National Academies Press: OpenBook

Twenty-Third Symposium on Naval Hydrodynamics (2001)

Chapter: Numerical Prediction of Scale Effects in Ship Stern Flows with Eddy-Viscosity Turbulence Models

« Previous: On Submerged Stagnation Points and Bow Vortices Generation
Suggested Citation:"Numerical Prediction of Scale Effects in Ship Stern Flows with Eddy-Viscosity Turbulence Models." National Research Council. 2001. Twenty-Third Symposium on Naval Hydrodynamics. Washington, DC: The National Academies Press. doi: 10.17226/10189.
×
Page 553
Suggested Citation:"Numerical Prediction of Scale Effects in Ship Stern Flows with Eddy-Viscosity Turbulence Models." National Research Council. 2001. Twenty-Third Symposium on Naval Hydrodynamics. Washington, DC: The National Academies Press. doi: 10.17226/10189.
×
Page 554
Suggested Citation:"Numerical Prediction of Scale Effects in Ship Stern Flows with Eddy-Viscosity Turbulence Models." National Research Council. 2001. Twenty-Third Symposium on Naval Hydrodynamics. Washington, DC: The National Academies Press. doi: 10.17226/10189.
×
Page 555
Suggested Citation:"Numerical Prediction of Scale Effects in Ship Stern Flows with Eddy-Viscosity Turbulence Models." National Research Council. 2001. Twenty-Third Symposium on Naval Hydrodynamics. Washington, DC: The National Academies Press. doi: 10.17226/10189.
×
Page 556
Suggested Citation:"Numerical Prediction of Scale Effects in Ship Stern Flows with Eddy-Viscosity Turbulence Models." National Research Council. 2001. Twenty-Third Symposium on Naval Hydrodynamics. Washington, DC: The National Academies Press. doi: 10.17226/10189.
×
Page 557
Suggested Citation:"Numerical Prediction of Scale Effects in Ship Stern Flows with Eddy-Viscosity Turbulence Models." National Research Council. 2001. Twenty-Third Symposium on Naval Hydrodynamics. Washington, DC: The National Academies Press. doi: 10.17226/10189.
×
Page 558
Suggested Citation:"Numerical Prediction of Scale Effects in Ship Stern Flows with Eddy-Viscosity Turbulence Models." National Research Council. 2001. Twenty-Third Symposium on Naval Hydrodynamics. Washington, DC: The National Academies Press. doi: 10.17226/10189.
×
Page 559
Suggested Citation:"Numerical Prediction of Scale Effects in Ship Stern Flows with Eddy-Viscosity Turbulence Models." National Research Council. 2001. Twenty-Third Symposium on Naval Hydrodynamics. Washington, DC: The National Academies Press. doi: 10.17226/10189.
×
Page 560
Suggested Citation:"Numerical Prediction of Scale Effects in Ship Stern Flows with Eddy-Viscosity Turbulence Models." National Research Council. 2001. Twenty-Third Symposium on Naval Hydrodynamics. Washington, DC: The National Academies Press. doi: 10.17226/10189.
×
Page 561
Suggested Citation:"Numerical Prediction of Scale Effects in Ship Stern Flows with Eddy-Viscosity Turbulence Models." National Research Council. 2001. Twenty-Third Symposium on Naval Hydrodynamics. Washington, DC: The National Academies Press. doi: 10.17226/10189.
×
Page 562
Suggested Citation:"Numerical Prediction of Scale Effects in Ship Stern Flows with Eddy-Viscosity Turbulence Models." National Research Council. 2001. Twenty-Third Symposium on Naval Hydrodynamics. Washington, DC: The National Academies Press. doi: 10.17226/10189.
×
Page 563
Suggested Citation:"Numerical Prediction of Scale Effects in Ship Stern Flows with Eddy-Viscosity Turbulence Models." National Research Council. 2001. Twenty-Third Symposium on Naval Hydrodynamics. Washington, DC: The National Academies Press. doi: 10.17226/10189.
×
Page 564
Suggested Citation:"Numerical Prediction of Scale Effects in Ship Stern Flows with Eddy-Viscosity Turbulence Models." National Research Council. 2001. Twenty-Third Symposium on Naval Hydrodynamics. Washington, DC: The National Academies Press. doi: 10.17226/10189.
×
Page 565
Suggested Citation:"Numerical Prediction of Scale Effects in Ship Stern Flows with Eddy-Viscosity Turbulence Models." National Research Council. 2001. Twenty-Third Symposium on Naval Hydrodynamics. Washington, DC: The National Academies Press. doi: 10.17226/10189.
×
Page 566
Suggested Citation:"Numerical Prediction of Scale Effects in Ship Stern Flows with Eddy-Viscosity Turbulence Models." National Research Council. 2001. Twenty-Third Symposium on Naval Hydrodynamics. Washington, DC: The National Academies Press. doi: 10.17226/10189.
×
Page 567
Suggested Citation:"Numerical Prediction of Scale Effects in Ship Stern Flows with Eddy-Viscosity Turbulence Models." National Research Council. 2001. Twenty-Third Symposium on Naval Hydrodynamics. Washington, DC: The National Academies Press. doi: 10.17226/10189.
×
Page 568

Below is the uncorrected machine-read text of this chapter, intended to provide our own search engines and external engines with highly rich, chapter-representative searchable text of each book. Because it is UNCORRECTED material, please consider the following text as a useful but insufficient proxy for the authoritative book pages.

lengths, word breaks, heading styles, and other typesetting-specific formatting, however, cannot be retained, and some typographic errors may have been accidentally inserted. Please use the print version of this publication as About this PDF file: This new digital representation of the original work has been recomposed from XML files created from the original paper book, not from the original typesetting files. Page breaks are true to the original; line NUMERICAL PREDICTION OF SCALE EFFECTS IN SHIP STERN FLOWS WITH EDDY-VISCOSITY TURBULENCE MODELS 553 Numerical Prediction of Scale Effects in Ship Stern Flows with Eddy-Viscosity Turbulence Models L.Eça (Instituto Superior Técnico, Portugal) M.Hoekstra (Maritime Research Institute, Netherlands) ABSTRACT This paper presents a numerical investigation of scale effects on ship stern flows. The most popular algebraic, one- equation and two-equation eddy-viscosity turbulence models are successfully applied to the calculation of the flow around the Mystery tanker from model up to full scale Reynolds number. It is shown that the choice of the turbulence model does have an influence on the results (notably the near-wake field), but the differences caused by changing the turbulence model tend to diminish with the increase of the Reynolds number. 1 INTRODUCTION One of the main advantages of Computational Fluid Dynamics (CFD) over traditional model testing is the (potential) capability to predict Reynolds number effects on the flow field around a ship. But before this advantage can be exploited, two major tasks arise. The first is to change the potential capability to a true capability by making sure that the numerical method can cope with the extreme requirements posed by flow simulations at full scale Reynolds number. The second task is to go through the verification and validation processes. Only a few attempts to predict scale effects with CFD have been reported. One such attempt is [1], presenting results, however, which are not numerically convincing. The present authors have shown to be more successful in computing ship stern flows from model up to full scale Reynolds numbers, [2], with verification of numerical errors, [3] and [4]. Unfortunately, there are virtually no reliable experimental data available for full scale ship stern flows, so that the validation process is obstructed. The best thing to do is then to increase the level of confidence by validating at model- scale Reynolds number and showing that well-known trends for Rn increasing are systematically reproduced. In the last two decades, a huge effort has been made to validate CFD predictions at model scale Reynolds number. However, the present status is far from being completely satisfactory. Notably the accurate prediction of the axial velocity field in regions of high streamwise vorticity has proved to be difficult. Some success has been claimed for second moment closures, e.g. [5]. But if one aims at selecting a turbulence closure that is numerically robust at model and full scale Reynolds numbers, eddy-viscosity models are still the only reasonable choice. In this paper we present a numerical investigation into the prediction of scale effects with eddy-viscosity turbulence models, including algebraic, one-equation and two-equation models. Two main goals are considered: • Investigate which turbulence models are numerically robust from model up to full scale Reynolds numbers, without the need of further tuning. • Evaluate the influence of the Reynolds number on the differences between solutions obtained with different eddy- viscosity turbulence models. With achieving these goals, we expect to increase the confidence in the use of CFD at full scale Reynolds numbers, using turbulence models that have been originally developed for thin shear layers at model scale Reynolds numbers. The paper is organised in the following way: section 2 gives a brief description of the turbulence models and their numerical implementation. The results of application to the flow around the Mystery tanker at Reynolds numbers from model scale up to full scale Rn are presented and discussed in section 3. Section 4 summarizes the conclusions of the paper. the authoritative version for attribution.

lengths, word breaks, heading styles, and other typesetting-specific formatting, however, cannot be retained, and some typographic errors may have been accidentally inserted. Please use the print version of this publication as About this PDF file: This new digital representation of the original work has been recomposed from XML files created from the original paper book, not from the original typesetting files. Page breaks are true to the original; line NUMERICAL PREDICTION OF SCALE EFFECTS IN SHIP STERN FLOWS WITH EDDY-VISCOSITY TURBULENCE MODELS 554 2 TURBULENCE MODELS In order to cover a broad spectrum of eddy-viscosity turbulence models, we have considered the following models: • Algebraic models – Cebeci & Smith, [6], (CS) – Baldwin & Lomax [7], (BL) • One-equation models – Spalart & Allmaras, [8], (SA) – Menter, [9], (MT) • Two-equation models – k-ε models Two-layer model, [10], (KE-TL) Chien's low-Rn model, [11], (KE) – k-ω models Standard, [12], (KW) Menter, [13], (KWM) – q-ζ model, [14], (QZ) The one-equation model of Baldwin & Barth, [15], and the SST version of Menter's k-ω model, [13], were also tested. However, results of these models will not be included in the present paper. The Baldwin & Barth model showed a very poor behaviour in the initial test runs, while the SST version of Menter's k-ω does not perform better than the other k- ω models tested. It is possible to improve the quality of the predictions of ship stern flows with the one-equation and two-equation turbulence models using a simple correction to the production term of the transport equations, [16]. However, in this paper we will adopt the standard versions of the models. 2.1 Algebraic Models The two algebraic models are well-known and based on a two-layer definition of the eddy-viscosity, vt, where the eddy-viscosity is obtained from the minimum of its values in the two layers. In the inner-layer, both models use the mixing-length approach with the Van Driest damping function in the near-wall region. In the Cebeci & Smith model, the eddy-viscosity in the outer region is obtained from (1) where qe stands for the velocity at the edge of the viscous region and δ* is the displacement thickness, which is an integral parameter defined for 2-D boundary-layer flows, and γk is the intermittency factor, which is given by (2) where δ is the thickness of the viscous region. In a ship stern flow calculation, performed in a curvilinear coordinate system, (ξ, η, ζ)1 some assumptions have to be made to compute qe, δ* and γk. We determine these quantities from information along each grid line normal to the wall and the viscous layer thickness is calculated from the total head. Details on the calculation of qe, δ* and γk are given in [16]. The main advantage of the Baldwin & Lomax model over the Cebeci & Smith model is the absence of δ and δ* from the definition of the length scale in the outer region. In the Baldwin & Lomax model, (vt)0 is given by (3) where (4) Fmax is the maximum of the function (5) and ymax is the value of yn where Fmax occurs. Udif is the difference between the maximum and minimum values of q along an η grid line, |W| is the magnitude of the vorticity vector, A−=26 and is the nondimensional distance to the wall in wall coordinates. Fkleb is the equivalent of γk, and is given by the authoritative version for attribution. (6) Ccp=1.6, C2=0.25 and Ckleb=0.3. Although the calculation of Fwake and ymax seems to be straightforward, it is not. The values of these two parameters are directly related to the vorticity, which is inversely proportional to the grid line distance in the physical space. This dependency of F on the vorticity makes its numerical calculation extremely sensitive to 1ξ is a stream wise coordinate, η is a coordinate normal to the ship surface and ζ is a girthwise coordinate.

lengths, word breaks, heading styles, and other typesetting-specific formatting, however, cannot be retained, and some typographic errors may have been accidentally inserted. Please use the print version of this publication as About this PDF file: This new digital representation of the original work has been recomposed from XML files created from the original paper book, not from the original typesetting files. Page breaks are true to the original; line NUMERICAL PREDICTION OF SCALE EFFECTS IN SHIP STERN FLOWS WITH EDDY-VISCOSITY TURBULENCE MODELS 555 the mean flow field, which depends on vt and is determined iteratively. To avoid undershoots and overshoots of the eddy- viscosity, the spatial variation of ymax must be limited. In the present implementation, the following limiters are adopted is the ymax value of a given ξ line at stream wise stationi and where stands for the ymax value on the previous upstream station for the same ξ line. It is our experience that it is very hard to converge the eddy-viscosity field when the limiters of ymax are turned off. On the other hand, if the limiters are always turned on there is a risk of ymax being determined by the upstream flow instead of the local flow quantities. Therefore, the apparent advantages of the Baldwin & Lomax model in the determination of (vt)0 are severely compromised by the numerical difficulties found in the determination of ymax. Details of the adjustment of the two models to make them suited for application to wakes are given in [16]. 2.2 One-Equation models 2.2.1 Spalart & Allmaras The new generation of one-equation turbulence models is based on a transport equation for the eddy viscosity rather than for the turbulence kinetic energy. The Spalart & Allmaras model proposed in [8] solves the following transport equation: (7) where (9) (8) Here and in the remainder of this paper S represents the rate of strain squared. The eddy-viscosity is obtained from The model constants are: The transport equation of (7), includes the term which does not contribute to the stability of its numerical solution. Therefore, equation (7) has been re-written as (10) In the wake, the distance to the wall is computed from (11) where yn is the distance measured along the normal grid lines and x−xte is the distance to the ‘trailing edge' of the ship measured along the symmetry plane. 2.2.2 Menter The one equation model proposed by Menter in [9] derives the following transport equation from the k-ε model: (12) The eddy-viscosity is given by (13) and the authoritative version for attribution. (14) The model constants are:

lengths, word breaks, heading styles, and other typesetting-specific formatting, however, cannot be retained, and some typographic errors may have been accidentally inserted. Please use the print version of this publication as About this PDF file: This new digital representation of the original work has been recomposed from XML files created from the original paper book, not from the original typesetting files. Page breaks are true to the original; line NUMERICAL PREDICTION OF SCALE EFFECTS IN SHIP STERN FLOWS WITH EDDY-VISCOSITY TURBULENCE MODELS 556 2.2.3 Boundary Conditions In a ship stern flow calculation we need to specify boundary conditions at the six boundaries of the domain: inlet, outlet, ship surface, ship symmetry plane, free surface2 and external boundary. The boundary conditions are specified in the same way for both one-equation turbulence models. At the ship surface the turbulent quantities are zero and symmetry conditions are applied on the free surface and at the ship's symmetry plane. At the outlet, the streamwise derivative of the turbulent quantities is assumed to be zero. Dirichlet boundary conditions are imposed for the turbulent quantities on the external boundary and are fixed by the value on the border of the inlet plane. The present calculations were performed only for the aft part of the ship; the inlet plane is located at mid-ship, and so there is already a viscous region in the inlet plane. A standard straightforward procedure was adopted to define the turbulent quantities at the inlet plane: vt is calculated with the Cebeci & Smith algebraic model and the turbulent quantities are then derived from the known eddy- viscosity. 2.3 Two-equation Models 2.3.1 Two-layer k- ε The two-layer k-ε model presented by Chen and Patel in [10] solves two transport equations in the outer flow region. In the near-wall region only the equation for the turbulence kinetic energy, k, is solved. The value of ε is derived from an algebraic length scale. The near-wall model is equivalent to the one-equation model of Wolfshtein, [17]. The eddy-viscosity is obtained from (15) and k and ε for a steady flow are obtained from the solution of the equations: (16) and (17) In the near-wall region, ε is determined from (18) with (19) and (20) The fµ function defined by the Wolfshtein one-equation model is (21) The standard k-ε constants are cµ=0.09, C1=1.44, C2=1.92, σk=1 and σε=1.3. The key feature of this two-layer model is the determination of the boundary between the inner and outer layers which is often defined by a criterion based on y−. However, with y− it is difficult to establish a criterion which is insensitive to the Reynolds number. In our approach, the inner-layer region is defined by the following criteria: The first criterion would be the natural choice to border the inner-layer region. However, in the iterative determination of the eddy-viscosity field it may lead to excessively large regions, which provoke numerical convergence problems. Therefore, we have added the second criterion which originates from the knowledge on flat plate boundary layers, where the ‘fully-turbulent' region starts at This approach does not guarantee that the fµ is close to 1 at the edge of inner-layer. Therefore, in the outer-layer the ε transport equation is solved but fµ is still obtained from (21). As in the one-equation models, (yn)*, defined by equation (11), is used in the wake to represent the distance to the wall. 2.3.2 Chien's k- ε model The low Reynolds k-ε model proposed by Chien, [11], does not distinguish between inner and outer layers and is directly applicable in the near-wall region. The eddy-viscosity is obtained from equation (15). The k and ε transport the authoritative version for attribution. equations of this model are: (22) 2In the present calculation we have used the double-model approximations and so the free surface is asymmetry plane.

lengths, word breaks, heading styles, and other typesetting-specific formatting, however, cannot be retained, and some typographic errors may have been accidentally inserted. Please use the print version of this publication as About this PDF file: This new digital representation of the original work has been recomposed from XML files created from the original paper book, not from the original typesetting files. Page breaks are true to the original; line NUMERICAL PREDICTION OF SCALE EFFECTS IN SHIP STERN FLOWS WITH EDDY-VISCOSITY TURBULENCE MODELS 557 and (24) (23) The near-wall damping functions are given by: (25) The model constants are 2.3.3 Standard k- ω model The k-ω model has been proposed by Wilcox, [12]. It obtains the eddy-viscosity from (26) and k and ω are obtained from the solution of the equations: (27) and (28) The ω equation can be integrated down to the wall and the model constants are: 2.3.4 Menter's k- ω model Menter's version of the k-ω model, [13], is a blending between the k-ε and the k-ω models. The objective of the model is to solve the ω equation in the near-wall region, which does not require extra damping functions, whereas in the outer region of the flow the ε transport equation is solved. As in the standard k-ω model, the eddy-viscosity is computed from (26) and the k transport equation is given by (27). The ω transport equation is re-written as: (29) where the constants of the model, for convenience symbolically denoted by are obtained from The set of constants is the one of the standard k-ω model and the set of constants has been derived from the k-ε model and is given by: The blending function, F1 is given by (30) with (32) (31) and 2.3.5 q- ζ model The q-ζ model is proposed in [14]. It is a two-equation model derived from the k-ε model with the objective of having turbulent quantities which go to zero at the wall. The two turbulent quantities of the model, q and ζ are related to k the authoritative version for attribution. and ε by (33) The transport equations of q and ζ are derived from the transport equations of k and ε with the relations: (35) (34) In the present implementation of the method, Chien's low Reynolds version of the k-ε model was adopted to obtain the transport equations of q and ζ. The eddy-viscosity is given by

lengths, word breaks, heading styles, and other typesetting-specific formatting, however, cannot be retained, and some typographic errors may have been accidentally inserted. Please use the print version of this publication as About this PDF file: This new digital representation of the original work has been recomposed from XML files created from the original paper book, not from the original typesetting files. Page breaks are true to the original; line NUMERICAL PREDICTION OF SCALE EFFECTS IN SHIP STERN FLOWS WITH EDDY-VISCOSITY TURBULENCE MODELS 558 and (36) (37) The near-wall damping functions are given by: (38) (39) (40) The model constants are 2.3.6 Boundary Conditions As in the one-equation models, boundary conditions are required for the transport equation of the turbulent quantities at the six boundaries of the computational domain of a ship stern flow calculation. At the symmetry plane of the ship and at the free surface, symmetry boundary conditions are applied to all turbulent quantities. Zero streamwise derivatives are applied at the outlet boundary. At the external boundary, we have applied Dirichlet boundary conditions. The values of the different turbulent quantities were set to obtain vt=0.01v and assuming that ω∞=U∞/L3. At the inlet plane, the turbulent quantities are fixed with a similar procedure as adopted for the one-equation models. However, in the two-equation models we need extra information because we have two turbulent quantities to define. We determine the k (or q) profiles from (41) with (42) where uτ is the friction velocity. The second turbulent quantity is obtained from k (or q) and the values of vt obtained from the Cebeci & Smith algebraic model. In the outer region, yn≥0.15δ, an Hermite cubic interpolation is used to obtain the turbulent quantities, assuming zero derivatives at the external boundary. At the ship surface, k, q and ζ are zero. In Chien's formulation of the k-ε model, ε is also zero at the wall. However, the wall boundary condition for ω asks for a few more words. According to [12], ω behaves in the vicinity of the wall as: (43) where Although almost everyone reports that the near-wall behaviour is a strong point of the model, the value of ω at the wall actually tends to infinity! Only in [12] it is recognised that the ω boundary condition at the wall may lead to numerical difficulties when fine grids are used in the near-wall region. The most popular implementation of the ω boundary condition at the wall is suggested by Menter in [13], which just fixes the wall value by (44) where ∆y1 is the distance of the first grid node to the wall. Obviously, this condition is just equation (43) multiplied by 10, and can be criticized for the following reasons: the authoritative version for attribution. 1. The factor 10 is completely arbitrary4. 2. It is based on the ω solution without viscous corrections in a region where viscous effects are dominant. 3. It is clearly grid dependent, because the distance of the first grid node to the wall appears explicitly in the definition of ωw. Two alternative wall boundary conditions are proposed by Wilcox in [12]. The first one is to calculate ω directly from equation (43) for yn<2.5 instead of solving the ω transport equation in the near-wall region. The second one is to derive ωw from the skin friction velocity using a ‘slightly rough wall' boundary condition. It has already been shown in [18] that the latter approach leads to unsatisfactory results for 3U is the undisturbed flow velocity and L is the ship length. ∞ 4Theoretically ω should be infinite.

lengths, word breaks, heading styles, and other typesetting-specific formatting, however, cannot be retained, and some typographic errors may have been accidentally inserted. Please use the print version of this publication as About this PDF file: This new digital representation of the original work has been recomposed from XML files created from the original paper book, not from the original typesetting files. Page breaks are true to the original; line NUMERICAL PREDICTION OF SCALE EFFECTS IN SHIP STERN FLOWS WITH EDDY-VISCOSITY TURBULENCE MODELS 559 the skin friction coefficient for the flow on a flat plate. Therefore, only the first option is left to avoid the use of a grid- dependent boundary condition. In this paper we shall compare two alternatives: i) adopt equation (43) with Nω including viscous corrections to define the near-wall values of ω, BC1; and ii) obtain ω at the wall from equation (44), BC2. 3 RESULTS AND DISCUSSION 3.1 General All calculations were carried out with the computer code PARNASSOS, [19], which solves the Reynolds Averaged Navier-Stokes equations in their complete form [20]. The test case in this paper is the flow around the Dyne (Mystery) tanker which has earlier been subject of comparative computations in the Gothenburg and Tokyo Workshops, [21] and [22]. Two main reasons justify this choice: the flow is sufficiently complex to test the accuracy of the turbulence models, and, in particular, it exhibits, at least at model scale Reynolds number, the so-called ‘hook shape' of the isolines of the axial velocity at the end of the stern, as a result of the existence of strong bilge vortices. A detailed numerical verification study has been performed with PARNASSOS for this test case, both at model scale Reynolds number, [3], and full scale Reynolds number, [4], which permits the selection of a grid with sufficient resolution. In the present study, five different Reynolds numbers have been considered: 5×106, 2×107, 108, 5×108 and 2×109, with the Reynolds number defined by A Cartesian coordinate system is introduced with the x axis along the undisturbed stream, the z axis vertical positive pointing upwards and y completing a right-hand system. The origin of the coordinate system is located on the forward perpendicular at the ship symmetry plane on the keel line. All the variables presented are made non-dimensional using U∞ and L as the velocity and length reference scales. The computational domain covers only the flow field near the stern. The inlet and outlet plane are x constant planes. The inlet plane is located at x=0.5L and the outlet plane at x=1.25L. The external boundary is an elliptical cylinder, given by: The remaining boundaries are the free surface, plane z=0.056L, the symmetry plane of the ship, y=0, and the hull surface. The volume grids were created with a proprietary elliptic PDE grid generator, based on the GRAPE approach [23]. The number of grid nodes in the streamwise and girthwise direction is the same for the five Reynolds numbers: Nξ=161 and Nζ=41. The number of grid nodes in the normal direction, Nη, increases with Rn. Nη=81 for the lowest Rn and 10 grid lines are added each time Rn is increased, which leads to Nη=121 for Rn=2×109. The grid line spacing in the normal direction is defined by one-dimensional stretching functions, which are tuned to obtain a maximum value of y− at the first layer of grid nodes away from the ship surface of approximately 0.5. Five significant flow parameters were selected to compare the different numerical solutions: • Friction resistance coefficient, • Pressure resistance coefficient, • Wake fraction, Wf: • Maximum cross-stream velocity at x=0.989L, (Vw)max, with • Minimum axial velocity component in the flow field, The maximum cross-stream velocity at x=0.989L is related to the bilge vortex intensity and identifies the existence of streamwise flow separation. The integrals included in the definitions of and Wf are evaluated with Gaussian quadrature rules assuming a bi-linear variation of the unknowns between the grid nodes. The area Ω for the calculation of the wake fraction is the propeller disc, which has been located at x=0.989L with the axis of the propeller at z=0.0166L; the disc radius is R=0.015L, while a zero hub radius has been assumed. the authoritative version for attribution.

lengths, word breaks, heading styles, and other typesetting-specific formatting, however, cannot be retained, and some typographic errors may have been accidentally inserted. Please use the print version of this publication as About this PDF file: This new digital representation of the original work has been recomposed from XML files created from the original paper book, not from the original typesetting files. Page breaks are true to the original; line NUMERICAL PREDICTION OF SCALE EFFECTS IN SHIP STERN FLOWS WITH EDDY-VISCOSITY TURBULENCE MODELS 560 3.2 Wall Boundary Condition for To investigate the influence of the numerical implementation of the wall boundary condition of ω, we have calculated the flow at Rn=5×106 with Menter's version of the k-ω model with the two options considered: BC1, which obtains ω from the theoretical value for y−<2.5 and BC2, which is based on an ad hoc definition of a finite value at the wall. Table 1 presents the five selected flow quantities and the mean and maximum values of vt obtained with BC1 and BC2. Table 1: Comparison of solutions obtained with Menter's k-ω model using different implementations of the ω wall boundary condition. Variable BC1 BC2 1.944 1.612 0.911 0.878 0.609 0.565 Wf 0.343 0.365 (Vw)max −0.075 −0.076 (vt)med×104 0.265 0.257 4 (vt)max×10 1.582 1.504 The differences obtained between the two solutions are certainly not negligible. As one might expect, the friction resistance coefficient, exhibits the largest difference. However, the limiting streamlines of both calculations, which are depicted in figure 1, are similar. Figure 1: Limiting streamlines for Menter's k-ω model with different wall boundary conditions. At the propeller plane, x=0.989L, there are significant differences between the isolines of the axial velocity, as shown in figure 2. With BC1 the speed is definitely lower in the inner wake than with BC2. Figure 2: Axial velocity isolines at x=0.989L for Menter's k-ω model with different wall boundary conditions. These results show that the flow prediction is clearly dependent on the numerical implementation of the ω wall boundary condition. They suggest that the ω behaviour at the wall can hardly be seen as a strong point of the k-ω models. From the results it is not clear which is the best choice, BC1 or BC2, however, as discussed above, the results of BC2 are inherently grid-dependent. Therefore, we will adopt BC1 for the remaining calculations with the k-ω models. 3.3 Scaling Effects the authoritative version for attribution. The results of the five selected flow quantities and the maximum value of vt are given in table 2 for the five Reynolds numbers and for the various turbulence models tested. ∆ stands for the maximum difference between the predictions of the different turbulence models at a given Reynolds number. The differences between predictions with different turbulence models come out as appreciable at model scale Reynolds number but tend to diminish with the increase of the Reynolds number. An exception is found in the maximum cross-stream velocity at the propeller plane; it is the only one of the selected flow variables which does not change monotonically with the Reynolds number. Figure 3 presents the friction resistance coefficients given in table 2. We have tried to compare the results with two friction lines, the ITTC line,

lengths, word breaks, heading styles, and other typesetting-specific formatting, however, cannot be retained, and some typographic errors may have been accidentally inserted. Please use the print version of this publication as About this PDF file: This new digital representation of the original work has been recomposed from XML files created from the original paper book, not from the original typesetting files. Page breaks are true to the original; line NUMERICAL PREDICTION OF SCALE EFFECTS IN SHIP STERN FLOWS WITH EDDY-VISCOSITY TURBULENCE MODELS 561 Table 2: Comparison of solutions obtained with the several turbulence models at different Reynolds numbers. Variable Rn CS SA MT KE-TL KE QZ KW KWM ∆ 5×106 1.573 1.816 1.606 1.561 1.598 1.445 1.939 1.944 0.499 2×107 1.269 1.469 1.335 1.280 1.315 1.197 1.546 1.547 0.350 108 1.022 1.177 1.098 1.059 1.082 0.988 1.213 1.221 0.233 5×108 0.846 0.965 0.916 0.885 0.900 0.824 0.981 0.983 0.141 2×109 0.732 0.825 0.791 0.767 0.776 0.713 0.841 0.843 0.130 5×106 0.598 0.788 0.741 0.713 0.687 0.655 0.923 0.911 0.325 2×107 0.527 0.720 0.648 0.638 0.645 0.614 0.829 0.817 0.320 108 0.466 0.649 0.583 0.580 0.604 0.589 0.744 0.732 0.278 5×108 0.423 0.605 0.547 0.543 0.570 0.575 0.690 0.675 0.267 2×109 0.402 0.578 0.535 0.542 0.565 0.570 0.648 0.638 0.236 5×106 0.528 0.632 0.619 0.563 0.576 0.532 0.616 0.609 0.104 Wf 2×107 0.462 0.547 0.535 0.493 0.512 0.479 0.532 0.530 0.085 108 0.397 0.452 0.454 0.421 0.437 0.415 0.451 0.449 0.057 5×108 0.347 0.393 0.393 0.369 0.380 0.365 0.384 0.386 0.046 2×109 0.315 0.352 0.353 0.334 0.341 0.330 0.345 0.345 0.038 5×106 0.209 0.280 0.290 0.287 0.266 0.263 0.347 0.343 0.138 (Vw)max 2×107 0.199 0.301 0.289 0.286 0.284 0.271 0.378 0.367 0.179 108 0.203 0.302 0.276 0.262 0.284 0.268 0.396 0.377 0.193 5×108 0.206 0.294 0.252 0.229 0.256 0.244 0.395 0.369 0.189 2×109 0.208 0.263 0.228 0.227 0.229 0.228 0.336 0.331 0.128 5×106 −0.012 −0.059 −0.068 −0.028 −0.015 −0.007 −0.075 −0.075 0.063 2×107 −0.004 −0.055 −0.040 −0.007 −0.016 −0.011 −0.065 −0.064 0.061 108 0.000 −0.035 −0.001 0.000 −0.001 0.000 −0.041 −0.041 0.041 5×108 0.000 −0.005 0.000 0.000 0.000 0.000 −0.027 −0.034 0.034 2×109 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 (vt)max ×104 5×106 2.327 1.627 1.259 1.113 1.095 1.046 1.492 1.582 1.281 2×107 1.979 1.338 1.026 0.927 0.965 0.889 1.318 1.361 1.090 108 1.647 1.093 0.845 0.798 0.823 0.742 0.931 1.141 0.905 5×108 1.374 0.917 0.711 0.660 0.708 0.603 0.763 0.862 0.800 2×109 1.173 0.780 0.626 0.556 0.586 0.505 0.700 0.756 0.668 and the Schoenherr line, Since our computation domain covers the aft half of the hull only, we have estimated the equivalent plate friction of the aftbody as where Sw is the wetted surface of the ship included in the computational domain, which is assumed to be half of the total wetted surface. The predictions of all the turbulence models exhibit the correct trend with the increase of the Reynolds number, but there is a clear difference in slope. As a further relevant result, we have plotted the mean wake fraction, Wf, as a function of the Reynolds number in figure 4. It is interesting to note that in both figures 3 and 4 there is good agreement between the SA model and the two k- ω models, KW and KWM. The calculated limiting streamlines at Rn=5×106, Rn=108 and Rn=2×109 are illustrated in figures 5 to 7 for the models CS, SA, MT, KW, KWM and KE. As in the previous results, the differences between the predictions of the various turbulence models tend to diminish with the increase of Rn. Once more, the the authoritative version for attribution.

lengths, word breaks, heading styles, and other typesetting-specific formatting, however, cannot be retained, and some typographic errors may have been accidentally inserted. Please use the print version of this publication as About this PDF file: This new digital representation of the original work has been recomposed from XML files created from the original paper book, not from the original typesetting files. Page breaks are true to the original; line NUMERICAL PREDICTION OF SCALE EFFECTS IN SHIP STERN FLOWS WITH EDDY-VISCOSITY TURBULENCE MODELS 562 results of the SA, KW and KWM models are very similar. At full scale, Rn=2×109, the CS, MT and KE also show good agreement. Figure 3: Friction resistance coefficient of the aft-body, as a function of the Reynolds number for the various turbulence models tested. Figure 5: Limiting streamlines at Rn=5×106. Figure 4: Wake fraction, Wf, as a function of the Reynolds number for the several turbulence models tested. The axial velocity isolines at the propeller plane, x=0.989L, at the same three Reynolds numbers, 5×106, 108 and 2×109, are presented in figures 8 to 10. The turbulence models included are again the CS, SA, MT, KW, KWM and KE. The U1 isolines exhibit a drastic influence of Rn. At model scale, the typical ‘hook shape' does appear for the k-ω models, and to some extent, for the one-equation models5 SA and MT. However, at x=0.989L, the ‘hook shape' tends to disappear with the increase of the Reynolds number. At Rn=2×109, none of the predictions exhibits a ‘hook shape' and differences between the results of the various models, including the algebraic CS model, are rather small. This effect of Rn is related to the stretching of the bilge vortex, generated within the ship boundary layer which reduces its thickness with the increase of Rn. Figures 11 to 13 illustrate the cross-stream velocity field at x=0.989L for the same three Reynolds numbers. The plots include the CS, MT and KWM models. Although there are some differences between the predictions of the three models even at Rn=2×109, the authoritative version for attribution. 5As mentioned before, these prediction can be easily improved with a simple correction to the production term of the transport equation of the turbulent quantity.

lengths, word breaks, heading styles, and other typesetting-specific formatting, however, cannot be retained, and some typographic errors may have been accidentally inserted. Please use the print version of this publication as About this PDF file: This new digital representation of the original work has been recomposed from XML files created from the original paper book, not from the original typesetting files. Page breaks are true to the original; line NUMERICAL PREDICTION OF SCALE EFFECTS IN SHIP STERN FLOWS WITH EDDY-VISCOSITY TURBULENCE MODELS 563 the stretching of the bilge vortex with the increase of Rn is clear for the three models. Figure 7: Limiting streamlines at Rn=2×109. Figure 6: Limiting streamlines at Rn=108. We should note that the present results do not imply that the typical ‘hook shape' of the axial velocity isolines disappears with the increase of Rn, it just appears further downstream. Figure 14 presents the velocity field at x=1.1L obtained with the KWM model for Rn=2×109. At this location, the bilge vortex is almost axisymmetric and the U1 plot shows the typical ‘hook shape', which is understandbly weaker than at the propeller plane at model scale, because the bilge vortex has not only rolled up in the near wake but it has also diffused. 4 CONCLUSIONS We have presented results of a numerical investigation of scaling effects in ship stern flows using algebraic, one- equation and two-equation turbulence models. The turbulence models were all implemented without any special tuning dependent on the Reynolds number. For the two-equation k-ω models, we have pointed out the deficiencies of a widely accepted numerical implementation of the wall boundary condition of ω. The results of the calculation of the flow around the Dyne (Mystery) tanker at five Reynolds numbers, 5×106, 2×107, 8, 5×108 and 2×109, suggest the following conclusions: 10 • It is possible to simulate numerically ship stern flows from model up to full scale Reynolds numbers with the most popular eddy-viscosity turbulence models, including algebraic, one-equation and two-equation models. • In global terms, the predictions exhibit the same trend in the flow field with the increase of the Reynolds number for all the turbulence models. the authoritative version for attribution.

About this PDF file: This new digital representation of the original work has been recomposed from XML files created from the original paper book, not from the original typesetting files. Page breaks are true to the original; line lengths, word breaks, heading styles, and other typesetting-specific formatting, however, cannot be retained, and some typographic errors may have been accidentally inserted. Please use the print version of this publication as the authoritative version for attribution. at Rn=5×106. x=0.989L obtained Figure 6: Axial velocity isolines at at x=0.989L obtained at Rn=108. Figure 7: Axial velocity isolines NUMERICAL PREDICTION OF SCALE EFFECTS IN SHIP STERN FLOWS WITH EDDY-VISCOSITY TURBULENCE MODELS 564

About this PDF file: This new digital representation of the original work has been recomposed from XML files created from the original paper book, not from the original typesetting files. Page breaks are true to the original; line lengths, word breaks, heading styles, and other typesetting-specific formatting, however, cannot be retained, and some typographic errors may have been accidentally inserted. Please use the print version of this publication as the authoritative version for attribution. at Rn=2×109. x=0.989L obtained Figure 8: Axial velocity isolines at Rn=5×106. Figure 9: Transverse velocity field at x=0.989L obtained at NUMERICAL PREDICTION OF SCALE EFFECTS IN SHIP STERN FLOWS WITH EDDY-VISCOSITY TURBULENCE MODELS 565

About this PDF file: This new digital representation of the original work has been recomposed from XML files created from the original paper book, not from the original typesetting files. Page breaks are true to the original; line lengths, word breaks, heading styles, and other typesetting-specific formatting, however, cannot be retained, and some typographic errors may have been accidentally inserted. Please use the print version of this publication as the authoritative version for attribution. Rn=108. Figure 10: Transverse velocity field at x=0.989L obtained at Rn=2×109. Figure 11: Transverse velocity field at x=0.989L obtained at NUMERICAL PREDICTION OF SCALE EFFECTS IN SHIP STERN FLOWS WITH EDDY-VISCOSITY TURBULENCE MODELS 566

lengths, word breaks, heading styles, and other typesetting-specific formatting, however, cannot be retained, and some typographic errors may have been accidentally inserted. Please use the print version of this publication as About this PDF file: This new digital representation of the original work has been recomposed from XML files created from the original paper book, not from the original typesetting files. Page breaks are true to the original; line NUMERICAL PREDICTION OF SCALE EFFECTS IN SHIP STERN FLOWS WITH EDDY-VISCOSITY TURBULENCE MODELS 567 Figure 12: Velocity field at x=1.1L obtained with Menter's k-ω model at Rn=2×109. Above: axial velocity, U1, isolines. Below: cross-stream velocity field. • The discrepancies between flow fields obtained with different turbulence models at a given Reynolds number tend to decrease with the increase of the Reynolds number. • Although the performance of the k-ω models seems to be very encouraging, the predictions depend on the numerical implementation of the ω boundary condition at a solid surface. Some implementations advised in the open literature are unacceptable. The present results reinforce the need for reliable experimental data at full scale Reynolds number for validation purposes. All eddy-viscosity turbulence models, used here, were essentially developed for boundary-layers at moderate Reynolds numbers. REFERENCES [1] Watson S.J.P., Bull P.W.—The Scaling of High Reynolds Number Viscous Flow Predictions Using CFD Techniques—Third Osaka Colloquium, Osaka, Japan. [2] Eça L., Hoekstra M.—Numerical Calculations of Ship Stern Flows at Full-Scale Reynolds Numbers Twenfirst Symposium on Naval Ship Hydrodynamics, Trondheim, June 1996. [3] Hoekstra M., Eça L.—An Example of Error Quantification of Ship-Related CFD Results—7th Numerical Ship Hydrodynamics Conference, Nantes, July 1999. [4] Eça L., Hoekstra M.—On the Numerical Verification of Ship Stern Flow Calculations—1st MARNET Workshop, Barcelona, November 1999. [5] Deng G.B., Visonneau M.—Comparison of Explicit Algebraic Stress Models and Second-Order Turbulence Closures for Steady Flows around Ships —7th Numerical Ship Hydrodynamics Conference, Nantes, July 1999. [6] Cebeci T., Smith A.M.O.—Analysis of Turbulent Boundary Layers.—Academic Press, November 1984. [7] Baldwin B.S., Lomax H.—Thin Layer Approximation and Algebraic Models for Separated Turbulent Flows—AIAA Paper 78–257, January 1978. [8] Spalart P.R., Allmaras S.R.—A One-Equations Turbulence Model for Aerodynamic Flows—AIAA 30th Aerospace Sciences Meeting, Reno, January 1992. [9] Menter F.R.—Eddy Viscosity Transport Equations and Their Relation to the k-ε Model—Journal of Fluids Engineering, Vol. 119, December 1997, pp. 876–884. [10] Chen H.C, Patel V.C.—Practical Near-Wall Turbulence Models for Complex Flows Including Separation.—AIAA 19th Fluid Dynamics, Plasma Dynamics and Lasers Conference, June 8–10, 1987. the authoritative version for attribution.

lengths, word breaks, heading styles, and other typesetting-specific formatting, however, cannot be retained, and some typographic errors may have been accidentally inserted. Please use the print version of this publication as About this PDF file: This new digital representation of the original work has been recomposed from XML files created from the original paper book, not from the original typesetting files. Page breaks are true to the original; line NUMERICAL PREDICTION OF SCALE EFFECTS IN SHIP STERN FLOWS WITH EDDY-VISCOSITY TURBULENCE MODELS 568 [11] Chien K.Y—Prediction of Channel and Boundary-Layer Flows with a Low-Reynolds-Number Turbulence Model — AIAA Journal, January 1992, pp. 33–38. [12] Wilcox D.C.—Turbulence Modeling for CFD—DWC Industries 1993. [13] Menter F.R.—Two-Equation Eddy-Viscosity Turbulence Models for Engineering Applications—AIAA Journal, Vol. 32, August 1994, pp. 1598– 1605. [14] Gibson M.M., Dafa Alla A.A.—Two-Equation Model for Turbulent Wall Flow—AIAA Journal, Vol. 33, August 1995, pp. 1514–1518. [15] Baldwin S.B., Barth T.J.—A One-Equation Turbulence Transport Model for High Reynolds Wall-Bounded Flows—AIAA Paper 91–0610, 29th Aerospace Sciences Meeting, Reno Nevada, January 1991. [16] Hoekstra M.—Numerical Simulation of Ship Stern Flows with a Space-Marching Navier-Stokes Method—PhD Thesis Delft University, 1999. [17] Wolfshtein M.—The Velocity and Temperature Distribution in One-Dimensional Flow with Turbulence Augmentation and Pressure Gradient.— International Journal of Heat and Mass Transfer, Vol. 12, 1969, pp. 301–318. [18] Goldberg U., Peroomian O., Chakravarthy S.,—A Wall Distance-Free k-ε Model with Enhanced Near-Wall Treatment—Journal of Fluids Engineering, Vol. 120, September 1998, 457–462. [19] Hoekstra M., Eça L.—PARNASSOS: An Efficient Method for Ship Stern Flow Calculation—Third Osaka Colloquium on Advanced CFD Applications to Ship Flow and Hull Form Design, Osaka, Japan, 1998. [20] van der Ploeg A., Eça L., Hoekstra M.—Combining Accuracy and Efficiency with Robustness in Ship Stern Flow Computation —Twen-third Symposium on Naval Ship Hydrodynamics, September 2000. [21] Larsson L., Patel V.C., Dyne G. (eds.)—Ship Viscous Flow.—Proceedings of 1990 SSPA-CTH-IIHR Workshop, Flowtech International AB, Research Report Nº2, Gothenburg, June 1991. [22] Proceedings of CFD Workshop Tokyo 1994, Ship Research Institute Tokyo, March 1994. [23] Sorenson R.L.—Three-Dimensional Grid Generation about Fighter Aircraft for Zonal Finite-Difference Computations —AIAA 86–0429. AIAA 24th Aerospace Sciences Conference, 1986, Reno, NV. the authoritative version for attribution.

Next: The Experimental and Numerical Study of Flow Structure and Water Noise Caused by Roughness of the Body »
Twenty-Third Symposium on Naval Hydrodynamics Get This Book
×
Buy Paperback | $550.00
MyNAP members save 10% online.
Login or Register to save!
Download Free PDF

"Vive la Revolution!" was the theme of the Twenty-Third Symposium on Naval Hydrodynamics held in Val de Reuil, France, from September 17-22, 2000 as more than 140 experts in ship design, construction, and operation came together to exchange naval research developments. The forum encouraged both formal and informal discussion of presented papers, and the occasion provides an opportunity for direct communication between international peers.

This book includes sixty-three papers presented at the symposium which was organized jointly by the Office of Naval Research, the National Research Council (Naval Studies Board), and the Bassin d'Essais des Carènes. This book includes the ten topical areas discussed at the symposium: wave-induced motions and loads, hydrodynamics in ship design, propulsor hydrodynamics and hydroacoustics, CFD validation, viscous ship hydrodynamics, cavitation and bubbly flow, wave hydrodynamics, wake dynamics, shallow water hydrodynamics, and fluid dynamics in the naval context.

  1. ×

    Welcome to OpenBook!

    You're looking at OpenBook, NAP.edu's online reading room since 1999. Based on feedback from you, our users, we've made some improvements that make it easier than ever to read thousands of publications on our website.

    Do you want to take a quick tour of the OpenBook's features?

    No Thanks Take a Tour »
  2. ×

    Show this book's table of contents, where you can jump to any chapter by name.

    « Back Next »
  3. ×

    ...or use these buttons to go back to the previous chapter or skip to the next one.

    « Back Next »
  4. ×

    Jump up to the previous page or down to the next one. Also, you can type in a page number and press Enter to go directly to that page in the book.

    « Back Next »
  5. ×

    To search the entire text of this book, type in your search term here and press Enter.

    « Back Next »
  6. ×

    Share a link to this book page on your preferred social network or via email.

    « Back Next »
  7. ×

    View our suggested citation for this chapter.

    « Back Next »
  8. ×

    Ready to take your reading offline? Click here to buy this book in print or download it as a free PDF, if available.

    « Back Next »
Stay Connected!