a lower thrust and torque coefficient can improve the flow behaviour with increasing the ef­ ficiency by up to 62%. Furthermore, the outcomes reveal that the lower expanded area ratio of 0.6 is more suitable for electric boats, creating a larger pressure difference of 1.079 MPa and generating extra potential thrust at the same advance ratio, which leads to greater open water efficiency. 1. Introduction The shipping industry is one of the largest modes of transport for global trade. Indeed 90 % of traded goods are carried by ships. An IMO Global Greenhouse Gas (GHG) study estimated that for the period 2007–2012, the shipping industry accounted for 3.1 % of annual global CO2 emissions [1–4]. According to more recent studies by CE Delft, the shipping industry, if left unchecked, could represent 10 % of GHG emissions by 2050 [5]. It can be seen that in the effort to keep global temperature increase below 2 ◦ C, action must be taken in order to reduce global shipping emissions. One way to reduce a vessel’s GHG emissions is to improve the fuel efficiency of the vessel. This can be achieved not only by improving the efficiency of the engine or motor, but also by optimizing the vessel’s propulsion system, namely the propeller. A modern screw propeller with an equivalent arrangement to a pair of tandem propellers on a single shaft was patented in 1838 by James Lowe. The propeller was designed comprising one or more blades where each blade was a portion of a curve which, if continued, would produce a screw [6]. This design was then improved upon by Brunel with respect to contrarotating designs and in 1845 was implemented in the design of Great Britain to much success, which forms the basis of the fixed pitch propellers seen today. The fixed pitch propeller (FPP) is the most commonly used propeller today due to its mechanical simplicity, resil­ ience and effectiveness. It consists of a main hub, connected to a drive shaft, from which a number of blades are attached at a certain pitch angle. The FFP comes in two forms; monoblock propellers which are propellers cast as a single piece and built up propeller, whose blades are cast separately from the boss and then bolted or fixed in some way after machining [6]. In addition, the Wageningen B series propeller is a standard series fixed pitch propeller that was designed and tested at the Netherlands Ship Model Basin in Wageningen. The open water charac­ teristics of 120 propeller models of the B-series were tested at the N.S.M. B and analysed with multiple regression analysis [7]. The hydrodynamic characteristics obtained from this experimental work provides a good foundation from which the calculated results can be validated. * Corresponding author. E-mail address: c.wen@exeter.ac.uk (C. Wen). https://doi.org/10.1016/j.ecmx.2023.100349 Received 9 October 2022; Received in revised form 5 January 2023; Accepted 6 January 2023 Available online 7 January 2023 2590-1745/© 2023 The Author(s). Published by Elsevier Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/). O. Lovibond et al. Energy Conversion and Management: X 17 (2023) 100349 Nomenclature c D KT KQ J n r T Q u VA Abbreviations CFD Computational Fluid Dynamics Exp Experimental EAR Expanded Area Ratio FFP Fixed Pitch Propeller GHG Green House Gas IMO International Maritime Organization MRF Multiple Reference Frame N.S.M.B Netherlands Ship Model Basin No. Number SST Shear Stress Transport RANS Reynolds Average Navier-Stokes Section chord length (m) Diameter of propeller (m) Thrust Coefficient (-) Torque Coefficient (-) Advance Coefficient (-) Rotational speed (rps) Propeller radius (m) Thrust (N) Torque (Nm) Flow velocity (m/s) Advance velocity (m/s) Density of fluid (kg/m3) Expanded area (m2) Propeller disc area (m2) Static pressure (Pa) Number of blades (-) ρ AE A0 p Z Units and Symbols ŋ Open Water Efficiency (-) There are several studies using computational fluid dynamics (CFD) simulations to analyse propeller hydrodynamic characteristics [8–11]. Triet et al. investigated the Wageningen B-Series propeller in open water conditions using the k-epsilon turbulence model [12]. This research, with others such as Tran Ngoc Tu [13] and Subhas et al. [14], utilised the Multi-Reference Frame (MRF) method to model the rotational mo­ tion of the propeller and received satisfactory results. The method consists of separating the model into two regions. A rotating region encapsulates the propeller and a static region covers the rest of the simulation domain. The dimensions of the two domains are comparable to Kutty and Rajendran [15], Subhas et al. [14] and Tran Ngoc Tu [13] and provide a good foundation from which the current model is con­ structed. Tran Ngoc Tu compared different approaches to calculating propeller open water characteristics using the Reynolds Average NavierStokes (RANS) equations method. These approaches are sliding grid, rotating reference frame and rotating domain. The calculations were performed on a hexahedral grid and two-equation of SST k-omega tur­ bulence model was used in the model. The result revealed that rotating reference frame is a suitable method for open water simulation and is now commonly deemed the Multi-Reference Frame (MRF). An under­ standing was gained that the rotating domain must be relatively tight to the propeller, width about 0.38D in order to get accurate results for both the propeller and the wake field. Similarly the static domain should be sufficiently large enough to not interfere with the wake field and also small enough to keep the mesh elements and computational time down. As a result, Triet et al. study provided the knowledge that the k-epsilon model is suitable for design purposes or coarse meshing and facilitates fast simulation time. This is because the model uses wall functions to calculate the near-wall region flows and theoretically requires a coarser mesh at the boundary layer to decrease simulation time. Huge consideration was given to other models. For instance, Guil­ mineau et al. simulated the wake field of a marine propeller using two different RANS equation models; the k-ω SST of Menter and an aniso­ tropic two-equation Explicit Algebraic Reynolds Stress Model (EARSM) [16]. Giancarlo investigated the RANS equations for turbulence modeling [17]. The author concluded that turbulence modeling is an attempt to devise a number of partial differential equations for turbulent-flow calculation based on appropriate approximations of the exact Navier–Stokes equations. This field is relatively challenging to be investigated numerically, which supports visualizing the flow patterns of the modeling. The present study aims to verify the basic knowledge for the flow simulations of a Wageningen B series propeller in electric boats using computational fluid dynamics modelling. The CFD simulation with detailed procedures was conducted to address and predict the impact of the propeller characteristics and map its performance. The study in­ vestigates the relationship between the expanded area ratio of the pro­ peller, the thrust coefficient, the torque coefficient, and efficiency based on the numerical results. 2. Theory and methodology The research aims to model the flow behaviour of a propeller in open water conditions. To achieve this, it is important to consider the con­ servation of mass and momentum in the analysis, as this will allow thrust and torque to be predicted, which is defined by the governing equations. Using this principle, the standard k − ε model will then be applied to model the hydrodynamic characteristics of the propeller as it represents the time-averaged turbulent nature of the flow around the propeller. Once this has been achieved, the coefficient formula will be applied to the hydrodynamic characteristics to produce the hydrody­ namic coefficients and allow comparison and validation against exper­ imental data. Finally, the data will be used to investigate the relationship between flow behaviour and thrust and torque coefficients to ascertain the most appropriate expanded area ratio of the propeller design for electric boats. The flow of a viscous incompressible fluid with constant properties is governed by the Navier–Stokes equations [17]: δui δ δp δ2 ui +v + (ui uj ) = δxi δt δxj δxj δxj (1) δui =0 δxj (2) where ui is the fluid velocity, p is the pressure divided by the density ρ, v is the fluid kinematic viscosity, and body forces do not appear explicitly, where the convective term of equation (1) is expressed in conservative form. The standard k − ε model is a semi-empirical model based on model transport equations for the turbulence kinetic energy (k) and its dissipation rate (ε) [18]. The model transport equation for k is derived from the exact equation. In contrast, the model transport equation for ε was obtained using physical reasoning and bears little resemblance to its mathematically exact counterpart. In the derivation of the k − ε model, the assumption is that the flow is fully turbulent, and the effects of molecular viscosity are negligible [19]. The stand­ ard k − ε model is therefore valid only for fully turbulent flows. The open-water propeller characteristics conventionally are repre­ sented in the form of the thrust and torque coefficients KT and KQ in terms of the advance coefficient J where: 2 O. Lovibond et al. Energy Conversion and Management: X 17 (2023) 100349 Fig. 4. Surface mesh on propeller, first mesh case (Isometric View). Fig. 1. 3D flow domain, including the stationary and rotating domain. Table 1 Mesh options & positioning. Mesh options Position Element size (mm) Face sizing Body sizing Body sizing Propeller blade Rotating domain Stationary domain 8 50 100 Table 2 Global meshing details. Fig. 2. Flow domain dimensions based on the propeller diameter D. Physics preference CFD Solver preference Element order Element size Max adaptive size Growth rate Mesh defeaturing size Curvature min size Curvature normal angle Fluent Linear Default (0.30714 m) Default (0.61428 m) Default (1.2) Default (1.5357 mm) Default (1.5357 mm) Default (18.0◦ ) Table 3 Mesh quality. Physics preference CFD Physics preference CFD Solver preference Element order Element size Fluent Linear Default (0.30714 m) Default (0.61428 m) Solver preference Element order Element size Fluent Linear Default (0.30714 m) Default (0.61428 m) Max adaptive size KT = Fig. 3. Unstructured mesh with tetrahedron elements, first mesh case (Sec­ tion View). KQ = 3 T ρn2 D4 Q ρn2 D5 Max adaptive size (3) (4) O. Lovibond et al. Energy Conversion and Management: X 17 (2023) 100349 Table 4 Mesh Sensitivity Analysis. J= Mesh no. No. of elements No. of elements x105 KT Error from KT (%) 1 2 3 254,100 550,646 1,343,546 2.541 5.50646 13.43546 0.3364614 0.3281122 0.3281318 − 4.024 − 6.671 − 6.664 VA nD (5) where, T is the propeller thrust, Q is the propeller torque, ρ is the fluid density, n is the number of propeller revolutions per second, D is the propeller diameter and VA is the speed of advance. The open water efficiency of the propeller is: η= J KT 2π KQ (6) And the percentage difference in Experimental (Exp) and CFD results is calculated using the following equations: ΔKT (%) = KTCFD − KTExp *100 KTExp (7) ΔKQ (%) = KQCFD − KQExp *100 KQExp (8) Δη(%) = ηCFD − ηExp *100 ηExp (9) The Expanded Area Ratio (EAR) is the most commonly used pro­ peller outline used by designers. It converts the face of a propeller from its helix to a flat plane. The expanded area is given by the relationship: ∫R AE = Z cdr (10) Fig. 5. Mesh convergence graph. rn where Z is the number of blades. To calculate this area, it is sufficient for most purposes to use a Simpson’s rule procedure with 11 ordinates. The expanded blade area ratio is simply the expanded blade area AE , divided by the propeller disc area A0 to give the relationship AE /A0 . This ratio is significant in propeller design and directly affects many propeller characteristics. Lee et al. reported that cavitation might be decreased by increasing the propeller blade area, but efficiency is consequently decreased [20]. This relationship has been emphasized and illustrated the open water efficiency of an MAU standard propeller with respect to an expanded area. Table 5 Cell zone conditions. Rotating Domain Stationary domain Motion Frame motion Relative to cell zone Rotation-axis origin Rotation axis direction Rotational velocity Motion Absolute (X, Y, Z) = (0,0,0) (X, Y, Z) = (0,0,1) Diff. velocity corresponding with n Stationary Table 6 Boundary conditions. Pressure outlet Outer enclosure wall Propeller blade 3. Numerical model descriptions Reference frame Absolute Velocity magnitude Coordinate system Turbulent intensity Backflow reference frame Backflow direction Turbulent intensity Wall motion Wall condition Wall motion Wall condition 6.22 m/s (X, Y, Z) = (0,0,1) 5% Absolute Normal to boundary 5% Stationary No slip Stationary No slip 3.1. Flow domain The numerical predictions presented in this study were performed using ANSYS FLUENT 2021R1 commercial CFD solver. As mentioned previously, the MRF model approach was used to predict the flow around the propeller numerically. The domain is defined and illustrated in Figs. 1 and 2. The domain is split into a global stationary domain and a subdivided rotating region, called the rotating domain. The rotating domain contains the entire propeller specified with dimensions of 1.15D in diameter and 0.4D in length. It should be noted that if the rotating domain is too small, predicted results may be inaccurate due to the effect of large vortices near the propeller. On the other hand, if the domain is too large however it will significantly increase the simulation time. The static domain must be large enough to prevent the full development of the upstream and downstream flow from affecting the results of the analysis. However, if the domain is too large, the computational time increases. For that reason, the inlet is located 2.5D upstream of the origin of the propeller and the outlet is located 5.8D downstream of the propeller to allow the turbulence to collapse freely before hitting a boundary wall. This is all encapsulated by a 3.3D square. Proper selection of the flow domain upstream and downstream distance is very important to prevent recir­ culation of the flow that will cause convergence problems [15]. Table 7 Simulation flow conditions. Case number Rotational speed n (rps) Velocity of advance (U) m/s Advance coefficient (J) 1 2 3 4 5 30 25 20 17.5 15 6.22 6.22 6.22 6.22 6.22 0.346 0.415 0.518 0.592 0.691 4 O. Lovibond et al. Energy Conversion and Management: X 17 (2023) 100349 Table 8 Propeller dimensions. Propeller Diameter (m) No of blades Expanded area ratio P/D ratio Rake Material Elements 1 2 3 0.6 0.6 0.6 4 4 4 0.65 0.6 0.7 1.10627 1.10627 1.10627 15 15 15 Ni-Al Bronze Ni-Al Bronze Ni-Al Bronze 550,646 556,324 563,826 3.3. Mesh sensitivity and model validation Table 9 Solver Settings. Pressure Link SIMPLE Pressure Velocity formulation Gradient Momentum Turbulent kinetic energy Turbulent dissipation rate Turbulence model Near wall treatment Models Solver Standard Absolute Least squares cell based First order upwind First order upwind First order upwind Standard k-epsilon Standard wall functions Single phase Steady The mesh sensitivity analysis was performed at J = 0.415 with the komega model on Propeller 1 and compared to experimental data ob­ tained by Barnitsas et al. (KT = 0.035) [7]. The study was conducted using three grids whose element number increases by the square rule. The summarized results of the mesh sensitivity are illustrated in Table 4. The aim of the mesh convergence study is to determine the mesh density at which the difference of propeller open water characteristic KT obtained from two subsequent meshes reaches a sufficiently low value [13]. The study was conducted by modelling each mesh at J = 0.415 and obtaining the predicted value of thrust (T). The thrust coefficient KT is then calculated using equation (3) which is then compared against the experimental thrust coefficient obtained from the N.S.M.B. [7]. The difference between the CFD value and the Experimental (Exp) value is the error. The percentage error from KT is then calculated and plotted against element number, as shown in Fig. 5. It can be seen from Fig. 5 that as the element number increases, the error in results also increases. This is unusual as generally, the trend should be that of which, with increased element number, the obtained results have increased accuracy. However, the higher element mesh results are in strong agreement with each other with a discrepancy of 0.00598 % despite a difference of 7.9 x105 elements. As a result, the decision was made to use the secondary mesh (element No. 550646) as a basis for the model as it receives greatly the same results as the more detailed mesh number 3 but with a significant reduction in computa­ tional time. The mesh sensitivity results are shown in Table 4. There are multiple reasons for the error in the results. One reason is that the simulation is performed under the single-phase condition. This 3.2. Meshing method The mesh was generated using the mesh tool in the ANSYS program. The grid is fully tetrahedral unstructured in both the stationary and rotating domain, as shown in Fig. 3. The selection is based on justifi­ cations that unstructured tetrahedral grids have the capabilities to dis­ cretize complex geometries fast and with minimum user intervention. The maximum cell size of the propeller and the rotating domain is 8 mm and 50 mm respectively. The propeller surface mesh can be seen in Fig. 4. The mesh was designed so that the cell sizes near the blade wall were small and steadily increased towards the stationary region. The global and local mesh details are displayed in Tables 1 and 2. The mesh quality is shown in Table 3. Table 10 Propeller 1 Hydrodynamic Characteristics. J 0.346 0.415 0.518 0.592 0.691 KT KT (Exp) ΔKT (%) KQ KQ (exp) ΔKQ (%) η η(Exp) Δη(%) 0.3477 0.3281 0.2962 0.2682 0.2270 0.3700 0.3600 0.3300 0.2900 0.2300 − 6.0367 − 8.8577 − 10.2300 − 7.5132 − 1.2965 0.0582 0.0555 0.0507 0.0468 0.0411 0.0750 0.0720 0.0630 0.0580 0.0440 –22.3375 –22.8804 − 19.5104 − 19.2424 − 6.5191 0.3300 0.3900 0.4819 0.5399 0.6071 − 0.5256 0.3900 0.4900 0.5500 0.6200 − 3.4513 − 0.0047 − 1.6446 − 1.8421 − 2.0823 Table 11 Propeller 2 Hydrodynamic Characteristics. J 0.346 0.415 0.518 0.592 0.691 KT KT (Exp) ΔKT (%) KQ KQ (exp) ΔKQ (%) η η (Exp) Δη (%) 0.3482 0.3294 0.2972 0.2710 0.2322 0.3700 0.3600 0.3300 0.2900 0.2400 − 5.8981 − 8.4891 − 9.9441 − 6.5410 − 3.2420 0.0578 0.0554 0.0511 0.0476 0.0423 0.0750 0.0710 0.0630 0.0560 0.0480 –22.8730 –22.0337 − 18.9460 − 15.0734 − 11.8825 0.3310 0.3928 0.4801 0.5373 0.6039 0.3400 0.3900 0.4800 0.5400 0.6200 − 2.6375 0.7081 0.0227 − 0.5018 − 2.5973 Table 12 Propeller 3 Hydrodynamic Characteristics. J 0.346 0.415 0.518 0.592 0.691 KT KT (Exp) ΔKT (%) KQ KQ (exp) ΔKQ (%) η η (Exp) Δη (%) 0.3504 0.3301 0.2956 0.2680 0.2274 0.3700 0.3600 0.3300 0.2900 0.2300 − 5.2923 − 8.3075 − 10.4205 − 7.5739 − 1.1109 0.0596 0.0567 0.0517 0.0478 0.0419 0.0750 0.0710 0.0650 0.0570 0.0480 − 20.5109 − 20.1717 − 20.4331 − 16.2078 − 12.6487 0.3233 0.3844 0.4715 0.5291 0.5967 0.3400 0.3800 0.4700 0.5300 0.6200 − 4.9226 1.1479 0.3248 − 0.1703 − 3.7630 5 O. Lovibond et al. Energy Conversion and Management: X 17 (2023) 100349 Fig. 6. Propeller 1 hydrodynamic characteristics. Fig. 7. Propeller 2 hydrodynamic characteristics. Fig. 8. Propeller 3 hydrodynamic characteristics. means that cavitation is not simulated; however, it would affect the results in the open water test performed at Wageningen. Another reason for the error is the lack of a propeller hub and shaft in the model’s geometry. Both of these solid objects will affect the flow over the pro­ peller and alter the physical hydrodynamic characteristics of the pro­ peller and were present at N.S.M.B. Another reason is that the fluid used 6 O. Lovibond et al. Energy Conversion and Management: X 17 (2023) 100349 Fig. 9. Propeller 2 static pressure contour for J = 0.415: front face left side and back face right side. Fig. 10. Propeller 3 static pressure contour for J = 0.415: front face left side and back face right side. Fig. 11. Pressure contour at J = 0.415: Propeller 3 left side and Propeller 2 right side. in the model is pure liquid H2O whereas the fluid used at Wageningen is impure sea water with a different density which can affect the experi­ mental result. The final reason is that the mesh quality could be improved given other meshing software such as ANSA, which could produce more accurate results [21]. Despite there being an error in re­ sults, the error is sufficiently low enough to validate the model against the experimental data obtained from the ship model basin in Wageningen with only 6.671% error. 3.4. Solver and boundary conditions The domain continuum was chosen as a fluid and the properties of pure liquid water were assigned to it with a density of 998.2 kg/m3 and a viscosity of 0.001003 kg/(m.s). The cell zone and boundary conditions 7 O. Lovibond et al. Energy Conversion and Management: X 17 (2023) 100349 (a) Propeller 1 Velocity Contours at J=0.415 Subhas et al. recommendation on CFD analysis of a Propeller Flow and Cavitation [14]. However, in future studies, the 2nd order upwind setting will be used to get more accurate results at the cost of greater computational time. The detailed setting is shown in Table 9. In the simulation, the numerical algorithm used an implicit method for the pressure -Linked equation known as SIMPLE as the default scheme. The problem was chosen to be a steady state where the mo­ mentum equation was solved to attain the velocity field and the pressure gradient term determined based on the initial condition or previous iteration using the pressure distribution. The solution gradient for the simulation is calculated based on the Least-Squares Cell-Based approach because it delivers the same range of accuracy with less intensive computation. Since the iterative domain contains a huge number of cells that require faster convergence, the First-Order Upwind with first-order accuracy. The Standard k-epsilon turbulent model was selected using the transport equation for the kinetic energy and the dissipation rate and fits the current simulation study because it was assumed for fully turbulent flows. In the calculations, the mesh should predict the velocity gradient within the boundary layer. This essentially requires capturing the thin viscous layer by the first cell adapted at the wall. However, if the first cell is placed at the log-layer region, then the standard wall function could be chosen, which can predict various ranges of Reynolds-number covering y + higher than 30 [24]. In general, both standard and real­ izable k-epsilon model variants used wall functions. The only matter is that the value of y + near the wall must not be below 30 as has been checked in this simulation.”. (b) Propeller 2 Velocity Contours at J=0.415 4. Results and discussion 4.1. Hydrodynamic characteristics The simulation was run for each propeller at 5 different advanced coefficients in order to find the hydrodynamic characteristics, KT , KQ and η. The results of these characteristics are then compared to experi­ mental data (Exp) and the error in the results is calculated. These results are displayed in Tables 10, 11 and 12. The advance coefficient J is changed by varying the revolution of the propeller while the advance velocity was kept constant at 6.22 m/s. The hydrodynamic data for all three propellers is being modelled. It can be seen from Figs. 6 -8 that the overall accuracy of results increases with an increasing advance ratio (J). The CFD efficiency almost exactly replicates the Experimental data with the greatest dif­ ference being − 4.92 % with Propeller 3 at J = 0.346. For the torque coefficient, accuracy increases and converges at the high end of J. Whereas with the thrust coefficient, the greatest inaccuracy is found in the middle region of J, most notably at J = 0.518. It is noticeable in Fig. 8 that propeller 3 experienced the lowest ef­ ficiency. This is in correlation with Chang Sup Lee’s prediction that increasing the expanded area ratio will cause the efficiency to decrease [14]. It can be seen from the Tables 10-12 that Propeller 3 consistently has the greatest KT for all values of J. This suggests that the greatest EAR of 0.7 can produce the largest thrust. Propeller 3 also has the largest KQ indicating it delivers the most torque and subsequently power out of all propellers. This does not mean it has the highest efficiency, however. The greatest efficiency is reached by Propeller 3 at the low end of J. Nevertheless, towards the higher end of J, Propeller 3′ s efficiency drops off and both Propeller 1 and 2 experience greater efficiency. This sug­ gests that a lower EAR produces an improved efficiency at high advance ratios. Yet a higher EAR produces an improved efficiency at low advance ratios. Interestingly, the highest accuracy results of KT and KQ which appear at J = 0.691, correlate with the lowest accuracy efficiency seen in Figs. 6-8. (c) Propeller 3 Velocity Contours at J=0.415 Fig. 12. Propeller velocity contours. are displayed in Tables 5 and 6, respectively. CFD simulation was conducted over a range of propeller advance ratios corresponding to the rotational velocity of the rotating domain, as summarized in Table 7. 3.5. Propeller geometry and Solver setting The geometry of the Wageningen B-series can be obtained from the official Wageningen B-Series Propeller generator, which allows you to create a propeller with B-series dimensions identical to those tested at Wageningen and the expanded area ratio as a variable. In this paper, three propellers that have identical dimensions but have three different expanded area ratios: 0.6, 0.65 and 0.7 will be investigated. The pro­ peller geometry can be seen in Table 8 [25]. As mentioned previously, a MRF method is used to model the pro­ peller rotation. A number of hypotheses are used to simplify the case of simulating propeller hydrodynamic characteristics. These include steady-state flow, non-cavitation and incompressible flow, hence the single phase model. The solver settings used in the k − ε model are dis­ played in Table 9. The first order upwind settings were used following 8 O. Lovibond et al. Energy Conversion and Management: X 17 (2023) 100349 Fig. 13. Isometric view for velocity vectors at J = 0.415 for Propeller 2. 4.2. Pressure and velocity distributions particles are accelerated to a higher velocity. Therefore, a greater force must be applied to the fluid, henceforth a greater reaction force is applied to the propeller culminating in improved thrust. The minimum velocity with respect to the boundary conditions in all three figures is 0 m/s. The isosurface velocity as three-dimensional analog of Propeller 2 is displayed in Fig. 13. It used to link up all the point with same values within a volume of space. The result showed that the fluid reaches its maximum velocity at the propeller tips. It can also be noted that Pro­ peller 2 has a higher maximum velocity of 46.95 m/s. This suggests once again that the lower EAR can produce a greater thrust, albeit by an almost negligible improvement of 0.32 %. In general, the propeller with the lowest EAR, as well as the lower thrust and torque coefficients, generates the highest slipstream velocity, improves the flow behaviour, and suggests suitability for optimisation for electric motors, which can generate maximum torque instantly. Fig. 9 shows the pressure distribution across the front and back face of Propeller 2 operating at an advance ratio of J = 0.415. It can be seen that the pressure at the back side of the propeller (the pressure face), is slightly greater than the pressure at the front side of the propeller, (the suction face). Indeed a maximum pressure of 435.9 kPa is reached on the pressure face compared to a maximum pressure of 121.3 kPa on the suction face. This difference in pressure between the front and back faces of the propeller is the creating force in the forward direction known as thrust. The greatest difference in pressure between the two faces is sit­ uated at the leading edge of the propeller and reaches a total pressure difference of 1.079 MPa. When compared to Propeller 3 displayed in Fig. 10 at J = 0.415, the maximum pressure experienced is 385.0 kPa, which is less than Propeller 2 with the lower EAR and the maximum pressure difference obtained is 1.042 MPa. This shows that a smaller EAR creates a slightly larger pressure difference and has the potential to create more thrust at the same advance ratio. The pressure difference between the front and back faces of the propeller can be clearly seen in Fig. 11. The maximum pressure differ­ ence appears between R/R0 = 0.8 and the tip of the blades. Here the fluid pressure difference reaches a maximum of 586.4 kPa for Propeller 3 and 594.4 kPa for Propeller 2. This is where the thrust is generated, and the results indicate that greater thrust can be produced with the smaller EAR consistent with Propeller 2. It is important to note the area of relatively low pressure situated aft of the propeller hub. This area is likely to cavitate due to the local pressure dropping below the vapour pressure of the fluid, and this effect is enhanced due to the vortical nature of the propellers slipstream and the positive feedback loop associated with cavitating vortices [22]. The velocity contours of each propeller are displayed in Fig. 12. All propellers are used as a comparison in the velocity section. It can be seen that at the same Advance Coefficient J = 0.415, Propeller 3 (Fig. 12.c) induces the lowest maximum velocity of 45.39 m/s and that Propeller 1 and 2 which have a lower EAR induce a greater maximum velocity by 0.39 % and 0.31 % respectively. This means that a lower EAR has the potential to create more thrust due to Newton’s third law [23]. The 5. Conclusion This research was devoted to a CFD simulation to evaluate three Wageningen B-series propellers’ performance. The study successfully provides the framework needed to numerically analyse a propeller’s open-water characteristics. The commercial CFD code FLUENT was found to be reliable in providing good initial predictions and thus is a viable alternative to open water tests at the design phase. The validation revealed that the CFD simulations used in this study are in good agreement with the experimental results recording an average error of 8.65 %. This diversion in the accuracy of the experiment can be explained by the effect of some minor factors, such as using pure water and the absence of the hub or the shaft in the validation. Moreover, the results for the thrust coefficient showed a slight underprediction for all advance ratios, with this underprediction max­ imising at J = 0.518 and minimising and converging for high advance ratios, namely J = 0.691. The results for the torque coefficient showed a much greater underprediction with a decrease in error towards the higher advance ratios. The open water efficiency was very accurate throughout the advance ratio, with a slight underprediction at the 9 O. Lovibond et al. Energy Conversion and Management: X 17 (2023) 100349 extremities of the advance ratio. Overall results showed a reliable capability to predict the performance of a Wageningen B-Series propeller. However, based on the pressure and fluid velocity analysis, the lower expanded area ratio showed the greatest potential in creating the most thrust. Concerning flow behaviour, the greatest thrust and torque co­ efficients correlate with the lowest propeller slipstream velocity, as seen in Propeller 3. This suggests that lower thrust and torque coefficients, which correspond with a lower expanded area ratio, as shown in Pro­ pellers 1 and 2, can generate an improved flow behaviour with a higher particle velocity. It could be noticed that up to 62 % of the efficiency increase could be gained with lower thrust and torque coefficients. In addition, the outcome revealed that the lower expanded area ratio of 0.6 is more suitable for electric boats, creating a larger pressure difference of 1.079 MPa and generating extra potential thrust at the same advance ratio, which leads to greater open water efficiency. [4] Selin H, Zhang Y, Dunn R, Selin NE, Lau AKH. Mitigation of CO2emissions from international shipping through national allocation. Environ Res Lett 2021;16(4): pp. https://doi.org/10.1088/1748-9326/abec02. [5] jf, “CE_Delft_7I69_Update_of_Maritime_Greenhouse_Gas_Emission_Projections_Def_ v1,” 2017, [Online]. Available: www.cedelft.eu. [6] Carlton J. Marine propellers and propulsion. Butterworth-Heinemann; 2018. [7] Barnitsas MM, Ray D, Kinley P. KT, KQ and efficiency curves for the Wageningen Bseries propellers. University of Michigan; 1981. [8] E. Javanmard, E. Yari, J. A. Mehr, and S. Mansoorzadeh, “Hydrodynamic characteristic curves and behavior of flow around a surface-piercing propeller using computational fluid dynamics based on FVM,” Ocean Eng., vol. 192, no. November 2018, p. 106445, 2019, 10.1016/j.oceaneng.2019.106445. [9] Song S, Demirel YK, Atlar M. Propeller performance penalty of biofouling: Computational fluid dynamics prediction. J Offshore Mech Arct Eng 2020;142(6): pp. https://doi.org/10.1115/1.4047201. [10] Helal MM, Ahmed TM, Banawan AA, Kotb MA. Numerical prediction of the performance of marine propellers using computational fluid dynamics simulation with transition-sensitive turbulence model. Proc Inst Mech Eng Part M J Eng Marit Environ 2019;233(2):515–27. https://doi.org/10.1177/1475090218763199. [11] L. do V. Machado and A. C. Fernandes, “Proposals to improve load and cavitation computational fluid dynamics analysis with moving reference frame of a conventional propeller,” J. Offshore Mech. Arct. Eng., vol. 143, no. 2, 2021, 10.1115/1.4048542. [12] Pham MT, Phan QT, Ngo KH. CFD simulation for the Wageningen B-Series propeller characteristics in open-water condition using k-epsilon turbulence model SEE PROFILE. Sci Technol Dev J - Eng Technol 2019;1(1):35–42. [13] Tu TN. Numerical simulation of propeller open water characteristics using RANSE method. Alexandria Eng J 2019;58(2):531–7. https://doi.org/10.1016/j. aej.2019.05.005. [14] Subhas S, Saji VF, Ramakrishna S, Das HN. CFD analysis of a propeller flow and cavitation. Int J Comput Appl Technol 2012;55(16):26–33. https://doi.org/ 10.5120/8841-3125. [15] Kutty HA, Rajendran P. 3D CFD simulation and experimental validation of small APC slow flyer propeller blade. Aerospace 2017;4(1):pp. https://doi.org/10.3390/ aerospace4010010. [16] E. Guilmineau et al., “Wake simulation of a marine propeller To cite this version : HAL Id : hal-01202477,” 2020. [17] Giancarlo A. Reynolds-Averaged Navier-Stokes equations for turbulence modeling. Appl Mech Rev 2009;62(4):40802. [18] B. E. Launder and D. B. Spalding, “Lectures in mathematical models of turbulence,” 1972. [19] Wilcox DC. Turbulence modeling for CFD, vol. 2. CA: DCW industries La Canada; 1998. [20] C. S. Lee, Y. D. Choi, B. K. Ahn, H. G. Jang, and M. S. Shin, “Performance optimization of marine propellers,” 11th Int. Symp. Pract. Des. Ships Other Float. Struct. PRADS 2010, vol. 1, no. 4, pp. 674–680, 2010, 10.2478/ijnaoe-2013-0038. [21] ANSA, “No Title,” [Online]. Available: https://www.ansa-usa.com/software/ansa/ . [22] Reisman GE, Wang Y-C, Brennen CE. Observations of shock waves in cloud cavitation. J Fluid Mech 1998;355:255–83. [23] T. E. of E. Britannica, “Newton’s laws of motion,” Encycl. Br., [Online]. Available: https://www.britannica.com/science/Newtons-laws-of-motion. [24] Ansys. ANSYS FLUENT 12.0 Theory Guide. (April); 2009. [25] https://www.wageningen-b-series-propeller.com/. Declaration of Competing Interest The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper. Data availability Data will be made available on request. Acknowledgment The second author, Anas F.A.Elbarghthi would like to acknowledge the support of the Student Grant Competition of the Technical Univer­ sity of Liberec under project No. SGS-2023-5323. References [1] “IMO, Third IMO GHG Study, International Maritime Organization, 2014.” https:// www.imo.org/en/OurWork/Environment/Pages/Greenhouse-Gas-Studies-2014. aspx. [2] Xing H, Spence S, Chen H. A comprehensive review on countermeasures for CO2 emissions from ships. Renew Sustain Energy Rev 2020;vol. 134, no. August: 110222. https://doi.org/10.1016/j.rser.2020.110222. [3] Wang X-T, Liu H, Lv Z-F, Deng F-Y, Xu H-L, Qi L-J, et al. Trade-linked shipping CO2 emissions. Nat Clim Chang 2021;11(11):945–51. 10