You are on page 1of 141

Ship Science

Evaluation of Slamming Impact Pressures and Forces on Ship-like Sections

Individual Project Final Report


Enrico Anderlini

Project Supervisor: Prof. P. Temarel

21st December 2011

Abstract
This report entails the study of slamming impact pressures and forces against rigid body dropping onto a calm free-surface. These have been performed numerically with CFX on wedge and ship-like section geometries for various impact speeds and heel angles, where in the latter case three-dimensional simulations have also been undertaken. The results have been compared against experimental measurements for validation (Aarsnes, 1996; Lewis, et al., 2010). The data from the wedge simulations do not agree well with the experiments because of limited domain size and hydroelastic effects. In addition, serious instabilities in the solver occur for the case where the wedge density is less than that of water due to errors in the rigid body solution feature of CFX. The ship-like section numerical results compare much better with the experimental measurements, especially for the two-dimensional cases. The three-dimensional simulations present lower slamming pressures and loads, require much greater computational memory and time and cannot model domains of small size because of the deforming mesh approach employed. Furthermore, the numerical and experimental data have been employed to validate the use of empirical theories developed in the 70s. Stavovy and Chuangs (1976) method seems to provide good impact force agreement, while Lloyds Registers rules (2011 a; 2011 b) tend to overestimate the results, as it is expected from a classification society. Nevertheless, the exponential time distribution used (Kawakami, et al., 1977, cited in Belik, 1982) is not appropriate in approximating the slamming behaviour of the ship-like section.

Acknowledgements
Many thanks to Prof. Temarel, who, when telling me at the project presentation: Be careful, though, that you may find yourself stuck at the end of November with the simulations still not running as you wish them to do, was absolutely right. Without his help, I would not have been able to solve all the issues I encountered during the project. I would also like to acknowledge the valuable lecture by Dr. Bressloff (2011) on good CFD practice, which aided me during the first stages of the project. In addition, I would like to mention the users of the forums at <www.cfd-online.com> [accessed on 20/12/2011] and especially Satish Patange from the ANSYS customer portal for the support they gave me with the software, especially concerning the rigid body solver errors. Finally, I would like to acknowledge the CFD surgery classes provided on a weekly basis by Dr. Xie, although they mainly focused on FLUENT solutions of fixed meshes.

ii

Table of Contents
Abstract .......................................................................................................................................... i Acknowledgements ....................................................................................................................... ii List of figures ................................................................................................................................. v List of tables .................................................................................................................................. x Nomenclature .............................................................................................................................. xii Abbreviations ............................................................................................................................... xv 1.01.11.22.03.03.13.23.33.43.53.63.74.05.05.15.25.35.45.56.06.16.26.36.46.57.08.0Introduction ...................................................................................................................... 1 Aim and Objectives ....................................................................................................... 1 Report Structure............................................................................................................ 2 Literature Review .............................................................................................................. 3 Methodology ..................................................................................................................... 8 Type of Simulations ....................................................................................................... 9 Geometry Selection ....................................................................................................... 9 Mesh Generation ........................................................................................................ 12 Turbulence Models...................................................................................................... 13 Multiphase Flow .......................................................................................................... 14 Boundary Conditions ................................................................................................... 14 Error Analysis............................................................................................................... 15 Theoretical and Empirical Approaches............................................................................ 16 Numerical and Empirical Results ..................................................................................... 20 Numerical Results for the Wedge by Lewis, et al. (2010) ........................................... 20 Numerical Results for Aarsnes (1996) Wedge ........................................................... 25 Numerical Results for the Ship-like Section ................................................................ 28 Numerical Results for the Three-dimensional Ship-like Section ................................. 44 Empirical Results ......................................................................................................... 47 Data Analysis and Discussion .......................................................................................... 52 Lewis, et al.s (2010) Wedge Analysis.......................................................................... 52 Aarsnes (1996) Wedge Analysis ................................................................................. 53 Ship-like Section Data Analysis.................................................................................... 53 Three-dimensional Ship-like Section Analysis ............................................................. 55 Discussion of the Empirical Results ............................................................................. 56 Conclusions ..................................................................................................................... 58 Future Work Recommendations ..................................................................................... 60 iii

9.010.011.0-

References ....................................................................................................................... 62 Appendix I-Further Information on Numerical Methods ............................................ 65 Appendix II-Computational Fluid Dynamics Theory .................................................... 66 Additional considerations on transient simulations ............................................... 68 Additional considerations on the geometry description and rigid body solution .. 68 Additional considerations on mesh generation ...................................................... 69 Additional considerations on turbulence models ................................................... 72 Additional considerations on the boundary conditions .......................................... 74 Additional considerations on multiphase flow ....................................................... 77 Additional considerations on error analysis............................................................ 78 Appendix III-CFX Geometry description ...................................................................... 80 Appendix IV-Further discussion of the empirical approaches .................................... 86 Appendix V-Preliminary studies for the CFX simulations ............................................ 88 Results from the turbulence and interphase models analysis ................................ 88 Results from a study in the influence of different mesh stiffness models.............. 91 Results from a study in domain size ........................................................................ 94 Discussion of the preliminary simulations .............................................................. 96 Appendix VI-Experimental results for the wedge by Lewis, et al. (2010) ................... 98 Appendix VII- Visual Results ...................................................................................... 101 Wedge by Lewis, et al. (2010) ............................................................................... 101 Wedge by Aarsnes (1996) ..................................................................................... 107 Two-dimensional ship-like section ........................................................................ 109 Heeled ship-like section ........................................................................................ 114 Three-dimensional ship-like section ..................................................................... 121 Appendix VIII- CEL equations .................................................................................... 124

11.111.211.311.411.511.611.712.013.014.014.114.214.314.415.016.016.116.216.316.416.517.0-

iv

List of figures
Figure 1: Plots for interpolating the k1 value for the lower part of the ship-like (bow) section, taken from Ochi and Motter (1973). ........................................................................................... 17 Figure 2: Pressure against time at P1 and P4 for a study in mesh density on the wedge by Lewis, et al. (2010) for low mass and impact speed. ............................................................................. 22 Figure 3: Pressure against time at P2 and P5 for a study in mesh density on the wedge by Lewis, et al. (2010) for low mass and impact speed. ............................................................................. 22 Figure 4: Pressure against time at P3 and P6 for a study in mesh density on the wedge by Lewis, et al. (2010) for low mass and impact speed. ............................................................................. 23 Figure 5: Pressure against time at P1 and P4 for the wedge by Lewis, et al. (2010) for varying mass and impact speed and constant mesh density. ................................................................. 24 Figure 6: Pressure against time at P2 and P5 for the wedge by Lewis, et al. (2010) for varying mass and impact speed and constant mesh density. ................................................................. 24 Figure 7: Pressure against time at P3 and P6 for the wedge by Lewis, et al. (2010) for varying mass and impact speed and constant mesh density. ................................................................. 25 Figure 8: Pressure versus time at P1 for the wedge by Aarsnes (1996). .................................... 26 Figure 9: Pressure versus time at P3 for the wedge by Aarsnes (1996). .................................... 26 Figure 10: Pressure versus time at P5 for the wedge by Aarsnes (1996). .................................. 27 Figure 11: Variation of the horizontal force per unit length with time for Aarsnes (1996) wedge. ..................................................................................................................................................... 27 Figure 12: Variation of the vertical force per unit length with time for Aarsnes (1996) wedge. ..................................................................................................................................................... 28 Figure 13: Total computation time as a function of the number of elements for the bow section simulations at low impact speed................................................................................................. 29 Figure 14: Pressure at P1 against time for the ship-like (bow) section with no heel angle and low impact speed for three mesh densities. ............................................................................... 30 Figure 15: Pressure at P2 against time for the bow section with no heel angle and low impact speed for three mesh densities................................................................................................... 31 Figure 16: Pressure at P3 against time for the bow section with no heel angle and low impact speed for three mesh densities................................................................................................... 31 Figure 17: Pressure at P4 against time for the bow section with no heel angle and low impact speed for three mesh densities................................................................................................... 32 Figure 18: Horizontal force per unit length against time for the bow section with no heel angle and low impact speed. ................................................................................................................ 32 Figure 19: Vertical force per unit length against time for the bow section with no heel angle and low impact speed. ................................................................................................................ 33 Figure 20: Variation of the pressure at P1 with time for the bow section with no heel angle and high impact speed. ...................................................................................................................... 33 Figure 21: Variation of the pressure at P3 with time for the bow section with no heel angle and high impact speed. ...................................................................................................................... 34 Figure 22: Variation of the horizontal force per unit length withtime for the bow section with no heel angle and high impact speed. ........................................................................................ 34

Figure 23: Variation of the vertical force per unit length with time for the bow section with no heel angle and high impact speed. ............................................................................................. 35 Figure 24: Variation of the pressure at P1 with time for the bow section with a 9.8 heel angle and low impact speed. ................................................................................................................ 35 Figure 25: Variation of the pressure at P2 with time for the bow section with a 9.8 heel angle and low impact speed. ................................................................................................................ 36 Figure 26: Variation of the pressure at P3 with time for the bow section with a 9.8 heel angle and low impact speed. ................................................................................................................ 36 Figure 27: Variation of the pressure at P4 with time for the bow section with a 9.8 heel angle and low impact speed. ................................................................................................................ 37 Figure 28: Variation of the pressure at P1 with time for the bow section with a 28.3 heel angle. ..................................................................................................................................................... 37 Figure 29: Variation of the pressure at P2 with time for the bow section with a 28.3 heel angle. ..................................................................................................................................................... 38 Figure 30: Variation of the pressure at P3 with time for the bow section with a 28.3 heel angle. ..................................................................................................................................................... 38 Figure 31: Variation of the pressure at P4 with time for the bow section with a 28.3 heel angle. ..................................................................................................................................................... 39 Figure 32: Variation of the pressure at P1 with time for the bow section with a 9.8 heel angle and high impact speed. ............................................................................................................... 39 Figure 33: Variation of the pressure at P2 with time for the bow section with a 9.8 heel angle and high impact speed. ............................................................................................................... 40 Figure 34: Variation of the pressure at P3 with time for the bow section with a 9.8 heel angle and high impact speed. ............................................................................................................... 40 Figure 35: Variation of the pressure at P4 with time for the bow section with a 9.8 heel angle and high impact speed. ............................................................................................................... 41 Figure 36: Horizontal force per unit length against time for the bow section with a 9.8 heel angle and low impact speed........................................................................................................ 41 Figure 37: Vertical force per unit length against time for the bow section with a 9.8 heel angle and low impact speed. ................................................................................................................ 42 Figure 38: Horizontal force per unit length against time for the bow section with a 28.3 heel angle. ........................................................................................................................................... 42 Figure 39: Vertical force per unit length against time for the bow section with a 28.3 heel angle. ........................................................................................................................................... 43 Figure 40: Horizontal force per unit length against time for the bow section with a 9.8 heel angle and high impact speed. ..................................................................................................... 43 Figure 41: Vertical force per unit length against time for the bow section with a 9.8 heel angle and high impact speed. ............................................................................................................... 44 Figure 42: Pressure variation with time at P1 for the bow section with no heel angle and low impact speed from three-dimensional simulations. ................................................................... 45 Figure 43: Pressure variation with time at P2 for the bow section with no heel angle and low impact speed from three-dimensional simulations. ................................................................... 45 Figure 44: Pressure variation with time at P3 for the bow section with no heel angle and low impact speed from three-dimensional simulations. ................................................................... 46

vi

Figure 45: Pressure variation with time at P4 for the bow section with no heel angle and low impact speed from three-dimensional simulations. ................................................................... 46 Figure 46: Vertical force per unit length against time for the bow section with no heel angle and low impact speed from three-dimensional simulations. ..................................................... 47 Figure 47: Impact pressure against time from empirical, numerical and experimental data for the wedge with low impact speed. ............................................................................................. 48 Figure 48: Impact pressure against time from empirical, numerical and experimental data for the wedge with high impact speed. ............................................................................................ 49 Figure 49: Impact pressure against time from empirical, numerical and experimental data for the bow section with low impact speed. .................................................................................... 50 Figure 50: Impact pressure against time from empirical, numerical and experimental data for the bow section with high impact speed. ................................................................................... 50 Figure 51: Vertical impact force per unit length against time from empirical, numerical and experimental data for the lower portion of the bow section with low impact speed................ 51 Figure 52: Vertical impact force per unit length against time from empirical, numerical and experimental data for the lower portion of the bow section with high impact speed. ............. 51 Figure 53: Section view of the wedge by Lewis, et al. (2010). .................................................... 83 Figure 54: Section view of the wedge by Aarsnes (1996). .......................................................... 83 Figure 55: Section view of the bow section. ............................................................................... 83 Figure 56: Domain dimensions for the wedge by Lewis, et al. (2010). ....................................... 84 Figure 57: Domain dimensions for the wedge by Aarsnes (1996). ............................................. 84 Figure 58: Domain dimensions for the bow section. Only the case with no heel angle is presented. ................................................................................................................................... 85 Figure 59: Variation of pressure at P1 and P4 with time for different turbulence and interphase models for the same mesh and wedge geometry. ..................................................................... 90 Figure 60: Variation of pressure at P2 and P5 with time for different turbulence and interphase models for the same mesh and wedge geometry. ..................................................................... 90 Figure 61: Variation of pressure at P3 and P6 with time for different turbulence and interphase models for the same mesh and wedge geometry. ..................................................................... 91 Figure 62: Variation of pressure at P1 and P4 with time for different mesh stiffness models for the same mesh and wedge geometry. ........................................................................................ 92 Figure 63: Variation of pressure at P2and P5 with time for different mesh stiffness models for the same mesh and wedge geometry. ........................................................................................ 93 Figure 64: Variation of pressure at P3 and P6 with time for different mesh stiffness models for the same mesh and wedge geometry. ........................................................................................ 93 Figure 65: Variation of pressure at P1 and P4 with time for different drop heights. ................. 95 Figure 66: Variations of pressure at P2 and P5 with time for different drop heights................. 95 Figure 67: Variation of pressure at P3 and P6 with time for different drop heights .................. 96 Figure 68: Pressure trace for a 0.50 m drop, with a wedge mass of 23.4 kg, taken from figure 18 in the paper by Lewis, et al. (2010). ............................................................................................ 98 Figure 69: Pressure trace for a 0.50 m drop, with a wedge mass of 33.4 kg, taken from figure 19 in the paper by Lewis, et al. (2010). ............................................................................................ 99 Figure 70: Pressure trace for a 0.75 m drop, with a wedge mass of 23.4 kg, taken from figure 20 in the paper by Lewis, et al. (2010). ............................................................................................ 99

vii

Figure 71: Pressure trace for a 0.75 m drop, with a wedge mass of 23.4 kg, taken from figure 21 in the paper by Lewis, et al. (2010). .......................................................................................... 100 Figure 72: Experimental and numerical flow description around the wedge, taken from Lewis (2011). ....................................................................................................................................... 100 Figure 73 A,B: Water volume fraction for the wedge by Lewis, et al. (2010) at t=0s for two different mesh densities. .......................................................................................................... 101 Figure 74: Hydrostatic pressure distribution around the wedge by Lewis, et al. (2010). ......... 102 Figure 75: Node density around the wedge by Lewis, et al. (2010) for a coarse mesh. ........... 102 Figure 76 A,B: Pressure and water volume fraction contours around the wedge at t=0.065 s. ................................................................................................................................................... 103 Figure 77 A,B: Pressure and water volume fraction contours around the wedge at t=0.080 s. ................................................................................................................................................... 103 Figure 78 A,B: Pressure and water volume fraction contours around the wedge at t=0.100 s. ................................................................................................................................................... 104 Figure 79: Pressure distribution around the wedge at t=0.115 s. ............................................ 104 Figure 80 A,B: Pressure and water volume fraction contours around the wedge at t=0.116 s. ................................................................................................................................................... 105 Figure 81 A,B: Pressure and water volume fraction contours around the wedge at t=0.149 s. ................................................................................................................................................... 106 Figure 82: Pressure distribution around the wedge at t=0.150 s. ............................................ 106 Figure 83: Node density and water volume fraction around the wedge by Aarsnes (1996) at t=0 s. ................................................................................................................................................ 107 Figure 84: Water volume fraction for the wedge by Aarsnes (1996) at impact time. .............. 107 Figure 85: Air entrapment under the wedge by Aarsnes (1996) at t=0.025 s. ......................... 108 Figure 86: Pressure distribution around the wedge by Aarsnes (1996) at t=0.025 s................ 108 Figure 87: Flow around the wedge by Aarsnes (1996) at t=0.100 s. ......................................... 109 Figure 88 A,B: Hydrostatic pressure and water volume fraction contours for the bow section with no heel angle and low impact speed at t=0 s.................................................................... 109 Figure 89 A,B: Pressure and water volume fraction contours for the bow section with no heel angle and low impact speed at t=0.175 s.................................................................................. 110 Figure 90 A,B: Pressure and water volume fraction contours for the bow section with no heel angle and low impact speed at impact time (t=0.012s). ........................................................... 111 Figure 91 A,B: Pressure and water volume fraction contours for the bow section with no heel angle and low impact speed at t=0.050 s.................................................................................. 111 Figure 92 A,B: Pressure and water volume fraction contours for the bow section with no heel angle and low impact speed at t=0.100 s.................................................................................. 112 Figure 93 A,B: Pressure and water volume fraction contours for the bow section with no heel angle and low impact speed at t=0.125 s.................................................................................. 112 Figure 94 A,B: Pressure and water volume fraction contours for the bow section with no heel angle and low impact speed at t=0.150 s.................................................................................. 113 Figure 95 A,B: Node position and flow around the bow section with no heel angle and low impact speed at t=0 and t=0.175 s. ........................................................................................... 113 Figure 96: Flow around the bow section at t=0.1 s with a 9.8 heel angle and low impact speed. ................................................................................................................................................... 114

viii

Figure 97: Pressure distribution around the bow section at t=0.1 s for a 9.8 heel angle and low impact speed. ............................................................................................................................ 114 Figure 98: Flow around the bow section at t=0.15 s with a 9.8 heel angle and low impact speed. ........................................................................................................................................ 115 Figure 99: Pressure distribution around the bow section at t=0.15 s for a 9.8 heel angle and low impact speed. ..................................................................................................................... 115 Figure 100: Flow around the bow section at t=0.175 s with a 9.8 heel angle and low impact speed. ........................................................................................................................................ 116 Figure 101: Pressure distribution around the bow section at t=0.175 s for a 9.8 heel angle and low impact speed. ..................................................................................................................... 116 Figure 102: Flow around the bow section at t=0.07 s with a 28.3 heel angle and low impact speed. ........................................................................................................................................ 117 Figure 103: Pressure distribution around the bow section at t=0.07 s for a 28.3 heel angle and low impact speed. ..................................................................................................................... 117 Figure 104: Flow around the bow section at t=0.1 s with a 28.3 heel angle and low impact speed. ........................................................................................................................................ 118 Figure 105: Pressure distribution around the bow section at t=0.1 s for a 28.3 heel angle and low impact speed. ..................................................................................................................... 118 Figure 106: Flow around the bow section at t=0.125 s with a 28.3 heel angle and low impact speed. ........................................................................................................................................ 119 Figure 107: Pressure distribution around the bow section at t=0.125 s for a 28.3 heel angle and low impact speed. .............................................................................................................. 119 Figure 108: Flow around the bow section at t=0.15 s with a 28.3 heel angle and low impact speed. ........................................................................................................................................ 120 Figure 109: Pressure distribution around the bow section at t=0.15 s for a 28.3 heel angle and low impact speed. ..................................................................................................................... 120 Figure 110: Node position in the section view around the three-dimensional bow section at t=0.175 s. ................................................................................................................................... 121 Figure 111: Enlargement of the flow around the tip of the three-dimensional bow section at t=0.175 s. ................................................................................................................................... 121 Figure 112: Pressure distribution around the three-dimensional wedge at t=0.175 s in the profile view of file x10. .............................................................................................................. 122 Figure 113: Flow around the three-dimensional wedge at t=0.175 s in the profile view of file x10. ............................................................................................................................................ 122 Figure 114: Pressure distribution around the three-dimensional wedge at t=0.175 s in the profile view of file x1.5. ............................................................................................................. 123 Figure 115: Water volume fraction contour around the three-dimensional wedge at t=0.175 s in the profile view of file x1.5.................................................................................................... 123

ix

List of tables
Table 1: Input y-velocity and heel angles for all simulations. ..................................................... 11 Table 2: Input mass and length for all simulations. .................................................................... 12 Table 3: Measures of mesh quality. ............................................................................................ 13 Table 4: Reynolds numbers of all simulations............................................................................. 13 Table 5: Main particulars of the bow section to a depth of 10% of the moulded depth. .......... 17 Table 6: Length and deadrise angle of the segments the bow section has been discretised into and corresponding k1 value. ........................................................................................................ 19 Table 7: Number of nodes and elements (hexa. are hexahedra), mesh density, time step intervals, Root Mean Square value of the Courant Number, total wall-clock time, minimum orthogonality angle, maximum expansion factor and aspect ratio for simulations on the wedge by Lewis, et al. (2010) for constant input speed and mass. ........................................................ 21 Table 8: Input velocity, mass, simulation time, Root Mean Square value of the Courant Number, total wall-clock time, minimum orthogonality angle, maximum expansion factor and aspect ratio for simulations on the wedge by Lewis, et al. (2010) for constant mesh density. l. stands for low, h. for high, m. for mass and s. for input speed. ............................................................. 23 Table 9: Number of nodes and elements (hexa. are hexahedra), mesh density, time step, Root Mean Square value of the Courant number, total wall-clock time, minimum mesh displacement, minimum orthogonality angle, maximum expansion factor and aspect ratio for simulations on the wedge by Aarsnes (1996). ..................................................................................................... 25 Table 10: Number of nodes and elements (hexa. are hexahedra), mesh density, time steps, Root Mean Square value of the Courant number, simulation time, total wall-clock time, number of processor-cores, minimum mesh displacement, minimum orthogonality angle, maximum expansion factor and aspect ratio for simulations on the bow section with no heel angle. ........................................................................................................................................... 29 Table 11: Number of nodes and elements (hexa. are hexahedra), mesh density, simulation time, Root Mean Square value of the Courant number, total wall-clock time, number of processorcores, minimum mesh displacement, minimum orthogonality angle, maximum expansion factor and aspect ratio for simulations on the bow section with heel angles. ........................... 30 Table 12: Number of nodes and elements (pyr. are pyramids), number of processor-cores, Root Mean Square value of the Courant number, total wall-clock time, minimum mesh displacement, minimum orthogonality angle, maximum expansion factor and aspect ratio for the threedimensional simulations of the bow section. ............................................................................. 44 Table 13: Maximum impact pressure and force per unit length for the wedge and bow section geometries from all empirical methods presented in section 4.0. wed. stands for wedge, sec. for bow section, l. for low, h. for high and s. for speed. ............................................................. 47 Table 14: Location of the six monitor points for the wedge by Lewis, et al. (2010) ................... 81 Table 15: Location of the five monitor points for the wedge by Aarsnes (1996) prior to rotation. ..................................................................................................................................................... 81 Table 16: Coordinate points employed to produce the surface of the bow section. The location of the four monitor points is also highlighted. ........................................................................... 81

Table 17: Location of the monitor points employed by CFX and the empirical procedure by Stavovy and Chuang (1976) to calculate the force per unit length against the lower portion of the bow section. .......................................................................................................................... 82 Table 18: Length and deadrise angle of the segments the lower part of the bow section has been discretised into. .................................................................................................................. 82 Table 19: Peak pressure magnitude calculated through the theory by Von Karman (1929, cited in Bertram, 2000) for the wedge for both impact speeds. ......................................................... 86 Table 20: Number of elements (tetra. are tetrahedra, pyr. pyramids), total wall-clock time, turbulence and interphase models for the simulations performed to study the influence of turbulence and interphase models for the wedge by Lewis, et al. (2010). ................................ 89 Table 21: Number of elements (tetra. are tetrahedra, pyr. pyramids), total wall-clock time and mesh stiffness model for the wedge by Lewis, et al. (2010) for an analysis of the mesh stiffness models. ........................................................................................................................................ 92 Table 22: Number of elements (tetra. are tetrahedra, pyr. pyramids), total wall-clock time, initial drop height and speed for a study in domain size and initial height above the water surface of the wedge by Lewis, et al. (2010)............................................................................... 94

xi

Nomenclature
: perimeter of the sides of the lower portion of the bow section [L] : sectional area of the bow section up to 10% of the moulded depth [L2] : half-breadth of the bow section to a depth of 10% of the moulded depth [L] : half-width of the flat bottom of the bow section [L] : half-width of the wet area of the wedge [L] : constant employed in the equation in the log-law layer [dimensionless]

: sectional area coefficient for lower part of the bow section [dimensionless] : Courant number [dimensionless] : maximum pressure coefficient [dimensionless] : 10% of the moulded depth of the bow section [L] : function of time : peak slamming impact force [M L T -2] : Froude number [dimensionless] m/s2: gravitational acceleration constant [L T -2] : specific static enthalpy [L2 T-2] in the energy and total enthalpy equations; height of the rigid body [L] otherwise : specific total enthalpy [L2 T-2] : Cartesian tensor subscripts, which can be 1, 2 or 3 for the x-, y- and z- directions , , : Cartesian unit vectors in the x-, y and z- directions : turbulence kinetic energy [L2 T-2] : constant of proportionality in the empirical formulations [dimensionless] : length (or thickness) of each strip [L] : actual length of the rigid body [L] : length of the sides of the sections [L]

xii

: length of the ship [L] : mass of the rigid body [M] : mass of each rigid body strip used in the simulations [M] : added mass [M] : maximum slamming impact pressure [M L-1 T-2] : Reynolds number [dimensionless] : initial drop height used in the simulation files [L] : initial experimental drop height [L] ; source term of the energy equation [M L-1 T-3] : source term of the momentum equation [M L-2 T-2] : source term of the additional variable equation [M2 L-6 T-1] : time [T] : flow temperature [] : pressure rise time [T] : flow velocity vector in Cartesian coordinates [L T -1]

: time-varying component of the velocity in Cartesian tensor notation [L T -1] : average of the product of the time-varying components of the flow velocity, in Cartesian tensor notation [L T-1] , : time-averaged components of the flow velocity in Cartesian tensor notation [L T -1]

: friction velocity [L T-1] : near-wall velocity [dimensionless] : flow speed very far from the body [L T -1] : initial flow velocity [L T-1] : forward velocity of the ship [L T -1] : Cartesian coordinates in tensor notation [L]

xiii

: measure of first node distance from the wall [dimensionless] : deadrise angle () [dimensionless] : dynamic diffusivity of an additional variable [M L-1 T-1] : distance of first node from the wall [L] : turbulence eddy dissipation rate [L2 T-3] : free-surface elevation [L] : vertical coordinate of a point on the wedge relative to the bottom of the wedge [L] : Von Karman constant [dimensionless] : thermal diffusivity [M L T -3 -1] in the energy equation; tank length [L] otherwise : dynamic viscosity [M L-1 T-1] : effective dynamic viscosity [M L-1 T-1] : dynamic eddy viscosity [M L-1 T-1] : kinematic viscosity [L2 T-1] : kinematic eddy viscosity [L2 T-1] : density [M L-3] : molecular stress tensor [M L-1 T-2], including both normal and shear stress components : wall shear stress [M L-1 T-2] : flow velocity potential [L2 T-1] : time-varying component of the additional variable [M L-3] : additional variable (non-reacting scalar), derived from the turbulence models [M L-3] : averaged component of the additional variable [M L-3] : turbulence frequency [T-1] The employed dimensions are: [L]: Length; [M]: Mass; [T]: Time; []: Temperature.

xiv

Abbreviations
2-D: two-dimensional 3-D: three-dimensional CEL: CFX Expression Language CFD: Computational Fluid Dynamics D: moulded Depth DNS: Direct Numerical Simulation FEA: Finite Element Analysis Gb: Gigabyte, 109 bytes. It is a measure of computing hardware size. ISSC: International Ship and offshore Structures Conference ISV: Increase near Small Volume ITTC: International Towing Tank Conference RANSE: Reynolds Averaged Navier-Stokes Equations RAM: Random Access Memory RMS: Root Mean Square SSG: Speziale, Sarkar and Gatski turbulence model

xv

1.0- Introduction
Slamming typically occurs in rough seas and is associated with large ship motions. A slam is represented by the substantial impact caused by the emersion of part of the hull and its immediate re-entry into the water with the impact velocity greater than a critical value (Lloyd, 1998). Common consequences are local damage from the impact load, large-scale buckling of the decks (Bertram, 2000) or a decrease in the fatigue life of the hull in the best case, as even moderate slamming results in the ship vibrating at its natural frequency, which is known as whipping (Lloyd, 1998). Hence, estimates of operational impact loads are of fundamental importance with regard to the safety of the ship design.

1.1-

Aim and Objectives

The main aim of this report is to produce a quick, economical (in terms of both computation time and money) and reliable method for estimating the pressures and loads associated with slamming events in order to produce structures of sufficient, but not excessive, strength. For this reason, firstly, a CFD simulation of a two-dimensional wedge dropping onto a calm water-surface has been performed using the commercial software ANSYS CFX, which has been validated against available experimental data (Lewis, et al., 2010). The simulation includes a study in the influence of mesh density, as suitable domain size and turbulence model were selected during a preliminary stage of the project. Due to incorrect pressure prediction and large oscillations in the results by CFX, the software was used to simulate the fall of a different wedge geometry (Aarsnes, 1996) in order to assess the source of the problems. Secondly, a similar analysis is performed for a two-dimensional ship-like (bow) section. The accuracy of the RANSE solver is assessed for two impact speeds and three heel angles. In this case, the solution of CFX can be validated not only against experimental measurements, but also against other numerical approaches (Brizzolara, et al., 2008). Thirdly, a study on the influence of tank dimensions on the accuracy of pressure and force per unit length measurements is to be performed for a fully three-dimensional case of the bow section drop in ANSYS CFX, due to the greater reliability of the software
1

with this geometry. The aim of this analysis is to test the correctness of validating twodimensional results (either numerical or analytical) against three-dimensional experimental results. In particular, a study on the influence of the distance from the longitudinal edges of the section to the tank walls, expressed as a fraction of the body length, has been performed. Finally, the validity of current empirical formulations in estimating the loads due to slamming events has been assessed against the available experimental and numerical data. These formulae (Stavovy & Chuang, 1976; Ochi & Motter, 1973) have mainly been developed during the 70s and have been adopted by classification societies, such as Lloyds Register (2011 a; 2011 b). Their usefulness lies in their ability to be integrated within wave load prediction programs, as performed by Belik, et al. (1980; 1983).

1.2-

Report Structure

The current report includes a background to the problem in section 2 and the description of the selected parameters in ANSYS CFX in section 3. Section 4 contains a summary of the analysed empirical formulations. The results are presented in sections 5 and their discussion follows in section 6. Section 7 presents the conclusions and section 8 provides comments on possible future improvements. Appendix I covers numerical approaches for slamming problems other than CFD more in detail. Appendix II

represents a summary of the theory behind CFD. Appendix III contains the data and drawings relating to the geometry files used in CFX. Appendix IV expands on section 4, namely the empirical methods applicable to slamming problems. In Appendix V,

preliminary studies on turbulence, interphase and mesh stiffness models as well as domain size are presented. Appendix VI encloses the results by Lewis, et al. (2010). Appendix VII is dedicated to the visual results obtained from CFX. Appendix VIII contains the .ccl file imported into CFX.

2.0- Literature Review


Ship responses in waves are investigated in terms of two different phenomena that can be linearly superimposed: wave excited responses, which are cyclic and continuous in time, and slamming responses, which are a series of decaying transients (Belik, et al., 1980). For the purpose of the design of structures that can withstand slamming and whipping loads, a global slamming analysis is necessary, which evaluates bending moments and shear forces on hull structures owing to fluid impact loadings (Kapsenberg & Thornwhill, 2010; ISSC, 2009), as employed by Belik, et al. (1980; 1983). Hence, the analysis of local slamming is fundamental not only in assessing local loads and associated damage, but also in estimating the impact loadings necessary for a global slamming analysis. Nevertheless, it should be noted that three-dimensional

effects can reduce the two-dimensional slamming impact force by up to 30% (Faltisen and Chezhian, 2005, cited in ISSC, 2009). Good examples of global slamming analysis have been presented by Belik, et al. (1988) and Oberhagenmann, et al. (2008). Since Von Karmans work (1929, cited in Bertram, 2000), wave impact loads have been analysed analytically, experimentally and numerically (Bertram, 2000). Despite the strong three-dimensionality of the phenomenon, most theories and numerical applications are for two-dimensional, rigid bodies for simplicity. Von Karman (1929, cited in Betram, 2000) used a momentum impact (based on momentum conservation) approach to estimate the impact slamming pressures on a two-dimensional wedge with small deadrise angle, which are underestimated. A more realistic impact theory was derived by Wagner (1932, cited in Bertram, 2000), which approximates the flow under the wedge as that around an expanding flat plate in a uniform flow. Although the small deadrise angle assumption was still present in his derivation, the theory was found to be unsuitable for due to air trapping effects (Bertram, 2000). Hence, despite being

a very useful tool to estimate the slamming pressure against the wedge, it cannot be used with the ship-like section, as it presents a 0 deadrise angle. However, it tends to slightly underestimate the peak pressures as well (Bertram, 2000). Improvements to the theory in order to account for air trapping and water compressibility have been proposed by Chuang (1967, cited in Betram, 2000) and Korobkin (1996, cited in Bertram, 2000) respectively. In addition, a detailed theoretical treatment of hydroelastic effects is presented by Faltinsen (1997, cited in Maki, et al., 2011; 2000). Nevertheless, none of
3

these has been considered due to their complexity. The two-dimensional studies by Wagner (1932, cited in Bertram, 2000) have been extended to the study of the water entry of three-dimensional cones into the water by Chuang (1969, cited in Stavovy and Chuang, 1976) with appropriate corrections (Bertram, 2000). This theory has not been used in the current individual project, although it is interesting to note that the threedimensional effects reduce the slamming impact pressures (Bertram, 2000). In the 1970s, analytical two-dimensional methods have been proposed to determine wave impact slamming pressures on any ship-like sections by Ochi and Motter (1973) and Stavovy and Chuang (1976), which use empirical formulations derived from experimental results to estimate the values of the constants of proportionality they employ. Indeed, they assume the peak slamming pressure to be proportional to the impact speed squared. The former approach considers the lower part of any ship sections (to about 10% of the draught) only, where the greatest slamming loads are encountered (Belik, 1982). In addition, an exponential time-distribution for the impact pressures and force is provided (Belik, 1982). Both these proposals have been found to be very interesting and they have been applied to the bow section by employing both procedures (Ochi & Motter, 1973; Stavovy & Chuang, 1976). In particular, Stavovy and Chuangs (1976) method, which considers the rigid body dropping onto a wavy surface at any heel angles, has been found by Beukelman (1980, cited in Belik, et al., 1980) to provide the most satisfactory agreement with measured impact pressures in global slamming analysis experiments. However, being based on the results from Chuangs (1969, cited in Stavovy and Chuang, 1976) cone theory and wedge experiments, it tends to be applicable for constant deadrise angle surfaces only. Hence, while it can be applied to the wedge straightforwardly, the bow section needs being discretised in a number of panels. Both Ochi and Motter (1973) and Stavovy and Chuangs (1976) approaches have been employed by Belik, et al. (1980; 1983; 1988) in their global slamming analysis. In addition, they considered flare slamming as well (Belik, et al., 1988) by including flare buoyancy and momentum to account for freeboard immersion. However, the impact forces due to bottom slamming only have been considered in the current project for simplicity. Furthermore, Ochi and Motters (1973) approach has been incorporated within Lloyds Rules (2011 a; 2011 b). Hence, the validity of this modified approach has also been evaluated.

Experimental measurements have been performed both on scale models in seakeeping tanks (ITTC, 2008; ISSC, 2009) as well as drop tests for a range of velocities or in freefall. The latter in particular are of interest in the current work, as they provide accurate data against which numerical results can be validated. They can be either elastic (Battley, et al., 2005; Peseux, et al., 2005, cited in ITTC, 2008) or rigid, constant or varying (e.g. a cone shape (Constantinescu, et al., 2009; Peseux, 2005, cited in ISSC, 2009) cross-section bodies or even flat plates (Huera-Huarte & Gharib, 2011). Of these, the current analysis regards rigid, two-dimensional bodies due to their simplicity, without affecting their practical importance. For this reason, tests on ship-like sections have been researched. Of these, wedges are the most common because of planing craft. A variety of deadrise angles has been used in different tests, such as those by Engle and Lewis (2003), which includes a comparison of the experimental data with the results of BEM and finite element method computations, as well as Chuangs (1967, cited in Stavovy and Chuang, 1976) empirical formulation. Tveitnes et al.s (2008) experiments consider constant velocity water entry problems and present only force data and crude pictures. Mehedi Sayeed et al.s (2010) free-fall experiment, performed with high frequency pressure gauges and accelerometers, is in particular very interesting, as the authors validate Stavovy and Chuangs (1976) empirical formulations for wedge drops with the available data. Hence, it can be used in the last part of the project.

Furthermore, experiments have been performed by Whelan (2004, cited in Lewis, 2011) on the dropping of a wedge onto a calm water surface, which also include a study on the influence of tank dimensions on the accuracy of the results. Nevertheless, for the validation of the CFD data of the first part of the project, the experiment on the free fall of a wedge by Lewis et al. (2010) has been selected instead due to the greater accuracy of the results deriving from higher frequency acquisition-data sensors and the accurate images that show the time development of the event. Furthermore, the wedge free-fall experiment performed by Aarsnes (1996) has been also analysed after problems with the numerical solutions for the Lewis, et al.s (2010) wedge have been encountered. Nevertheless, only data relating to a heeled wedge are available due to confidentiality issues. For both experiments, the data include pressure, force and acceleration

measurements. In addition, tests have also been performed on ship sections, the most influential of which are those by Yang et al. (2007) and Aarsnes (1996). The latter data was used for the validation of the CFD results of the second and third parts of the

project, as there is a recent paper which compares the experimental results with those from numerous numerical sources (Brizzolara, et al., 2008). Different numerical procedures have been used in the solution of slamming problems, which have increased in number and variety due to the recent technological improvements of the computing industry (ISSC, 2009). Most of the procedures are applied to two-dimensional drop tests on to a calm water surface, similar to the cited experiments. There are cases of more complex three-dimensional simulations of whole ship hulls (ITTC, 2008), such as that by Kapsenberg and Thornwhill (2010), who successfully use CFX, or three-dimensional drops onto a disturbed free-surface (Kim, 2005, cited in ITTC, 2008), but these have not been considered. Neither have the simulations of hydroelastic effects been taken into account, which typically use coupled or one-way transfer of information from a CFD to a FEA program (Paik, et al., 2009; Maki, et al., 2011). From a study on the influence of three-dimensional effects on the impact force on a wedge, Yang and Qiu (2010) found a decrease in the magnitude of the impact force for increasing tank dimensions. One of the least computationally-intensive techniques is the Boundary Element Method (BEM), which was firstly applied successfully by Zhao et al. (1996, cited in ISSC, 2009) to wedge slamming problems and has been widely used afterwards, although it tends to overestimate slamming pressures (Brizzolara, et al., 2008). Another numerical technique, which has become very popular recently, is the Smoothed Particle Hydrodynamics method. It is described in detail by Viviani, et al. (2007). Similarly, the Moving Particle Semi-implicit is another Lagrangian approach based on particles and their interactions (Lee, et al., 2009). Nevertheless, these procedures were not selected for the current project, as their details go beyond the knowledge of the student. For this reason, a well-established software package that presented a user friendly interphase and a good documentation was sought. One explicit element code that could have been used is LS-DYNA (2011). Although it has been used successfully for trimaran cross structures (Cao and Wu, 2007, cited in ISSC, 2009), it is not available at the University of Southampton. For the same reason, FLOW-3D (2011) and Abaqus (2011), which has been successfully been employed by Constantinescu, et al. (2009), could not be selected. Hence, the choice fell on the ANSYS packages available at the University of Southampton, which includes both CFX, a finite-volume solver, and FLUENT, a finite-element solver (ANSYS, 2009). CFX was favoured, as it had already been employed in the same two-dimensional wedge and
6

bow section simulations by Lewis (2011), showing very good agreement with the experimental measurements. Nevertheless, Lewis (2011) method, which includes fixed body with an inflow of water onto its surface, tends to be applicable for high impact speeds only. Hence, a moving rigid body approach was selected instead due to the favourable experience of Yang, et al. (2007). Indeed, Yang, et al. (2007) have proven the suitability of CFD techniques in estimating the peaks and temporal and spatial distributions of slamming pressures on two-dimensional wedge and containership stern sections against experimental measurements. In particular, both fixed and moving grid systems were employed, where the moving mesh was reconstructed at every time step, which has been found to correctly simulate the rigid body drop. A description of the cited numerical procedures is presented in Appendix I.

3.0- Methodology
Ansys CFX is a RANSE, finite-volume solver. It has been selected due to its robustness, flexibility and simple user interphase. Version 12.1 has been used in all cases. The Reynolds-Averaged Navier Stokes equations can be expressed in Cartesian tensor notation as (ANSYS, 2009): ( ( ( ) ( ) ( ) ( ) ( ) ; ) [ ( ) (1)

; )]

(2)

; (3)

(4)

In equations (1), (2), (3) and (4) the over-bars and capital letters refer to averaged components and the terms represent the Reynolds stresses (Wilson, 2011). Equation (2) is the

Equation (1) is the Reynolds averaged continuity equation.

Reynolds averaged momentum equation. Equation (3) is the Reynolds averaged energy equation. Equation (4) is the additional variable equation required for closure, where the additional variable is provided by the turbulence model. A more thorough

explanation of the fluid dynamics involved in CFX is presented in Appendix II. In the finite volume technique, the region of interest is divided into small sub-regions, called control volumes, which correspond to control surfaces around each node (ANSYS, 2009). The equations are discretized and solved iteratively for each control volume, which allows an approximation of each flow variable to be known at specific points throughout the domain (ANSYS, 2009). Hence, the value of each averaged flow parameter can be known at any time instant at any nodes within the mesh. Greater information for all parts of this section can be found in Appendix II.

3.1-

Type of Simulations

The impulsive nature of slamming requires a transient simulation. A second-order backward Euler transient scheme has been used from the very first due its greater accuracy (ANSYS, 2009), despite good engineering practice would require the use of first-order accurate schemes first (Bressloff, 2011). The total simulation time has been selected using common sense, based on the experimental results (Aarsnes, 1996; Brizzolara, et al., 2008; Lewis, et al., 2010). Various time steps have been employed for different mesh densities in order to maintain Courant number equality, as recommended (ANSYS, 2009). In particular, Courant number RMS values lower than 1 are mandatory for correct transient simulations (ANSYS, 2009), where the Courant number is proportional to the ratio of the time step and each element size, as per equation (18) in Appendix II. Therefore, finer meshes require smaller time steps. Nevertheless, for the two-dimensional wedge simulations, two time-steps have been employed for each mesh file in order to assess the effect on the Courant number and the validity of the results. The finest meshes and in particular the three-dimensional files would have required finer time steps (at least 0.0005 s), in accordance with their grid density. Nevertheless, a compromise had to be found with the time constraints of the project and the hard disk space available to the student, which should not exceed 50 Gb on the University of Southampton Network. Indeed, most simulations have been

performed through parallel processes on the University of Southampton Lyceum Cluster Unit of computers (iSolutions, 2011) for time-saving purposes, where the computations were necessarily performed on the students workspace. For this reason, the largest input files required to be split. Each part was then solved starting from the final results of the previous part.

3.2-

Geometry Selection

The section view of the Lewis, et al.s (2010) wedge, Aarsnes (1996) wedge and bow section and their domains can be seen in figures 53 to 58 in Appendix III. The

coordinates of the points used to plot the splines that represent the contour of the shiplike section can be similarly seen in Appendix III, as well as the location of the monitor points on the two wedges. It should be noted that all the bow section surfaces have been divided in two parts, where the lower part reaches up to 10% of the depth, so as to provide a means of validating the empirical procedures presented in section 4.0. The
9

height and width of the domain of the ship-like section have been taken as 10 times the basis dimensions of the geometry, from good engineering practice (Bressloff, 2011). This favoured a correct meshing of the NURBS surfaces. On the other hand, the domains of the wedge simulations have the same dimensions as the section of the tanks employed in the experiments (Aarsnes, 1996; Lewis, et al., 2010). The location of the origin does not affect the validity of the simulations results, as long as care is taken with the sign of hwat in the input CEL equations presented in appendix VIII. It should be noted that the bow section has been tilted by 9.8 and 28.3 in the heeled section simulations, in accordance with the experimental procedure (Aarsnes, 1996). The dimensions of the bodies have been taken from Lewis, et al.s (2010) and Aarsnes (1996) using full-scale for complete similarity with the experiments, with the exception of the height of the rigid bodies above the water level. In fact, employing the actual initial height would have resulted in mesh folding, as well as geometry files of excessive size requiring an enormous number of nodes. Therefore, it was decided to place the rigid bodies closer to the water surface and to impart an initial velocity equal to that the body would experience if subjected to free-fall (which can be considered a reasonable assumption by considering the experimental methodology). In the case of Aarsnes (1996) experiments, the impact velocity was in fact known. Hence, this was set as the initial velocity, with a height of the section above the water level of 5 cm so as to enable possible air trapping effects. The initial height of Lewis, et al.s (2010) wedge above the free surface has been set to 10 cm as a compromise achieved after preliminary scaling studies, which showed that greater initial drop height caused higher impact pressure magnitudes, but greater oscillations and instability in the solver and results, as described in Appendix V. Table 1 contains all initial vertical drop speeds employed for the bow section and wedge simulations. The latter have been found through the constant acceleration equations of motion ((22) and (23)) visible in Appendix III. No corrections have been made for the Reynolds number, as it was assumed that the experiments were performed in fresh water in the absence of relative information (however, the error is believed to be small even if salt water had in fact been used).

10

TABLE 1: INPUT Y-VELOCITY AND HEEL ANGLES FOR ALL SIMULATIONS. Heel angle () 0 0 20.3 0 0 9.8 9.8 28.3 Experimental initial drop height (m) 0.50 0.75 0.317 0.018 0.318 0.020 0.318 0.020 Experimental impact speed (m/s) N.A. N.A. 2.43 0.58 2.43 0.61 2.43 0.61 Initial drop speed (m/s) -1.401 -1.786 2.43 0.58 2.43 0.61 2.43 0.61 Simulation name a-d; l.s. h.s. Aarsnes wedge l.s., 3-D h.s 9.8 9.8-2.43 28.3

Rigid Body Lewis wedge Aarsnes wedge Ship-like section

Time (s) 0.285569 0.364030 -

The lateral dimensions of the tank for the three-dimensional simulations are the same as for the two-dimensional case, but the length does not match the experimental value, as it would result in mesh folding. For this reason, a study has been performed into the influence of domain dimensions on the impact pressure data and in particular on the magnitude of the three-dimensional effects. The initial tank or domain length was taken as 10 times the body length (1 m (Aarsnes, 1996)), as recommended by good CFD practice (Bressloff, 2011). The length of the domain has then been reduced to 4, 2 and 1.5 times the domain length. Beyond this point, mesh folding occurs. The names of the three-dimensional simulations have been based on the ratio, namely x10, x4, x2 and x1.5. In addition, it should be noted that only half body is modelled, as a symmetry plane is employed so as to reduce the number of nodes. No mesh is required for the bodies, as the rigid body feature of ANSYS CFX is employed (ANSYS, 2009), which allows the input of body mass, position of centre of mass (which has been assumed to coincide with the centre of volume), forces acting onto the body (gravity only) and initial downward velocity, as the motion was restrained to the y-axis only. For the two-dimensional cases, the mass of each strip is calculated as kg, where is the total mass of the experimental rig, is the length of the

experimental rigid bodies and the length of the strip. The results are reported in table 2 for all simulations.

11

TABLE 2: INPUT MASS AND LENGTH FOR ALL SIMULATIONS. Exp. mass (kg) 23.4 33.4 298.3 267.9 267.9 Exp. length (m) 0.735 0.735 1 1 1 Strip length (m) 0.01 0.01 0.01 0.02 0.5 Strip mass (kg) 0.318 0.454 2.983 5.358 133.95 Simulations a; b; c; d; l.m., l.s.; l.m., h.s. h.m., l.s.; h.m., h.s. Aarsnes wedge 2-D bow section 3-D bow section

3.3-

Mesh Generation

The mesh generation process, realised in CFX-Mesh, is based on the compromise between spacing, resolution and geometry description (Lewis, 2011). Indeed, the mesh should be fine enough to represent the selected geometry correctly and ensure acceptable levels of conservation of mass and momentum, while avoiding excessive computation time, which is typically proportional to the number of cells squared (Bressloff, 2011). The use of anisotropic meshes (ANSYS, 2009) has been adopted to solve this issue, which relies on finely spaced grids only in the areas of interest. Hence, the grid is denser near the body surface, while the rest of the mesh contains coarser elements. In addition, an inflated boundary layer has been used near the bottom and top boundaries in order to facilitate mesh deformation for all cases. However, the

simplicity of CFX-Mesh as a meshing tool meant that the first layer height for this region had to be the same as for the inflated boundary layers placed next to the rigid body surface. This was set as 0.05 m for most meshes and 0.005 m for the files with the finest grid in order to obtain acceptable values. (equation (20) in section 11.3), which

The concept of the dimensionless parameter

represents the distance of the first node from a boundary, is particularly important in CFD simulations. The recommended range is (Bressloff, 2011), as this

represents the region of the turbulent boundary layer where the log-law is applicable, as viscous and turbulent models coexist. The performed simulations do not respect these criteria despite all efforts, thus meaning the first node typically lies in the viscous sublayer as values as low as 0.9 are present. This is believed to be the cause for solver

errors with some turbulence models, as explained in section 3.4. In addition, it was checked for every time step that the expansion factor, orthogonality and aspect ratio, which are measures of mesh quality, were within the prescribed limits (ANSYS, 2009), which can be seen in table 3.
12

TABLE 3: MEASURES OF MESH QUALITY. Maximum Expansion Factor 20 Minimum Orthogonality Angle 20 Maximum Aspect Ratio 100

The mesh deformation is based on the stiffness model. Since the standard model used was inversely proportional to the volume of each element to avoid mesh folding as recommended in the ANSYS tutorials (ANSYS, 2009), coarser meshes were in fact allowed greater mesh deformation. Mesh refinement has been based on multiplying the number of elements of a basis mesh file by 2, 8 and 64, as per good CFD practice (Bressloff, 2011). Two-dimensional meshes were obtained for thin domains using the extruded 2-D meshing approach of CFX-Mesh, while three-dimensional meshes through the advancing front/Delunay method (ANSYS, 2009). Hence, two-dimensional meshes are composed of hexahedra and prisms only, while three-dimensional grids have tetrahedra and pyramids as well.

3.4-

Turbulence Models
(Homsy, et al., 2007), have been calculated for all

The Reynolds numbers,

cases in Appendix II and are reported in table 4 for convenience. Nevertheless, due to the impulsive nature of slamming, these values should be treated with caution, as explained in Appendix II.
TABLE 4: REYNOLDS NUMBERS OF ALL SIMULATIONS Low-speed wedge 338995 High-speed wedge 412408 Aarsnes wedge 722388 Low-speed section 138507-145672 High-speed section 580298

For the available

, turbulence is to be expected (Homsy, et al., 2007), which is

characterised by eddy formation and highly irregular particle paths (Wilson, 2011). CFX modifies the Navier-Stokes equations with averaged and fluctuating quantities in order to account for the turbulence effects, resulting in the formulation of RANSE. As explained in Appendix II, closure of these equations is achieved with two main models: eddy viscosity and Reynolds stress models. The former are the most common and include the k- and shear stress transport models. The k- is believed to be the most
13

robust and applicable to the majority of cases (ANSYS, 2009). The latter is more appropriate for the description of boundary layer problems for low Reynolds numbers (ANSYS, 2009) and was found by Lewis (2011) to be more accurate. The Reynolds stress SSG belongs to the second category of turbulence models and tends to be more accurate, although also computationally-intensive (ANSYS, 2009). Both k- and SSG models require the use of a scalable wall function to solve problems for low values,

which ignores the viscous sub-layer (ANSYS, 2009). In the case of the shear stress transport, an automatic wall function has been employed instead, which considers the viscous sub-layer also for low thickness errors (ANSYS, 2009). From a preliminary study (Appendix V), the selected turbulence model for all simulations has been the shear stress transport due to its greatest applicability, despite the larger number of iteration loops it needs for the same convergence criterion. Indeed, the k- and SSG models would cause overflow errors in the solver for low values, values, thus resulting in less boundary displacement

typical of the finer grids, possibly due to the near-wall treatment and scalable wall function of these models.

3.5-

Multiphase Flow

Two Eulerian, incompressible fluids have been modelled in the domain (air and water), to which an inhomogeneous multiphase model has been applied as recommended by ANSYS (2009) for simulations involving a free surface, although Lewis (2011) found the homogenous model to be more appropriate. The selected interphase model is the free surface, based on a preliminary analysis (Appendix V) and ANSYS (2009) recommendations. An interphase compression layer of two has been adopted to enable air entrapment. The hydrostatic pressure and water depth were set manually through step functions, visible in Appendix VIII.

3.6-

Boundary Conditions

The employed boundary conditions are smooth, no-slip walls on the rigid body surface, which are input to follow the rigid body solution (its downward motion), a stationary no-slip wall at the bottom of the domain, a stationary free-slip wall at the top of the domain and two symmetry planes at the front and back of the mesh, so as to simulate a
14

two-dimensional flow. For the side walls, the adoption of a fictional opening boundary condition was necessary in order to specify the hydrostatic pressure. ANSYS (2009) has been found to place free-slip walls next to it automatically. Where the slamming impact force on the lower portion of the section had to be determined, the side surfaces of the rigid body were divided in two parts in the geometry file: a lower surface up to 10% of the section depth and the remainder. Both surfaces have been matched to the same boundary condition, while the total impact force has been calculated as the sum of the loads acting against the two separate surfaces. For the three-dimensional mesh, one of the two symmetry planes is still used in order to halve the total number of nodes, while the other plane is replaced by a stationary, smooth, no-slip wall.

3.7-

Error Analysis

A thorough error analysis for CFD simulations is treated by Lewis (2011), as reported in Appendix II. Although an accurate procedure has been developed by Roache (1997, cited in Lewis, 2011) in order to study the influence of grid independence in transient simulations, this has been considered as too complex for the current knowledge of the student and a different, simplified approach has been used. The verification of the discretisation errors is mainly based on the convergence criteria set in the CFX solver. The minimum number of coefficient loops has been set to 2, while the maximum one to 25 (10 for the three-dimensional meshes due to constraints posed by the memory size), where a compromise was found between computational resources and desired accuracy. Tutorial 32 (ANSYS, 2009) employs a lower value of 5, but an increase was required in order to favour the convergence of the shear stress transport turbulence model, as mentioned in section 3.4. The minimum convergence criterion was set with a RMS value of , as recommended by ANSYS (2009). It has been always met except in

the first few time-steps. In addition, a study on the influence of grid density on the accuracy of the results was performed by employing different mesh files. From good CFD practice examples

(Bressloff, 2011), the values should converge to an asymptotic value.


15

4.0- Theoretical and Empirical Approaches


In order to treat slamming analytically within the framework of potential theory, the fluid is assumed to be inviscid, irrotational and incompressible and the surface tension and gravity effects are ignored (Bertram, 2000). Only impacts where the rigid body drops vertically are considered. Von Karmans (1929, cited in Bertram, 2000) approach is reported in Appendix IV only, as it greatly underestimates the impact pressures. A more realistic impact theory was derived by Wagner (1932, cited in Bertram, 2000), which also considers the local uprise of the water (Faltinsen, 2005). Wagner (1932, cited in Faltinsen, 2005) assumes that the flow under the wedge can be approximated as that around an expanding flat plate in a uniform flow, for small and negligible gravity effects. By employing the

velocity potential, the wave surface elevation and the half width of the wetted surface area, it is possible to obtain the maximum impact pressure as (Faltinsen, 1993): . (5)

The derivation of equation (5) is presented in Appendix IV. Wagners theory can applied to arbitrarily shaped bodies as long as the deadrise angle is neither too large nor too small ( ) due to air trapping, hydroelasticity and compressibility effects (Bertram, 2000; Faltinsen, 2005). Although Wagners (1932, cited in Bertram, 2000) method tends to underestimate the peak pressures (Bertram, 2000), it provides better estimates than Von Karmans (1929, cited in Bertram, 2000). Nevertheless, it cannot be used for the ship-like section due to its 0 deadrise angle. Ochi and Motter (1973) presented a method for estimating the peak impact pressure against any ship-like sections based on the impact speed and a dimensionless coefficient: . This assumes that the (6)

value is independent of sea severity and ship speed up to is a function of the section shape only, in

(Ochi & Motter, 1973). Hence,

particular up to 10% of the loaded draught (Ochi & Motter, 1973), since this is assumed to be the part subjected to the greatest slamming load. However, it should be noted that flare slamming may affect the results for the studied section.
16

In addition, in this

particular case the draught has been substituted with the section depth in absence of any other information. In equation (6), the non-dimensional coefficient has been fitted to experimental data and obtained through regressional analysis, with an average percentage difference of 16.2% for a sample of 15 experiments (Ochi & Motter, 1973). Although an expression for is provided in the paper by Ochi and Motter (1973) based

on the conformal mapping, three-parameter transformation parameters, the value employed within this project has been graphically interpolated from figure 1 (Ochi & Motter, 1973).

FIGURE 1: PLOTS FOR INTERPOLATING THE K 1 VALUE FOR THE LOWER PART OF THE SHIP-LIKE (BOW) SECTION, TAKEN FROM OCHI AND MOTTER (1973).

The required particulars of the analysed ship-like section up to the waterline at 10% of the moulded depth are provided in table 5. The resulting value is 5.3.

TABLE 5: MAIN PARTICULARS OF THE BOW SECTION TO A DEPTH OF 10% OF THE MOULDED DEPTH. (m) 0.0212 (m) 0.005 (m) 0.024 0.23 0.73 2p (m) 0.0705

The approach by Ochi and Motter (Ochi & Motter, 1973) is also employed by Lloyds Register (2011 a; 2011 b), a classification society, for the determination of the slamming impact pressure against the hull of naval and trimaran ships. The value, defined as the hull form shape coefficient, for equation (6) is estimated as for . Since the deadrise angle is 0, . For naval ships,

flare slamming impact pressures can also be estimated if the forward ship velocity is
17

known (Lloyds Register, 2011 a).

Similarly, for all other ships, Lloyds Register

provides estimates of the slamming peak impact pressures and loads based on the ship length, draught and forward speed, combined with a statistical approach. Assessment of the suitability of these formulae goes beyond the current project, as it would require a global slamming analysis, as discussed in section 2.0. For both Ochi and Motter (1973) and Lloyds Registers (2011 a; 2011 b) methods, the impact pressure has been assumed not to vary significantly onto the lower part of the section. Hence, the slamming peak impact force against the lower section can be approximated as the peak impact pressure times the perimeter. Stavovy and Chuangs (1976) approach for estimating the slamming impact pressure is the most widely applicable and thorough, as it allows for the wave action and planing as well. However, it can be used for constant deadrise sections only, i.e. wedges. In the current analysis, the simple case of the drop of a two-dimensional section with neither heel angle nor forward speed onto a calm water surface only is considered. Hence, the equation for the maximum impact pressure reduces to (Stavovy & Chuang, 1976): . In equation (7), the non-dimensional coefficient (7)

, fitted to experimental data and the

results from Wagners (1932, cited in Bertram, 2000) wedge equations and Chuangs (1967, cited in Bertram, 2000) cone theory is given by (Belik, 1982): For For , , ; For , ( ) . (8.2) (8.3) , can be evaluated as: (9) ; (8.1)

The peak slamming force against the wedge sides, of length .

Equations (7), (8.3) and (9) can straightforwardly be used to calculate the slamming impact pressure and force against the wedge. In addition, the method can be extended to the lower portion (about 10%, as per Ochi and Motter (1973)) of the ship-like section by discretising the surface in a number of segments (Belik, 1982). Equations (7), (8.1),

18

(8.2), (8.3) and (9) should be used as appropriate to calculate the peak impact pressures and force against this part of the ship-like section. The coordinate of the points

employed for this purpose match the location of the nodes of the two-dimensional simulation with a medium mesh density (section 5.3) in order to obtain a correct validation. They are available in Appendix III. Table 6 contains a summary of the segment lengths, deadrise angles and values calculated from equations (8.1), (8.2)

and (8.3). Segment 1-7, where point 7 is taken at a depth of 1% D, provides only a marginally different approximation of segment 1-2, but results in a very different value, as it is clear from table 6.
TABLE 6: LENGTH AND DEADRISE ANGLE OF THE SEGMENTS THE BOW SECTION HAS BEEN DISCRETISED INTO AND CORRESPONDING K 1 VALUE. Segment 1-2 Length (m) () 0.005 0 72 Segment 2-3 0.0075 32.2 2.78 Segment 3-4 0.0071 58.1 0.75 Segment 4-5 0.0077 66.1 0.57 Segment 5-6 0.0076 66.8 0.56 Segment 1-7 0.0120 13.2 15.28

The temporal distribution of the impact force and pressure can be assumed to be expressed in the form (Kawakami, et al., 1977, cited in Belik, 1982): ( ). (10)

This approximation has been derived from experiments (Belik, 1982). The pressure rise time, , is defined as the interval from the instant of impact to the time at which

the force or pressure reaches its maximum value (Belik, 1982). A reasonable value of from Belik (1982) is 0.002 s. Hence, the variation of impact force per unit length (or pressure by substituting with ) can be determined as (Belik, 1982): . (11)

The empirical slamming impact pressures can be validated against the experimental measurements for P1 for all geometries. In addition, for the ship-like section, the numerical results for the impact force per unit length acting against the lower part of the section can be employed for the validation of the empirical data, due to their reliability. On the other hand, this cannot be done for the wedge.
19

5.0- Numerical and Empirical Results


As already mentioned in section 3.2, all numerical results have been read through monitor points. For the pressure measurements, the relative pressure was calculated at the nodes closest to the specified monitor points rather than at the point itself, which does not account for atmospheric pressure. Hence, inaccuracies must be present, although the fineness of the meshes close to the bodies surfaces should have mitigated this issue. The forces against a surface are calculated by CFX as the integration over the whole area of the relative dynamic pressures at all the nodes belonging to the region. For simplicity, the horizontal and vertical force components have been calculated against the total body surface in all cases, even though this procedure includes the suction effects on the top surface. Nevertheless, these have been believed to be of minor importance, as the small heel angles employed did not allow the immersion of the top surface. In order to prove this, for the Separated file in section 5.3, both the impact force against the whole section and against the sides only have been calculated. The forces are presented per unit length for all cases in order to enable the comparison with the empirical results. Hence, this results in a different presentation of the force values as compared to the papers by Aarsnes (1996) and Brizzolara, et al. (2008), since the force components against Aarsnes wedge and bow section have been calculated for the central portion of the rigid body, which has a length of 0.1 m out of a total body length of 1 m. Whenever a number of processor-cores greater than 1 is a reported, the Lyceum Cluster Unit has been used to compute the results. Greater information on the properties of the unit can be found in (iSolutions, 2011). A processor number of 1 corresponds to a Xeon quad-core processor that runs at 3.20 GHz with 12 Gb RAM, which has been employed through direct rather remote access at the University of Southampton.

5.1-

Numerical Results for the Wedge by Lewis, et al. (2010)

Table 7 contains the data relating to the simulations performed with Lewis, et al.s (2010) wedge. The number of nodes and elements, the time step interval, the RMS Courant number, the total computation time and the measures of mesh quality are presented. The simulation time is 0.150 s for all cases. The influence of mesh density has been studied, where the mesh refinement is based on the number of elements as
20

described in section 3.3. Two time steps have been used for each of the two central meshes in order to study the influence on the Courant number. As it can be seen from figures 2 to 4, the files in this section have been named using letters starting from a, the coarsest mesh, and finishing at d, the finest mesh. In addition, the name may be followed by the time step interval in seconds. The simulations have been performed with 4 processor-cores.
TABLE 7: NUMBER OF NODES AND ELEMENTS (HEXA. ARE HEXAHEDRA), MESH DENSITY, TIME STEP INTERVALS, ROOT MEAN SQUARE VALUE OF THE COURANT NUMBER, TOTAL WALL-CLOCK TIME, MINIMUM ORTHOGONALITY ANGLE, MAXIMUM EXPANSION FACTOR AND ASPECT RATIO FOR SIMULATIONS ON THE WEDGE BY LEWIS, ET AL. (2010) FOR CONSTANT INPUT SPEED AND MASS. Item a b No. nodes 2334 4174 No. prisms 1263 2262 No. hexa. 465 840 Tot. elements 1728 3102 Mesh dens. basis x2 Time step (s) 0.005 0.0025 0.001 c d 20632 132622 15744 118406 2180 6340 17924 124746 x8 x64 0.0025 0.001 0.001 RMS C.N. 0.13 0.15 0.11 0.24 0.09 0.09 T.W.C.t. (s) 425 514 620 4828 4862 34495 min. O.A. 37.9 38.1 38.3 42.7 42.7 37.2 Max. E.F. 16 15 15 6 6 7 Max. A.R. 28 22 20 9 9 5

Figures 2 to 4 present the pressure variation with time at P1 and P4, P2 and P5 and P3 and P6 respectively. The plots of the second monitor point tend to be obscured by the oscillations for all figures. The convergence criteria presented an oscillatory behaviour as well, but not as pronounced. measurements are available. No force data is presented, as no experimental

21

30000 25000 20000 Pressure (Pa) 15000 10000 5000 0 0 -5000 0.01 0.02 0.03 Time (s) 0.04 0.05 0.06
P1-a P4-a P1-b_0.005 P4-b_0.005 P1-b_0.0025 P4-b_0.0025 P1-c_0.0025 P4-c_0.0025 P1-c_0.001 P4-c_0.001 P1-d P4-d P1-Experiment P4-Experiment

FIGURE 2: PRESSURE AGAINST TIME AT P1 AND P4 FOR A STUDY IN MESH DENSITY ON THE WEDGE BY LEWIS, ET AL. (2010) FOR LOW MASS AND IMPACT SPEED.

30000 25000 20000 Pressure (Pa) 15000 10000 5000 0 0 -5000 0.02 0.04 Time (s) 0.06 0.08
P2-a P5-a P2-b_0.005 P5-b_0.005 P2-b_0.0025 P5-b_0.0025 P2-c_0.0025 P5-c_0.0025 P2-c_0.001 P5-c_0.001 P2-d P5-d P2-Experiment P5-Experiment

FIGURE 3: PRESSURE AGAINST TIME AT P2 AND P5 FOR A STUDY IN MESH DENSITY ON THE WEDGE BY LEWIS, ET AL. (2010) FOR LOW MASS AND IMPACT SPEED.

22

30000 25000 20000 Pressure (Pa) 15000 10000 5000 0 0 -5000 0.02 0.04 Time (s) 0.06 0.08
P3-a P6-a P3-b_0.005 P6-b_0.005 P3-b_0.0025 P6-b_0.0025 P3-c_0.0025 P6-c_0.0025 P3-c_0.001 P6-c_0.001 P3-d P6-d P3-Experiment P6-Experiment

FIGURE 4: PRESSURE AGAINST TIME AT P3 AND P6 FOR A STUDY IN MESH DENSITY ON THE WEDGE BY LEWIS, ET AL. (2010) FOR LOW MASS AND IMPACT SPEED.

A further study in the effect of variations in mass and initial velocity on the pressure distribution is presented in figures 5 to 7. The data relative to each simulation is contained in table 8, where a time step of 0.0025 s has been used in all cases. The initial speed and strip mass have been calculated in section 3.2. The low mass, low

impact speed file coincides with file c_0.0025 in the previous figures and the mesh is shared by all other simulations. The zero time coincides with the zero time of the simulation in this case (see section 3.2), rather than the time of impact. No

experimental measurements are plotted, but the experimental results (Lewis, et al., 2010) can be seen for all cases in Appendix VI.
TABLE 8: INPUT VELOCITY, MASS, SIMULATION TIME, ROOT MEAN SQUARE VALUE OF THE COURANT NUMBER, TOTAL WALL-CLOCK TIME, MINIMUM ORTHOGONALITY ANGLE, MAXIMUM EXPANSION FACTOR AND ASPECT RATIO FOR SIMULATIONS ON THE WEDGE BY LEWIS, ET AL. (2010) FOR CONSTANT MESH DENSITY. L. STANDS FOR LOW, H. FOR HIGH, M. FOR MASS AND S. FOR INPUT SPEED. Item l.m., l.s. l.m., h.s. h.m., l.s. h.m., h.s. In. Vel (m/s) -1.401 -1.786 -1.401 -1.786 Strip mass (kg) 0.318 0.318 0.454 0.454 S. T. (s) 0.15 0.13 0.135 0.135 RMS C.N. 0.36 0.28 0.09 0.13 T.W.C.t. (s) 4471 4401 4039 3428 min. O.A. 42.8 42.6 42.6 42.5 Max. E.F. 6 6 6 6 Max. A.R. 9 9 9 9

23

25000

20000 P1-l.m., l.s. Pressure (Pa) 15000 P4-l.m.,l.s. P1-l.m., h.s. P4-l.m., h.s. 10000 P1-h.m., l.s. P4-h.m., l.s. 5000 P1-h.m., h.s. P4-h.m., h.s. 0 0.05 0.06 0.07 0.08 0.09 0.1 0.11 0.12 Time (s)

FIGURE 5: PRESSURE AGAINST TIME AT P1 AND P4 FOR THE WEDGE BY LEWIS, ET AL. (2010) FOR VARYING MASS AND IMPACT SPEED AND CONSTANT MESH DENSITY.

16000 14000 12000 Pressure (Pa) 10000 8000 6000 4000 2000 0 0.05 0.07 0.09 Time (s) 0.11 0.13

P2-l.m., l.s. P5-l.m., l.s. P2-l.m., h.s. P5-l.m., h.s. P2-h.m., l.s. P5-h.m., l.s. P2-h.m., h.s. P5-h.m., h.s.

FIGURE 6: PRESSURE AGAINST TIME AT P2 AND P5 FOR THE WEDGE BY LEWIS, ET AL. (2010) FOR VARYING MASS AND IMPACT SPEED AND CONSTANT MESH DENSITY.

24

12000

10000 P3-l.m., l.s. P6-l.m., l.s. P3-l.m., h.s. P6-l.m., h.s. P3-h.m., l.s. 4000 P6-h.m., l.s. P3-h.m., h.s. 2000 P6-h.m., h.s.

8000 Pressure (Pa)

6000

0 0.05 0.08 Time (s) 0.11 0.14

FIGURE 7: PRESSURE AGAINST TIME AT P3 AND P6 FOR THE WEDGE BY LEWIS, ET AL. (2010) FOR VARYING MASS AND IMPACT SPEED AND CONSTANT MESH DENSITY.

5.2-

Numerical Results for Aarsnes (1996) Wedge

Table 9 shows the number of nodes and elements, the time step interval, the RMS Courant number, the total computation time, the minimum mesh displacement and the measures of mesh quality for Aarsnes (1996) wedge. The total simulation time is set to 0.05 s. The simulations have been performed by the student through direct access. Figures 8 to 10 present the numerical and experimental temporal distribution of pressure at P1, P3 and P5 respectively. No experimental measurements were available for P2, P4 and P6. In figures 11 and 12, the variation with time of the horizontal and vertical force components per unit length can be seen.
TABLE 9: NUMBER OF NODES AND ELEMENTS (HEXA. ARE HEXAHEDRA), MESH DENSITY, TIME STEP, ROOT MEAN SQUARE VALUE OF THE COURANT NUMBER, TOTAL WALL-CLOCK TIME, MINIMUM MESH DISPLACEMENT, MINIMUM ORTHOGONALITY ANGLE, MAXIMUM EXPANSION FACTOR AND ASPECT RATIO FOR SIMULATIONS ON THE WEDGE BY AARSNES (1996). No. nodes 39268 No. prisms 34802 No. hexa. 1910 Tot. elements 36712 Time step (s) 0.0005 RMS C.N. 0.12 T.W.C.t. (s) 13257 min. M.D. -0.00132 min. O. A. () 51.3 Max. E. F. 7 Max. A. R. 4

25

90000 80000 70000 60000 Pressure (Pa) 50000 40000 30000 20000 10000 0 0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 Time (s) CFX Experimental

FIGURE 8: PRESSURE VERSUS TIME AT P1 FOR THE WEDGE BY AARSNES (1996).

160000 140000 120000 Pressure (Pa) 100000 80000 60000 40000 20000 0 0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 Time (s)

CFX Experimental

FIGURE 9: PRESSURE VERSUS TIME AT P3 FOR THE WEDGE BY AARSNES (1996).

26

160000 140000 120000 Pressure (Pa) 100000 80000 60000 40000 20000 0 0 0.002 0.004 0.006 0.008 0.01 0.012 0.014 Time (s)

CFX Experimental

FIGURE 10: PRESSURE VERSUS TIME AT P5 FOR THE WEDGE BY AARSNES (1996).

3000 2500 x-Force per unit length (N/m) 2000 1500 1000 500 0 0 -500 -1000 0.005 0.01 0.015 0.02 CFX Experiment

Time (s)

FIGURE 11: VARIATION OF THE HORIZONTAL FORCE PER UNIT LENGTH WITH TIME FOR AARSNES (1996) WEDGE.

27

9000 8000 7000 y-Force per unit length (N/m) 6000 5000 4000 3000 2000 1000 0 0 -1000 0.004 0.008 Time (s) 0.012 0.016 CFX Experimental

FIGURE 12: VARIATION OF THE VERTICAL FORCE PER UNIT LENGTH WITH TIME FOR AARSNES (1996) WEDGE.

5.3-

Numerical Results for the Ship-like Section

Table 10 presents the different mesh densities, time steps, RMS Courant numbers, total computation and simulation times, minimum mesh displacement, minimum orthogonality angle, maximum expansion factor and maximum aspect ratio for each of the files employed in the simulation of the ship-like section with no heel angle. The measures of mesh and time step quality should be compared against the prescribed limits (ANSYS, 2009), available in sections 3.3 and 3.1. The mesh refinement,

described in section 3.3, has been performed for the same low input y-velocity. The High speed file uses the same grid as the low impact speed simulation with a medium mesh density, but a greater input speed as per table 1 in section 3.2. The Separated file, which has a number of elements similar to the Medium mesh, presents calculations of the vertical force acting against the whole section (as per all other cases) as well as against the sides only (excluding the top surface).

28

TABLE 10: NUMBER OF NODES AND ELEMENTS (HEXA. ARE HEXAHEDRA), MESH DENSITY, TIME STEPS, ROOT MEAN SQUARE VALUE OF THE COURANT NUMBER, SIMULATION TIME, TOTAL WALL-CLOCK TIME, NUMBER OF PROCESSOR-CORES, MINIMUM MESH DISPLACEMENT, MINIMUM ORTHOGONALITY ANGLE, MAXIMUM EXPANSION FACTOR AND ASPECT RATIO FOR SIMULATIONS ON THE BOW SECTION WITH NO HEEL ANGLE. Item Coarse Medium Fine Separated High sp. No. nodes 15120 27000 104666 31614 27000 No. prisms 12128 22566 97044 27530 22566 No. hexa. 1310 1950 3230 1750 1950 Tot. elements 13438 24516 100274 29280 24516 Mesh dens. basis x2 x8 x2 x2 Time step (s) 0.0025 0.001 0.001 0.001 0.001 RMS C.N. 0.43 0.13 0.08 0.09 0.18 S.t. (s) 0.175 0.175 0.175 0.175 0.1 T.W.C. t. (s) 4533 8952 26956 9032 5633 No. Proc. 4 8 16 8 8 min. M.D. -0.00488 -0.00195 -0.00195 -0.00195 -0.00285 min. O. A. () 50.3 49.7 45.9 50.6 49.4 Max. E. F. 7 6 5 5 6 Max. A. R. 10 6 8 6 6

Since all simulations in this section have been performed on the Lyceum Cluster Unit at the University of Southampton (iSolutions, 2011), all processor-cores have the same characteristics. Hence, it is possible to estimate the increase in the computational resources as a function of total number of elements. This is done by multiplying the total wall clock time by the number of processor-cores for the first three meshes in table 10 in order to calculate the total computation time a single processor-core would require. The resulting values are plotted against the total number of elements to produce a linear relationship, fitted with a trendline. This is shown in figure 13.
500 Total Computation Time (s) x10-3 450 400 350 300 250 200 150 100 50 0 0 20 40 60 80 100 120 Total No. of Elements x10-3 y = 4.7542x - 45.375

FIGURE 13: TOTAL COMPUTATION TIME AS A FUNCTION OF THE NUMBER OF ELEMENTS FOR THE BOW SECTION SIMULATIONS AT LOW IMPACT SPEED.

In addition, simulations have been performed with the bow section tilted by 9.8 and 28.3. Two speeds have been used in the former case; where the input vertical velocity
29

data can be seen in table 1 in section 3.2. Table 11 contains the data of the mesh density, simulation time, RMS Courant numbers, total computation time and mesh quality parameters for the tilted bow section simulations. Only time steps of 0.001 s have been used. The two 9.8 files use the same mesh.
TABLE 11: NUMBER OF NODES AND ELEMENTS (HEXA. ARE HEXAHEDRA), MESH DENSITY, SIMULATION TIME, ROOT MEAN SQUARE VALUE OF THE COURANT NUMBER, TOTAL WALL-CLOCK TIME, NUMBER OF PROCESSOR-CORES, MINIMUM MESH DISPLACEMENT, MINIMUM ORTHOGONALITY ANGLE, MAXIMUM EXPANSION FACTOR AND ASPECT RATIO FOR SIMULATIONS ON THE BOW SECTION WITH HEEL ANGLES. No. nodes 19764 20158 19764 No. Prisms 13419 13659 13419 No. hexa. 2830 2900 2830 Tot. elem. 16249 16559 16249 RMS C.N. 0.05 0.05 0.07 T.W.T. (s) 3253 5152 4399 No. proc. 8 8 1 min. M.D. -0.00197 -0.00176 -0.00281 min. O. A. () 48.6 49.6 48.5 Max. E. F. 7 9 7 Max. A. R. 6 6 6

Item 9.8 28.3 9.82.43

S.T. (s) 0.175 0.150 0.080

Figures 14 to 17 show the change in relative pressure with time at points P1, P2, P3 and P4 for the 0 heel angle, low velocity simulations, for different mesh densities. The experimental measurements (Aarsnes, 1996) are also presented for validation of the numerical results.
8000 7000 6000 5000 Pressure (Pa) 4000 3000 2000 1000 0 0 -1000 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 Coarse Medium Fine Experiment

Time (s)

FIGURE 14: PRESSURE AT P1 AGAINST TIME FOR THE SHIP-LIKE (BOW) SECTION WITH NO HEEL ANGLE AND LOW IMPACT SPEED FOR THREE MESH DENSITIES.

30

7000 6000 5000 4000 Coarse 3000 2000 1000 0 0 -1000 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 Medium Fine Experiment

Pressure (Pa)

Time (s)

FIGURE 15: PRESSURE AT P2 AGAINST TIME FOR THE BOW SECTION WITH NO HEEL ANGLE AND LOW IMPACT SPEED FOR THREE MESH DENSITIES.

7000 6000 5000 4000 Coarse 3000 2000 1000 0 0 -1000 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 Medium Fine Experiment

Pressure (Pa)

Time (s)

FIGURE 16: PRESSURE AT P3 AGAINST TIME FOR THE BOW SECTION WITH NO HEEL ANGLE AND LOW IMPACT SPEED FOR THREE MESH DENSITIES.

31

7000 6000 5000 Pressure (Pa) 4000 3000 2000 1000 0 0 -1000 0.05 0.1 Time (s) 0.15

Coarse Medium Fine Experiment

FIGURE 17: PRESSURE AT P4 AGAINST TIME FOR THE BOW SECTION WITH NO HEEL ANGLE AND LOW IMPACT SPEED FOR THREE MESH DENSITIES.

Figures 18 and 19 present the horizontal and vertical forces per unit length against time for the straight, low-speed bow section. In figure 18, the CFX curve corresponds to the simulation with a mesh of medium density.
600 400 x_Force per unit length (N/m) 200 0 0 -200 -400 -600 -800 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 CFX Experimental

Time (s)

FIGURE 18: HORIZONTAL FORCE PER UNIT LENGTH AGAINST TIME FOR THE BOW SECTION WITH NO HEEL ANGLE AND LOW IMPACT SPEED.

32

2500

2000 y-Force per unit length (N/m)

1500

Coarse Medium

1000

Fine Separated-with top

500

Separated-no top Experiment

0 0 -500 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 Time (s)

FIGURE 19: VERTICAL FORCE PER UNIT LENGTH AGAINST TIME FOR THE BOW SECTION WITH NO HEEL ANGLE AND LOW IMPACT SPEED.

Figures 20 and 21 show the relative pressures at P1 and P3 against time, for the straight, high impact speed ship-like section. Figures 22 and 23 present its horizontal and vertical force components per unit length.
20000 18000 16000 14000 Pressure (Pa) 12000 10000 8000 6000 4000 2000 0 0 0.02 0.04 0.06 Time (s) 0.08 0.1 0.12 CFX Experiment

FIGURE 20: VARIATION OF THE PRESSURE AT P1 WITH TIME FOR THE BOW SECTION WITH NO HEEL ANGLE AND HIGH IMPACT SPEED.

33

16000 14000 12000 10000 Pressure (Pa) 8000 CFX 6000 4000 2000 0 0 -2000 0.02 0.04 0.06 Time (s) 0.08 0.1 0.12 Experiment

FIGURE 21: VARIATION OF THE PRESSURE AT P3 WITH TIME FOR THE BOW SECTION WITH NO HEEL ANGLE AND HIGH IMPACT SPEED.

400

200 x-Force per unit length (N/m)

0 0 -200 0.02 0.04 0.06 0.08 0.1 0.12 CFX Experimental -400

-600

-800

Time (s)

FIGURE 22: VARIATION OF THE HORIZONTAL FORCE PER UNIT LENGTH WITHTIME FOR THE BOW SECTION WITH NO HEEL ANGLE AND HIGH IMPACT SPEED.

34

5000 4000 y-Force per unit length (N/m) 3000 2000 1000 0 0 -1000 -2000 -3000 0.02 0.04 0.06 0.08 0.1 0.12 CFX Experimental

Time (s)

FIGURE 23: VARIATION OF THE VERTICAL FORCE PER UNIT LENGTH WITH TIME FOR THE BOW SECTION WITH NO HEEL ANGLE AND HIGH IMPACT SPEED.

In figures 24 to 35, the relative pressures with time at P1, P2, P3 and P4 can be seen for a heel angle of 9.8 and low impact speed, a tilt angle of 28.3 and a roll angle of 9.8 and a high impact speed respectively.
6000 5000 4000 Pressure (Pa) 3000 CFX 2000 1000 0 0 -1000 0.02 0.04 0.06 0.08 Time (s) 0.1 0.12 0.14 0.16 Experiment

FIGURE 24: VARIATION OF THE PRESSURE AT P1 WITH TIME FOR THE BOW SECTION WITH A 9.8 HEEL ANGLE AND LOW IMPACT SPEED.

35

8000 7000 6000 5000 Pressure (Pa) 4000 CFX 3000 2000 1000 0 0 -1000 0.02 0.04 0.06 0.08 Time (s) 0.1 0.12 0.14 0.16 Experiment

FIGURE 25: VARIATION OF THE PRESSURE AT P2 WITH TIME FOR THE BOW SECTION WITH A 9.8 HEEL ANGLE AND LOW IMPACT SPEED.

9000 8000 7000 6000 Pressure (Pa) 5000 4000 3000 2000 1000 0 0 -1000 0.02 0.04 0.06 0.08 Time (s) 0.1 0.12 0.14 0.16 CFX Experiment

FIGURE 26: VARIATION OF THE PRESSURE AT P3 WITH TIME FOR THE BOW SECTION WITH A 9.8 HEEL ANGLE AND LOW IMPACT SPEED.

36

12000 10000 8000 Pressure (Pa) 6000 CFX 4000 2000 0 0 -2000 0.02 0.04 0.06 0.08 Time (s) 0.1 0.12 0.14 0.16 Experiment

FIGURE 27: VARIATION OF THE PRESSURE AT P4 WITH TIME FOR THE BOW SECTION WITH A 9.8 HEEL ANGLE AND LOW IMPACT SPEED.

3500 3000 2500 Pressure (Pa) 2000 1500 1000 500 0 0 0.02 0.04 Time (s) 0.06 0.08 0.1 CFX Experiment

FIGURE 28: VARIATION OF THE PRESSURE AT P1 WITH TIME FOR THE BOW SECTION WITH A 28.3 HEEL ANGLE.

37

14000 12000 10000 Pressure (Pa) 8000 6000 4000 2000 0 0 0.02 0.04 0.06 Time (s) 0.08 0.1 0.12 CFX Experiment

FIGURE 29: VARIATION OF THE PRESSURE AT P2 WITH TIME FOR THE BOW SECTION WITH A 28.3 HEEL ANGLE.

20000 18000 16000 14000 Pressure (Pa) 12000 10000 8000 6000 4000 2000 0 0 0.02 0.04 0.06 Time (s) 0.08 0.1 0.12 CFX Experiment

FIGURE 30: VARIATION OF THE PRESSURE AT P3 WITH TIME FOR THE BOW SECTION WITH A 28.3 HEEL ANGLE.

38

30000

25000

20000 Pressure (Pa)

15000

CFX Experiment

10000

5000

0 0 0.02 0.04 0.06 Time (s) 0.08 0.1 0.12

FIGURE 31: VARIATION OF THE PRESSURE AT P4 WITH TIME FOR THE BOW SECTION WITH A 28.3 HEEL ANGLE.

20000 18000 16000 14000 Pressure (Pa) 12000 10000 8000 6000 4000 2000 0 0 0.02 0.04 Time (s) 0.06 0.08 CFX Experiment

FIGURE 32: VARIATION OF THE PRESSURE AT P1 WITH TIME FOR THE BOW SECTION WITH A 9.8 HEEL ANGLE AND HIGH IMPACT SPEED.

39

20000 18000 16000 14000 12000 Pressure (Pa) 10000 8000 6000 4000 2000 0 -2000 0 0.02 0.04 Time (s) 0.06 0.08 CFX Experiment

FIGURE 33: VARIATION OF THE PRESSURE AT P2 WITH TIME FOR THE BOW SECTION WITH A 9.8 HEEL ANGLE AND HIGH IMPACT SPEED.

25000

20000

15000 Pressure (Pa)

10000

CFX Experiment

5000

0 0 -5000 0.02 0.04 0.06 0.08

Time (s)

FIGURE 34: VARIATION OF THE PRESSURE AT P3 WITH TIME FOR THE BOW SECTION WITH A 9.8 HEEL ANGLE AND HIGH IMPACT SPEED.

40

35000 30000 25000 Pressure (Pa) 20000 15000 10000 5000 0 0 -5000 0.02 0.04 Time (s) 0.06 0.08 CFX Experiment

FIGURE 35: VARIATION OF THE PRESSURE AT P4 WITH TIME FOR THE BOW SECTION WITH A 9.8 HEEL ANGLE AND HIGH IMPACT SPEED.

Figures 36 to 41 present the horizontal and vertical force components per unit length for a heel angle of 9.8 and low impact speed, a roll angle of 28.3 and a tilt angle of 9.8 and a high impact speed respectively.
1000 800 x-Force per unit length (N/m) 600 400 200 0 0 -200 -400 -600 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 CFX Experiment

Time (s)

FIGURE 36: HORIZONTAL FORCE PER UNIT LENGTH AGAINST TIME FOR THE BOW SECTION WITH A 9.8 HEEL ANGLE AND LOW IMPACT SPEED.

41

2000

1500 y-Force per unit length (m)

1000 CFX 500 Experimental

0 0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18

-500

Time (s)

FIGURE 37: VERTICAL FORCE PER UNIT LENGTH AGAINST TIME FOR THE BOW SECTION WITH A 9.8 HEEL ANGLE AND LOW IMPACT SPEED.

2000

1500 x-Force per unit length (N/m)

1000 CFX 500 Experiment

0 0 0.03 0.06 0.09 0.12 0.15 0.18

-500

Time (s)

FIGURE 38: HORIZONTAL FORCE PER UNIT LENGTH AGAINST TIME FOR THE BOW SECTION WITH A 28.3 HEEL ANGLE.

42

3500 3000 y-Force per unit length (N/m) 2500 2000 1500 1000 500 0 0 -500 0.02 0.04 0.06 0.08 Time (s) 0.1 0.12 0.14 0.16 CFX Experiment

FIGURE 39: VERTICAL FORCE PER UNIT LENGTH AGAINST TIME FOR THE BOW SECTION WITH A 28.3 HEEL ANGLE.

2500

2000 x-Force per unit length (N/m)

1500

1000

CFX Experiment

500

0 0 -500 0.02 0.04 0.06 0.08

Time (s)

FIGURE 40: HORIZONTAL FORCE PER UNIT LENGTH AGAINST TIME FOR THE BOW SECTION WITH A 9.8 HEEL ANGLE AND HIGH IMPACT SPEED.

43

4500 4000 3500 y-Force per unit length (N/m) 3000 2500 2000 1500 1000 500 0 0 -500 0.02 0.04 Time (s) 0.06 0.08 CFX Experiment

FIGURE 41: VERTICAL FORCE PER UNIT LENGTH AGAINST TIME FOR THE BOW SECTION WITH A 9.8 HEEL ANGLE AND HIGH IMPACT SPEED.

5.4-

Numerical Results for the Three-dimensional Ship-like Section

Table 12 presents the number of elements, processors, the RMS Courant number, the total computation time and the mesh quality parameters for the three-dimensional simulations. The names of the files are based on the ratio of the tank and the rigid body lengths, as described in section 3.2. The total simulation time is 0.175 s.
TABLE 12: NUMBER OF NODES AND ELEMENTS (PYR. ARE PYRAMIDS), NUMBER OF PROCESSOR-CORES, ROOT MEAN SQUARE VALUE OF THE COURANT NUMBER, TOTAL WALL-CLOCK TIME, MINIMUM MESH DISPLACEMENT, MINIMUM ORTHOGONALITY ANGLE, MAXIMUM EXPANSION FACTOR AND ASPECT RATIO FOR THE THREE-DIMENSIONAL SIMULATIONS OF THE BOW SECTION. Item x10 x4 x2 x1.5 No. pyr. 990 1025 1028 1011 No. prisms 49052 46203 45111 44953 Tot. elements 921604 960954 935641 952205 No. Proc. 8 1 1 16 RMS C.N. 0.3 0.31 0.31 0.31 T.W.C.t. (s) 79042 117560 102617 123577 min. M.D. -0.00203 -0.00202 -0.00201 -0.00201 min. O. A. () 14.8 18.6 20.1 11.9 Max. E. F. 221 312 377 241 Max. A. R. 32 17 17 23

Figures 42 to 45 present the relative pressure results at P1, P2, P3 and P4 for the threedimensional simulation of the bow section drop. The results are presented against the
44

experimental measurements (Aarsnes, 1996) and the numerical results for the twodimensional simulations, which correspond to a mesh of medium density (section 5.3).
7000 6000 5000 4000 3000 2000 1000 0 0 -1000 0.03 0.06 0.09 Time (s) 0.12 0.15 0.18 x10 x4 x2 x1.5 2-D Experiment

FIGURE 42: PRESSURE VARIATION WITH TIME AT P1 FOR THE BOW SECTION WITH NO HEEL ANGLE AND LOW IMPACT SPEED FROM THREE-DIMENSIONAL SIMULATIONS.

Pressure (Pa)

7000 6000 5000 4000 3000 2000 1000 0 0 -1000 0.03 0.06 0.09 Time (s) 0.12 0.15 0.18 x10 x4 x2 x1.5 2-D Experiment

FIGURE 43: PRESSURE VARIATION WITH TIME AT P2 FOR THE BOW SECTION WITH NO HEEL ANGLE AND LOW IMPACT SPEED FROM THREE-DIMENSIONAL SIMULATIONS.

Pressure (Pa)

45

7000 6000 5000 4000 3000 2000 1000 0 0 -1000 0.03 0.06 0.09 Time (s) 0.12 0.15 0.18 x10 x4 x2 x1.5 2-D Experiment

FIGURE 44: PRESSURE VARIATION WITH TIME AT P3 FOR THE BOW SECTION WITH NO HEEL ANGLE AND LOW IMPACT SPEED FROM THREE-DIMENSIONAL SIMULATIONS.

Pressure (Pa)

7000 6000 5000 4000 3000 2000 1000 0 0 -1000 0.03 0.06 0.09 Time (s) 0.12 0.15 0.18 x10 x4 x2 x1.5 2-D Experiment

FIGURE 45: PRESSURE VARIATION WITH TIME AT P4 FOR THE BOW SECTION WITH NO HEEL ANGLE AND LOW IMPACT SPEED FROM THREE-DIMENSIONAL SIMULATIONS.

Pressure (Pa)

46

In figure 46, the slamming vertical force per unit length can be seen for the experimental data and the two-dimensional and three-dimensional CFX results.
2500

2000 y-Force per unit length (N/m)

1500

x10 x4

1000

x2 x1.5

500

2D Experiment

0 0 -500 0.03 0.06 0.09 0.12 0.15 0.18

Time (s)

FIGURE 46: VERTICAL FORCE PER UNIT LENGTH AGAINST TIME FOR THE BOW SECTION WITH NO HEEL ANGLE AND LOW IMPACT SPEED FROM THREE-DIMENSIONAL SIMULATIONS.

5.5-

Empirical Results

For all cases, the peak impact pressure and force per unit length are presented in table 13. These have been calculated according to the equations and procedures described in section 4.0. The temporal distribution of all empirical peak impact pressures and loads are calculated through equations (10) and (11) in section 4.0.
TABLE 13: MAXIMUM IMPACT PRESSURE AND FORCE PER UNIT LENGTH FOR THE WEDGE AND BOW SECTION GEOMETRIES FROM ALL EMPIRICAL METHODS PRESENTED IN SECTION 4.0. WED. STANDS FOR WEDGE, SEC. FOR BOW SECTION, L. FOR LOW, H. FOR HIGH AND S. FOR SPEED. Method Wagner Stav.-Chuang Mod.Sta.-Ch. Ochi-Motter Lloyds P1-wed., l. s. (Pa) 13943 11660 N.A. N.A. P1-wed.,h. s. (Pa) 20636 17258 N.A. N.A. P1-sec., l. s. (Pa) N.A. 24148 5125 889 4695 P1-sec., h. s. (Pa) N.A. 423877 89965 15601 82421 F-sec., l. s. (N/m) N.A. 265 63 331 F-sec., h. s. (N/m) N.A. 4649 1100 5811

47

For the wedge, the temporal distribution of the slamming impact pressure has been plotted against the pressure variation with time at P1 for the experimental measurements (Lewis, et al., 2010) and the CFX solution for a finely dense mesh with the upper RMS Courant number (file c_0.0025, as per section 5.1). The graphs are shown in figures 47 and 48 for a low (1.57 m/s) and a high impact speed (1.91 m/s) respectively, where the impact speed values have been calculated through the free-fall equations ((22) and (23)) shown in Appendix III with
30000 25000 20000 15000 Wagner 10000 5000 0 0 -5000 -10000 0.01 0.02 0.03 0.04 0.05 0.06 Stavovy-Chuang Experiment CFX

m.

Pressure (Pa)

Time (s)

FIGURE 47: IMPACT PRESSURE AGAINST TIME FROM EMPIRICAL, NUMERICAL AND EXPERIMENTAL DATA FOR THE WEDGE WITH LOW IMPACT SPEED.

48

40000 35000 30000 25000 Pressure (Pa) 20000 15000 10000 5000 0 0 -5000 0.01 0.02 0.03 Time (s) 0.04 0.05 0.06 Wagner Stavovy-Chuang Experiment CFX

FIGURE 48: IMPACT PRESSURE AGAINST TIME FROM EMPIRICAL, NUMERICAL AND EXPERIMENTAL DATA FOR THE WEDGE WITH HIGH IMPACT SPEED.

For the ship-like section, figures 49 and 50 present the variation of the peak impact pressure calculated with the empirical formulae presented in section 4.0 for the low and high vertical impact velocities respectively. The graphs also contain the plots of the experimental and numerical (medium mesh) relative pressures at P1 for validation. In figure 50, the results from Stavovy and Chuangs (1976) method are not reported due to their exaggerated magnitude, whose peak can be seen in table 13. The Modified

Stavovy-Chuang plot employs segment 1-7 rather than 1-2, as described in section 4.

49

30000 25000 20000 Pressure (Pa) 15000 10000 5000 0 0 -5000 0.05 0.1 Time (s) 0.15 Medium Experiment Stavovy-Chuang Ochi-Motter Lloyds Register Modified Stavovy-Chuang

FIGURE 49: IMPACT PRESSURE AGAINST TIME FROM EMPIRICAL, NUMERICAL AND EXPERIMENTAL DATA FOR THE BOW SECTION WITH LOW IMPACT SPEED.

100000 90000 80000 70000 Pressure (Pa) 60000 50000 40000 30000 20000 10000 0 0 0.02 0.04 0.06 Time (s) 0.08 0.1 0.12 Modified StavovyChuang Lloyds Register Ochi-Motter CFX

Experiment

FIGURE 50: IMPACT PRESSURE AGAINST TIME FROM EMPIRICAL, NUMERICAL AND EXPERIMENTAL DATA FOR THE BOW SECTION WITH HIGH IMPACT SPEED.

Figures 51 and 52 present the variation with time of the numerical and empirical vertical force per unit length against the lower portion of the section only (up to 10% depth) for the low and high impact velocities respectively.
50

350 300 y-Force per unit length (N/m) 250 Coarse 200 150 100 50 0 0 -50 0.05 0.1 Time (s) 0.15 Medium Fine Separated Stavovy-Chuang Ochi-Motter Lloyds Register

FIGURE 51: VERTICAL IMPACT FORCE PER UNIT LENGTH AGAINST TIME FROM EMPIRICAL, NUMERICAL AND EXPERIMENTAL DATA FOR THE LOWER PORTION OF THE BOW SECTION WITH LOW IMPACT SPEED.

7000 6000 y-Force per unit length (N/m) 5000 4000 3000 2000 1000 0 0 0.02 0.04 0.06 0.08 0.1 Time (s)

CFX Stavovy-Chuang Ochi-Motter Lloyds Register

FIGURE 52: VERTICAL IMPACT FORCE PER UNIT LENGTH AGAINST TIME FROM EMPIRICAL, NUMERICAL AND EXPERIMENTAL DATA FOR THE LOWER PORTION OF THE BOW SECTION WITH HIGH IMPACT SPEED.

51

6.0- Data Analysis and Discussion


6.1Lewis, et al.s (2010) Wedge Analysis

As it is evident from the figures in section 5.1, all results in this section are subjected to oscillations of very large magnitude starting within the slamming event time. This has been interpreted to be caused by the density of the wedge, which is less than that of water for both employed masses. This would require the buoyancy force to dominate after the first initial transient impact and the body would need to float toward the end of the simulation. However, as shown in Appendix VII, the rigid body continues in its downward motion, with an alternation of positive and negative pressures below it for consecutive time steps. This has been interpreted as a bug, where the rigid body solution (section 3.2 and 11.2) overrides all other flow equations. ANSYS has been contacted, but replied that the feature is supported in version 13, which is not available to the student. Therefore, all results in section 5.1 are strongly biased by this issue and should be discarded. Furthermore, the pressure magnitude is affected by the limited domain size. Nevertheless, from figures 2 to 4, it is evident that finer meshes result in greater magnitude of the relative pressure, thus presenting closer agreement with the experimental measurements as expected. In addition, for them a longer time elapses from impact before the oscillatory behaviour begins. Conversely, finer time steps (or lower Courant numbers) counterintuitively cause the magnitude of the peaks to lower. Overall, file c_0.0025, which presents a not excessively fine mesh and the upper time step, would be the wise choice for future studies, considering the linear increase in the computation time with number of elements and the convergence of the magnitude of the pressure data to an asymptotic value, as per good CFD practice (Bressloff, 2011). Figures 5 to 7 in section 5.1 show that, while a change in the initial and hence impact velocity results in changes in both magnitude and time occurrence of the pressure peaks, mass variations cause only differences in the magnitude of the relative pressure. This is valid for all monitor points and is in good agreement with the experimental observations (Lewis, et al., 2010), presented in Appendix VI. In addition, from the comparison of figures 47 and 48 in section 5.5, it can be seen that the disagreement with the experimental data augments for increasing initial downward velocity at P1.
52

6.2-

Aarsnes (1996) Wedge Analysis

As it can be seen from figures 8 to 10 in section 5.2, the magnitude of the pressure calculated by CFX is much lower than the experimental measurements for all monitor points (by about 50%). Nevertheless, the temporal domain tends to be accurate, with the exception of monitor point 1, which lies closer to the tip of the wedge, in a region where air entrapment occurs, as visible in figure 85 in Appendix VII. In this case, the density of the rigid body is greater than that of water. Therefore, as compared with section 6.1, the persistent downward motion of the wedge imposed by the rigid body equations is correct, which is the reason for the lack of oscillations. Hence, the issues associated with the rigid body solution cannot be the cause for the low pressure magnitude in this case. This is believed to be due to the neglecting of the hydroelastic effects associated with slamming, which become significant for low deadrise angle (in this case only apparent due to the heel angle) and high impact speed (Faltinsen, 2000). On the other hand, the vertical force component per unit length calculated by CFX is very similar to the experimental measurements (figures 11 and 12 in section 5.2). In fact, it even presents a slightly larger magnitude. Conversely, the horizontal force per unit length is still underestimated, which is probably caused by the neglected hydroelastic effects.

6.3-

Ship-like Section Data Analysis

Overall, the numerical results for the ship-like section are in good agreement with the experimental measurements, as it can be seen from the figures in section 5.3. However, in all cases, the CFX solution is never accurate for the first time steps. Typically, the more evident differences occur at P1 and P4 as expected, as these points lie at the bottom and under the flare of the section respectively. In addition, the solver becomes less accurate with increasing heel angle and impact speed as per the other numerical approaches employed by Brizzolara, et al. (2008), although the great exception is represented by P4 for the 28.3 heel angle simulation. Usually, not only the magnitude of the numerical results for the relative pressure, but also its temporal distribution is very accurate, as per the results from FLOW-3D (2011) and SPH (Viviani, et al., 2007), which showed the closest agreement with the experimental measurements in the paper by Brizzolara, et al. (2008). In fact, CFX actually behaves even better than the
53

competing approaches, as it presents fewer oscillations. The most worrying feature of the CFX pressure results is their tendency to grow rather than decay toward the end of the analysed simulation time. This is caused by the rigid body motion overriding the other equations, which results in the bow section proceeding in its downward travel. Nevertheless, this does not cause instability problems in the solver as per the Lewis, et al.s (2010) wedge (section 6.1), since the mass of the section is greater than that of the water it displaces (therefore, it cannot float). The numerical data for the vertical and horizontal forces per unit length present greater discrepancies from the experimental measurements that the pressure results. The

numerical peak vertical force component per unit length is slightly underestimated and shifted towards a lower time after impact for both 0 heel angle simulations. In addition, in both cases, the curve is rising for a time greater than 0.15 s as per the pressure at P1. The horizontal force component per unit length has been presented for the 0 heel angle drops so as to show the greater accuracy of the numerical results than the experimental data. Indeed, symmetry of the section would require the flow in the two verses of the xdirection to cancel each other out. The greater magnitude of the oscillations in the experiments can be explained by considering rotatory oscillations of the section after impact, which have not been allowed in the software (the motion has been constrained to translation along the y-axis only as explained in section 3.2). The greater negative values of the experimental horizontal force per unit length for the high impact speed case suggests calibration errors or a side motion of the rigid body, which can in fact question the validity of this experiment for a time greater than 0.08 s. On the other hand, a horizontal force component is to be expected for the drop of the section with an initial heel angle. For a heel angle of 9.8 and a low impact speed, the magnitude of the numerical results for the x-force per unit length is precise, but it presents a shift in the time distribution and no double peak. For a heel angle of 28.3, the temporal domain of the CFX solution for both horizontal and vertical forces per unit length is correct, but the magnitude is underestimated for the former and overestimated for the latter. This is in agreement with the observations of Aarsnes (1996) wedge (section 6.2) and it is possibly caused by the neglecting of hydroelastic effects. For a heel angle of 9.8 and a high impact speed, CFX anticipates the occurrence of the peaks of both variables.

54

From the study in the node density (figures 14 to 19), there is no clear distinction in the pressure and force results for the different meshes. The coarser grid is slightly less accurate and presents fewer oscillations, also due to its coarser time steps for Courant number equality. Although the dependence of the total computation time on the total number of elements is typically in the form of a second order polynomial (Bressloff, 2011), a linear relationship is in fact found as per figure 13 in section 5.3, probably due to the need to maintain a finely spaced mesh close to the section surface even for the coarsest grid. Nevertheless, owing to the excessive computation time of the fine mesh, the node density of the medium grid has been used in all subsequent simulations, although there are differences in the number of nodes and elements for the files presenting a heel angle because of the different geometry of the domain. Furthermore, from figure 19 in section 5.3, it is clear that there is no evident difference in the vertical force data per unit length whether the force acting against the section top surface is considered or not. This justifies the assumption of the negligibility of the suction effects against this region and supports the simpler approach of selecting a monitor point including the whole boundary condition for the section surface.

6.4-

Three-dimensional Ship-like Section Analysis

From figures 42 to 45 in section 5.4, it is clear that the three-dimensional effects slightly decrease the magnitude of the relative pressure, siting the three-dimensional results inbetween the numerical two-dimensional data and the experimental measurements. The great exception is represented by P4, where the pressure is largely underestimated, probably due to the location of this point under the flare. The results from the threedimensional domains are very close for all curves, except for the file x1.5, which presents the smallest ratio and consequently the more cramped mesh towards the

final stages of the slamming event, as visible in figures 114 and 115 in Appendix VII. Its pressure has a slightly larger magnitude than the other results. The investigation in the minimum size of the domain for a three-dimensional simulation showed all the faults of the selected approach, which relies on mesh deformation. Indeed, firstly, reaching the very small tank length employed in the experiment for the camera recordings (Aarsnes, 1996) was impossible due to fatal errors in the solver because of the occurrence of mesh folding. Secondly, even for the relatively large domain sizes selected, the expansion factor and orthogonality angles of the grids
55

exceeded the prescribed limits as evident in table 12 in section 5.4, resulting in an incorrect description of the section surface, as shown in figure 111 in Appendix VII. Figure 46 in section 5.4 shows that the three-dimensional effects also cause a reduction in the vertical force per unit length as compared to a purely two-dimensional simulation. The experimental measurements present a larger magnitude, thus supporting the validity of the assumption by Aarsnes (1996) of approximating two-dimensional results by taking the readings along the central portion of the rigid body only.

6.5-

Discussion of the Empirical Results

For both wedge impacts (figures 47 and 48 in section 5.5), both Wagner (1932, cited in Faltinsen, 1993) and Stavovy and Chuangs (1976) methods underestimate the magnitude of the impacting pressure, presenting results similar to those computed by CFX. This is very unusual, as Stavovy and Chuangs (1976) approach is believed to be accurate for a deadrise angle of 25. Hence, it is possible that the initial downward velocity assumed for both numerical and empirical computations may be wrong, thus causing the underestimate in the magnitude of the pressure measurements in both cases. On the one hand, the exponential time distribution of the empirical results is in very good agreement with the experimental measurements, presenting a pronounced peak. This is as expected, since slamming events are typically characterised by an initial sharp peak, followed by a slower variation owing to the rate of change of fluid momentum and to the alterations in the buoyancy force (Belik, et al., 1980). On the other hand, the time of occurrence of the peak pressure is shifted to 0.002 s after the impact time because of the value.

The peak impact pressure for the ship-like section is calculated in the same procedure for all empirical methods, as the theoretical equations can be considered variants of equation (6) in section 4.0. Hence, the validity of each technique depends purely on the magnitude of the value, visible in section 4.0. From figure 49 and 50 in section 5.5, it is evident that Stavovy and Chuangs (1976) method greatly overestimates the peak impact pressure, even for low impact speed. This is mainly due to the 0 deadrise angle, despite its very small width, which results in a very large value. In fact, the modified

procedure provides much more reasonable results, while relying only on a minor geometrical approximation. The approach recommended by Lloyds Register (2011 a;
56

2011 b) is found to only minimally underestimate the impact pressure for low initial speed, while it significantly overestimates it for the higher impact speed. Similarly, Ochi and Motters (1973) procedure largely underestimates the pressure for the low impact speed, while for the higher initial downward velocity the magnitude of the impact pressure matches the numerical and experimental pressure at P1. The

exponential time distribution is more accurate for the high impact speed, where an initial pressure peak is present. Nevertheless, it is evident that the employed time profile is not suitable to represent the second peak, which is the most important one for the low impact speed case. A triangular profile, as first proposed by Ochi and Motter (1971, cited in Belik, 1982), would be more efficient for the slamming event of the ship-like section. Also for the vertical force component per unit length acting against the lower portion of the section, the validity of the exponential time distribution of the empirical solution is dubious, as visible in figures 51 and 52 in section 5.5. Considering the magnitude only, for the low impact speed, Stavovy and Chuangs (1976) approach provides the closest estimate to the numerical and experimental data, while Ochi and Motters (1973) technique underestimates the magnitude of the vertical force per unit length and Lloyds Registers (2011 a) rules overestimate it, even though only by a small amount. For the high impact speed case, Ochi and Motter (1973) provide the closest approximation, although it is already an overestimate. The magnitude of the force per unit length of both Stavovy and Chuang (1976) and Lloyds Registers (2011 a; 2011 b) approaches largely exceeds the experimental and numerical results. For the latter, this can be partly explained by considering the assumption of a constant impact pressure acting against the whole lower portion of the section, described in section 4.0, which may not be completely appropriate. In the method by Stavovy and Chuang (1976), the results are negatively affected by the enormous pressure value calculated for the first segment of the section, which presents a 0 deadrise angle, as per table 6 in section 4.0. It should be noted that the results from Lloyds Register (2011 a; 2011 b) rules are expected to overestimate the slamming pressures for safety concerns. In addition, too few experiments have been analysed to draw any significant conclusions on the empirical approaches. Similarly, when three-dimensional effects are considered (as experienced by a ship at sea), Stavovy and Chuangs (1976) approach is found to provide a suitable approximation (Beukelman, 1980, cited in Belik, et al., 1980).
57

7.0- Conclusions
The rigid body solution, presented as a beta feature in ANSYS CFX v12.1, has been found to be unable to model the free drop of a wedge geometry having a density less than that of water. This is caused by the overriding of all other flow equations by the rigid body solver. For this reason, ANSYS has been contacted and notified the issue. In the simulation of the fall of a different wedge geometry, which presents high impact speed and a low effective deadrise angle due to heel, hydroelasticity and air-entrapment have been found to be significant. Hence, for neither wedge simulation did CFX show good agreement with experimental measurements (Lewis, et al., 2010; Aarsnes, 1996) for the slamming impact pressure and force data. Conversely, CFX was found very satisfactory in the estimate of the magnitude and spatial and temporal distributions of the impact pressure and loads against a bow section for various heel angles and impact speeds, as compared with experimental data (Aarsnes, 1996). However, three-dimensional simulations have been found to present worse agreement with experimental measurements than two-dimensional models, while requiring much greater computational resources and time. In addition, they result in mesh folding for small domain dimensions and cause severe mesh deformation around the body surface for the selected approach, which relies on node motion. The empirical formulations (Wagner, 1932, cited in Faltinsen, 1993; Stavovy and Chuang, 1976) present lower peak impact pressure as compared with the experimental results for the wedge slamming event, possibly due to the incorrect estimate of the impact velocity. The exponential time distribution shows a correct peak accentuation, but requires a shift of the peak pressure towards higher slamming event times. For the bow section, most empirical methods (Lloyds Register, 2011 a; 2011 b; Ochi & Motter, 1973) underestimate the slamming pressure magnitude for low impact speeds, while they greatly overestimate it for higher initial downward velocities. Since they all rely on variations of the same equation ((6) in section 4.0), it is believed that assuming the slamming pressure to be proportional to the impact speed square may be an oversimplification, although a larger amount of data is required for validation. In addition, the exponential profile employed for the time distribution is found to be unsatisfactory. For the estimation of the slamming vertical impact force, the theory by Stavovy and
58

Chuang (1976), which divides the lower portion of the section surface in a number of panels, is found to provide best agreement with the experiment, while the method adopted by the student of assuming the impact pressure to vary only slightly against the lower portion of the section was found to be unsuitable. As expected, Lloyds Register (2011 a; 2011 b) rules overestimate the slamming pressure for safety concerns.

59

8.0- Future Work Recommendations


Since the causes for the inaccurate behaviour of CFX with the wedge free-fall are not fully understood, as explained in sections 6.1, 6.2 and 6.5, more simulations should be performed with a wedge geometry, possibly with a density greater than that of water, no heel angle and known impact speed. Of all the papers presented in section 2.0, Aarsnes (1996) measurements on wedge free drops are the only ones for which sufficient data are provided. Nevertheless, the student could not access Appendix A (Aarsnes, 1996), where the results are stored. In addition, it would be interesting to model the wedge as an immersed solid and/or to employ the rigid body solution in ANSYS version 13 or later for the wedge geometries presented in section 3.2 in order to discover possible improvements over the results shown in sections 5.1 and 5.2. The empirical formulae discussed in section 4.0 provide the required economical means of evaluating the maximum impact loads due to slamming of a ship during the preliminary design stage. However, a more complex theoretical treatment may be necessary as mentioned in section 7.0. Nevertheless, more data should be available for this purpose, preferably including results from global slamming analyses. In addition, for the bow section, a different equation should be adopted for the time distribution of the pressure measurements in order to match the experimental measurements, possibly with the use of ramp functions. Due to the high cost of the ANSYS package, a free software RANSE solver such as OpenFoam (2011) could be used in order to provide a more economical approach. The experimental data and the results from CFX can be used as a benchmark for the validation of the results. In addition, the effects of compressibility and/ or

hydroelasticity can be considered as well, possibly with the coupling with an external FEA solver. Nevertheless, it is known that more initial experience will be required by the user. By considering the accuracy of the numerical results for the bow section, it is envisaged the use of a RANSE solver for simulating the transient behaviour of an actual ship or model in waves. This corresponds to a general slamming analysis and good examples are provided by Oberhagenmann, et al. (2008) and Kapsenberg and Thornwhill (2010). Nevertheless, it is believed that previous knowledge in the software, longer computation
60

time and coarser meshes will be necessary. Furthermore, it is recognised that less reliable experimental data is available for this case. As the three-dimensional simulations performed in this individual project provide worse agreement with the experiment that two-dimensional cases and also result in mesh folding for low ratio, it would be interesting to try using a different solving approach. This can be in the form of a fixed-mesh, moving-flow technique, as the one employed by Lewis (2011), although it is recognised that this tends to work for high impact speeds only. This may also be done using open software.

61

9.0- References
Aarsnes, J. V., 1996. Drop tests with ship sections- effect of roll angle. Marintek Report No. 603834.00.01. Abaqus, 2011. Abaqus/CFD. [Online] Available at: http://www.abacom.de/products/abaqus_cfd.html [Accessed 27 11 2011]. ANSYS, 2009. ANSYS CFX-Help guide, s.l.: ANSYS Inc.. Battley, M., Stenius, I., Breder, J. & Edinger, S., 2005. Dynamic characterisation of marine sandwich structures. Sanwich Structures 7: Advancing with sandwich structures and materials. Part 5, pp. 537-546. Bedford, A. M. & Fowler, W., 2007. Engineering mechanics: statics and dynamics. 5th ed. s.l.:Prentice Hall. Belik, O., 1982. Symmetric response of ships in regular and irregular waves. s.l.:University College London, PhD Thesis. Belik, O., Bishop, R. E. D. & Price, W. G., 1980. On the slamming response of ships to regular head waves. Trans. RINA, Volume 122, pp. 325-338. Belik, O., Bishop, R. E. & Price, W. G., 1983. A simulation of ship responses due to slamming in irregular head waves. Trans. RINA, Volume 125, pp. 237-254. Belik, O., Bishop, R. E. & Price, W. G., 1988. Influence of bottom and flare slammin on structural responses. Trans. RINA, Volume 130, pp. 325-337. Bertram, V., 2000. Practical Ship Hydrodynamics. Oxford: Butterworth-Heinemann. Bressloff, N. W., 2011. Good Practice in CFD. Lecture at the University of Southampton on 6th Otober 2011. Brizzolara, S. et al., 2008. Comparison of experimental and numerical loads on an impacting bow section. Ships and Offshore Structures, Volume 34, pp. 305-324. Constantinescu, A., El Malki Alaoui, A., Neme, A. & Rigo, P., 2009. Numerical and experimental studies of simple geometries in slamming. Proceedings of the 19th International Offshore and Polar Engineering Conference, Osaka, Japan, 21-26 June 2009, pp. 589-598. Engle, A. & Lewis, R., 2003. A comparison of hydrodynamic impacts prediction methods with two dimensional drop test data. Marine Structures, Issue 16, pp. 175-182. Faltinsen, O. M., 1993. Sea loads on ships and offshore structures. Cambridge: Cambridge University Press. Paperback edition.

62

Faltinsen, O. M., 2000. Hydroelastic Slamming. Journal of Marine Science and Technology, Issue 5, pp. 49-65. Faltinsen, O. M., 2005. Hydrodynamics of high-speed marine vehicles. Cambridge: Cambridge University Press. FLOW-3D, 2011. CFD Software: FLOW-3D Overview. [Online] Available at: http://www.flow3d.com/flow3d/flow3d_over.html [Accessed 02 11 2011]. Homsy, G. M. et al., 2007. Multimedia Fluid Mechanics, s.l.: s.n. Huera-Huarte, F. J. & Gharib, D. J., 2011. Experimental investigation of water slamming loads on panels. Ocean Engineering, Issue 38, pp. 1447-1355. iSolutions, 2011. The Lyceum Linux Teaching Cluster Service. [Online] Available at: http://www.soton.ac.uk/isolutions/computing/hpc/compute/index.html [Accessed 12 12 2011]. ISSC, 2009. Impulsive Pressure Loading and Response Assessment. [Online] Available at: http://139.30.101.246/ISSC2012/Material/2009/IMPULSIVE%20PRESSURE%20LOADING%20A ND%20RESPONSE%20ASSESSMENT.pdf [Accessed 31 October 2009]. ITTC, 2008. Final report and recommendations to the 25th ITTC. Proceedings of the 25th ITTC, Volume I, pp. 224-232. Kapsenberg, G. K. & Thornwhill, E. T., 2010. A practical approach to ship slamming in waves. 28th Symposium on Naval Hydrodynamics, Pasadena, California, 12-17 September 2010. Lee, B. H. et al., 2009. Numerical simulation of impact loads using a particle method. Ocean Engineering, 37(2-3, Feb. 2010), pp. 164-173. Lewis, S. G., 2011. Prediction of planing craft motions in waves using 2D CFD coupled with strip theory. s.l.:DoE Thesis. University of Southampton. Lewis, S. G., Hudson, D. A., Turnock, S. R. & Taunton, D. J., 2010. Impact of a free-falling wedge with water: synchronized visualization, pressure and acceleration measurements. The Japan Society of Fluid Mechanics and IOP Publishing Ltd., Issue 42. Lloyd, A. R., 1998. Seakeeping. Revised Edition ed. Chichester, UK: Seakeeping. Lloyds Register, 2011 a. Rules and Regulations for the classification of naval ships. Volume I, Part 5, Chapter 3, Section 4. Lloyds Register, 2011 b. Rules and Regulations for the classification of trimaran ships. Volume I, Part 5, Chapter 5, Section 2.

63

LS-DYNA, 2011. LS-DYNA: A combined Implicit/Explicit solver. [Online] Available at: http://www.ls-dyna.com/ [Accessed 02 11 2011]. Maki, K. J., Lee, D., Troesch, A. W. & Vlahopoulos, N., 2011. Hydroelastic impact of a wedgeshaped body. Ocean Engineering, Issue 38, pp. 621-629. Mehedi Sayeed, T., Peng, H. & Veitch, B., 2010. Experimental investigation of slamming loads on wedge. Dhaka, Bangladesh, Proceedings of MARTEC 2010. Oberhagenmann, J. et al., 2008. Stern slamming of an LNG carrier. Proceedings of the ASME 27th International Conference on Offshore Mechanics and Arctic Engineering, 15-20 June 2008, Estoril, Portugal. Ochi, M. K. & Motter, L. E., 1973. Prediction of slamming characteristics and hull responses for ship design. SNAME Transactions, pp. 144-176. OpenFoam, 2011. OpenFoam: open source CFD. [Online] Available at: http://www.openfoam.com/ [Accessed 13 12 2011]. Paik, K.-J., Carrica, P. M., Lee, D. & Maki, K., 2009. Strongly coupled fluid-structure interaction method for structural loads on surface ships. Ocean Engineering, Issue 36, pp. 1346-1357. Stavovy, A. B. & Chuang, S.-L., 1976. Analytical determination of slamming pressures for highspeed vehicles in waves. Journal of Ship Research, 20(4), pp. 190-198. Tveitnes, T., Fairlie-Clarke, A. C. & Varyani, K., 2008. An experimental invesitgation into the constant velocity water entry of wedge-shaped sections. Ocean Engineering, Issue 35, pp. 1463-1478. Viviani, M., Brizzolara, S. & Savio, L., 2007. Evaluation of slamming loads on ship bow section adopting SPH and RANSE methods. Proc. IAHR. Paper SS-09-06. Wilson, P. A., 2011. Marine Hydrodynamics Lecture Notes. SESS3001: Marine Hydrodynamics. Univeristy of Southampton, unpublished. Yang, Q. & Qiu, W., 2010. Computation of slamming forces on three-dimensional bodies with a CIP method. Proceedings f the 29th International Conference on Ocean, Offshore and Arctic Engineering, 6-11 June, 2010, Shanghai, China. Yang, S. H. et al., 2007. Experimental and numerical study on the water entry of symmetric wedges and a stern section of modern containership. 10th International Symposium on Practical Design of Ships and Other Floating Structures, Houston, Texas.

64

10.0- Appendix I-Further Information on Numerical Methods


The Boundary Element Method is based on the two-dimensional solution of velocity potentials with the body velocity being constant or variable, while the free surface is updated with a simplified equipotential boundary condition on the free surface (Brizzolara, et al., 2008). The Smoothed Particle Hydrodynamic is a Lagrangian meshless CFD method. It

discretises the fluid in a number of particles and solves the flow using the Navier-Stokes and continuity equations in a discrete form through a kernel function (Viviani, et al., 2007). Its results tend to be oscillatory for constant speed or forced drops (Brizzolara, et al., 2008). The Moving Particle Semi-implicit method has been developed by Koshizuka and Oka (1996, cited in Lee, 2009). It is a Lagrangian meshless method for incompressible, freesurface simulations (Lee, et al., 2009). The LS-DYNA approach models the fluid domain with a multi-material Eulerian formulation (Brizzolara, et al., 2008). FLOW-3D is a commercial finite volume-finite differences, RANSE solver that does not require the specification of wall and boundaries in the geometry, as it employs a fractional area volume obstacle representation (Brizzolara, et al., 2008). ABAQUS is a finite element solver, which allows deforming-mesh Lagrangian Eulerian (ALE) description of the equations of motion, heat transfer and turbulent transport (Abaqus, 2011; Constantinescu, et al., 2009). In the paper by Brizzolara, et al. (2008) the FLUENT solver has also been used. However, wrong boundary conditions have been set, so that no useful data can be used in the comparison with the ANSYS CFX results of this paper.

65

11.0- Appendix II-Computational Fluid Dynamics Theory


This section has been mainly based on the ANSYS theory guide (2009), as well as the book by Homsy, et al. (2007). CFX is finite-volume solver, which has been now (version 12.1) fully integrated within the Workbench suite by ANSYS (2009). This has resulted in a full integration of problem definition (ANSYS CFX-Pre), analysis (CFX-5 solver) and results presentation (ANSYS CFD-Post) as well as geometry (Ansys DesignModeler) and mesh construction tools (either CFX-Mesh or Mesh). The main equations of flow that CFD programs try to solve are the continuity equation and the Navier-Stokes equations. From Homsy, et al. (2007), for incompressible flows, the continuity equation reduces to: . This is a mathematical statement of the conservation of mass. (12) Similarly, for

incompressible the Navier-Stokes equations are provided by (Homsy, et al., 2007): ( ) . (13)

The continuity, momentum and total energy equations for all flows, including a source term, can be seen in the theory guide by ANSYS (2009). These are known as transport equations (ANSYS, 2009). A source term for buoyancy calculations should be

included, as per equations (2), (3) and (4) in section 3.0, which is simply represented by the product of the gravitational constant and the difference between the pressure of the selected fluid and a reference pressure (typically set to be that of the lightest fluid employed in the simulation) (ANSYS, 2009). This assumes that the full buoyancy model is used, as the current simulations did not include a unique fluid, where the density variation is driven only by a small density variation, thus requiring the Boussinesq complex model (ANSYS, 2009). For closure, constitutive equations of state for density and for enthalpy are necessary (Homsy, et al., 2007; ANSYS, 2009). The incompressible equations of state have been used in the current analysis for all simulations, although a more appropriate approach would have been to treat air as an ideal gas and to include compressibility effects for water as well. All these equations

66

can be found in the theory guide by ANSYS (2009). On the other hand, no conjugate heat transfer equations (ANSYS, 2009) have been necessary, as the solution of the heat transfer equations across the rigid body were not necessary. In fact, the performed CFD analysis concerns a turbulent flow, as explained in section 3.4, where the flux has a highly unstable and random behaviour, presenting turbulences or eddies. Nevertheless, if larger time scales are analysed, even turbulent flows may be defined to exhibit averaged characteristics with an additional time varying component (ANSYS, 2009). Hence, for instance a velocity component and a time-varying component : . (14) can be divided into an averaged

In equation (14), the averaged component is obtained as: . (15)

As explained in section 3, turbulence models are required that introduce average and fluctuating quantities to produce Reynolds-Averaged Navier-Stokes (RANS) equations (ANSYS, 2009), which can be seen in equations (1), (2), (3) and (4) in section 3.0. In equations (2) and (3) in section 3.0, is the molecular shear tensor (N/m2), which

includes both normal and shear components of the stress. Although the continuity equation ((1) in section 3.0) has not been altered, the momentum and scalar transport equations contain turbulent flow terms additional to the molecular diffusive fluxes. These terms, , are the Reynolds stresses (Wilson, 2011). Furthermore, equation (3) contains another turbulence flux term, . The mean total enthalpy is given by: . In equation (4), is the Reynolds flux and the additional variable is given by: . (17) (16)

Greater detail on these turbulent equations is provided in sections 11.4 and 11.5, although it should be noted that their treatment goes beyond the scope of this project. Numerous references are provided by the ANSYS theory guide (2009) in addition to all relevant equations.
67

11.1- Additional considerations on transient simulations


The time dependence of the flow can be specified in ANSYS CFX as either steady state or transient (ANSYS, 2009), where transient implies a dependence of the flow properties on time. In these simulations, real time information is necessary in defining the time intervals at which the CFX-Solver calculates the flow field (Lewis, 2011). For transient flow simulations, the Courant number ( ) is important in determining the is required for

time step. In fact, even for CFX, which is an implicit code, a small

accuracy in certain simulations. For a one-dimensional grid, the Courant number is defined by (Lewis, 2011): . In equation (18), is the fluid speed, is the time-step and (18) is the cell size. The

Courant number calculated in CFX is a multidimensional generalization of this expression where the velocity and length scale are based on the mass flow into the control volume and the dimension of the control volume (ANSYS, 2009). The second-order, backward Euler transient scheme is an implicit scheme, suitable for both fixed and varying time steps (ANSYS, 2009). Nevertheless, it is fundamental to notice that the transient scheme for turbulence equations will remain first order and the transient scheme for volume fraction equations will be set to a bounded second order scheme, as it is appropriate (ANSYS, 2009).

11.2- Additional considerations on the geometry description and rigid body solution
The rigid body feature is still a beta feature in ANSYS version 12.1 and the results should thus be evaluated with care. In this tool, the rigid body can be defined by its outer contour. In addition, the mass, its centre of gravity and the inertia matrix can be specified as well. Furthermore, the motion can be constrained to translational or

rotational or part of them. In the current case, since the motion has been constrained to translational in the y-direction only, the inertia matrix terms were set to zero for both cases, since no rotation is allowed. However, as described in section 6, this feature is found to be faulty for modelling the fall of bodies onto a fluid, if they present a smaller
68

density than the fluid itself. Indeed, it was observed that the rigid body solution overrode all other flow equations, resulting in the body moving along the negative yaxis until mesh folding occurred. For this reason, ANSYS has been contacted. Despite the symmetry of both geometries about the 0y axis, no symmetry plane could be used to shrink the number of elements due to the deforming mesh, which would otherwise have resulted in a moving domain. Since the movement of the domain is to be specified by user input, this would have greatly undermined the validity of the results.

11.3- Additional considerations on mesh generation


The mesh design process is based on the compromise among the following requirements (Lewis, 2011): Spacing. A finely spaced grid is necessary to ensure acceptable levels of

conservation of mass and momentum (Lewis, 2011). However, finer meshes result in longer computation times and memory requirements (ANSYS, 2009), where the computational requirements of iterative algorithms are typically proportional to the number of cells squared (Bressloff, 2011). In addition, a smaller cell size needs smaller type steps as defined by the Courant number (Lewis, 2011) for transient solutions. Resolution. A sufficiently small grid size is necessary for the resolution of the fluid flow in all its regions (Lewis, 2011), which is fundamental for turbulent boundary layers. In particular, the position of the first node near the wall boundaries can significantly influence the accuracy of the results. Pressure peaks are thus not resolved in detail for impulsive flows (Lewis, 2011). Geometry. A good representation of the geometry is fundamental, especially in its most influential features. In Ansys CFX-Mesh it is possible to specify the values of the Reynolds number, and a basis length in order to create inflated boundaries near the desired bodies that require a specific the actual value. However, this feature was found to greatly underestimate

value and thus has not been used. is complicated, but it is fundamental in CFD. It is a non-

The concept of

dimensional measure of the distance of the first node from the boundary and is given by
69

equation (20) (Lewis, 2011). Equation (19) shows the non-dimensional form of the velocity from the wall (Lewis, 2011). ; (19)

(20)

In equation (19), and

( )

(ANSYS, 2009). It is evident that both the concepts of

are valid only within the turbulent boundary layer near a no-slip wall

(Homsy, et al., 2007). Indeed, on the wall, the turbulent kinetic energy, , is zero, while the turbulent dissipation, , has a non-zero value. The turbulent boundary layer can be thus subdivided into three parts (ANSYS, 2009): The viscous sub-layer, which occurs very close to the wall (for turbulence effect must stop near the wall. The log-law layer ( ), where both the viscous and the turbulent ), as the

forces are significant. In this region, equations (19) and (20) become and respectively (ANSYS, 2009).

The outer layer lies at the extremities of the boundary layer, where it merges with the external flow. Here, the turbulent forces are dominant.

Since a Direct Numerical Simulation of the turbulent boundary layer would exert too heavy a load on the memory requirements of current computers, turbulence models have been developed that approximate the Navier-Stokes equations, but are typically applicable in the log-law layer (ANSYS, 2009). This is described in further detail in the next two subsections, although the importance of the value lying within the

prescribe range of 30-100 (Bressloff, 2011) should now be clear. Otherwise, another possibility would be having all values lower than 5, as also the viscous sub-layer is

stable (Wilson, 2011). It could be then treated as a laminar layer up to the first node from the no-slip boundary. Although values as low as 0.9 have been obtained, it is very difficult to have all values below 5, as it would require enormous computational value was 40.

resources. In fact, in the current simulations the lowest maximum

70

For all simulations, a maximum 0.05 m node distance and minimum radius of influence of 0.1 m have been applied around the body surface, although this is a soft constraint (it can be overridden by the first layer thickness of the inflated boundary layers (ANSYS, 2009)). This enables the location of monitor points close to the position of the pressure gauges employed by Lewis, et al. (2010) and Aarsnes (1996). The monitor points are fundamental, since they store the selected data (pressure measurements) at the node closest to the specified coordinates. Hence, even though the node is moving, the monitor follows its displacement (ANSYS, 2009). Nevertheless, there is an inevitable associated error, as hardly ever does a node lie on the required location, which has been accepted, as described in section 5.0. The theoretical positions of the monitor points can be seen in Appendix III. In addition, the concepts of elements orthogonality and aspect ratio should be considered as they are a measure of mesh quality. Indeed, controlling mesh quality is of primary importance in minimizing discretisation errors (ANSYS, 2009). Orthogonality regards the comparison of the angle between adjacent element edges against the optimal value for an orthogonal shape. As reported in section 3.3, 20 is the minimum

orthogonality angle (ANSYS, 2009). The element aspect ratio concerns the degree of stretching of the mesh elements (ANSYS, 2009). The most relevant measure for the CFX-Solver involves the aspect ratio at each node, which is calculated as the ratio of the maximum to minimum integration point surface areas in all elements adjacent to the node (ANSYS, 2009). A suitable aspect ratio is for values larger than 100 as reported in section 3.3 (ANSYS, 2009). The mesh expansion factor is the final method of mesh quality control (ANSYS, 2009). The prescribed maximum recommended value is 20 (ANSYS, 2009). Furthermore, in the case of the deforming mesh, consideration to nodes motion was necessary in order to avoid solver errors caused by a folding mesh. For this reason, inflated layers of cells were employed at the top and bottom of the geometry, as mentioned in section 3.3. In addition, the mesh elements are coarser away from the rigid body (grid anisotropy) as compared to the fixed geometry so as to facilitate the shifting of the nodes, as well as to reduce the computational resources.

71

11.4- Additional considerations on turbulence models


In order to determine the nature of the flow to be modelled, the Reynolds number need to be estimate for all cases, where the Reynolds number is given by equation (21): . (21)

Hence, the Reynolds number, which is a dimensionless parameter, can be considered as the ratio of inertial and viscous forces. For the wedge case, it has been assumed that m, which is the wedge height, m/s and m/s from m and

equations (22) and (23) in Appendix III. For the ship-like section,

m/s (the former value applies to the straight section, whereas the latter to the heeled one) and m/s. As mentioned in section 3.4, fresh water m2/s (ANSYS,

has been assumed for all simulations. Therefore,

2009). The results are reported in table 4 in section 3.4. Nevertheless, the impulsive nature of the slamming event as well as the lack of initial knowledge in the velocity profile after the rigid bodies enter the water mean that the calculated highly approximated. This sub-section is chiefly based on the theory, modelling and user guides by ANSYS (2009). In order to avoid a long list of citations, no other direct citations have been included in the text unless they are taken from other authors. Turbulent flows are characterised by eddy formation and highly irregular particle paths, where every particle has a different pathline and streamline (Wilson, 2011). Hence, it is a complex threedimensional, unsteady process, which consists of many scales. However, it strongly influences the characteristics of the flow. Turbulence occurs for , i.e. when the values are

inertia forces in the fluid are significant as compared to the viscous forces (Homsy, et al., 2007). Therefore, in most simulations turbulent flows actually arise, which would generally involve length scales smaller than the smallest finite volume element. In addition, the Direct Numerical Simulation of these flows would need enormous resources of computing power. Nevertheless, considering larger time scales than those of turbulent fluctuations, turbulent flows exhibit average characteristics, with an additional time-varying, fluctuating component (Homsy, et al., 2007). Hence, a less computationally intensive approach is to modify the original Navier-Stokes equations with the introduction of averaged and fluctuating quantities to produce Reynolds
72

Averaged Navier-Stokes equations (RANSE). Indeed, these equations calculate the mean flow quantities only, modelling all scales of the turbulence field, without solving the turbulent fluctuations. Therefore, RANSE obtain a great saving in computational effort as compared with DNS. Turbulence models based on the RANSE are known as Statistical Turbulence Models. However, a side effect of the averaging procedure is the introduction of additional unknown terms containing the products of the fluctuating quantities, which can be considered as additional stresses in the fluid. Known as turbulent or Reynolds stresses, they are difficult to measure directly. In order to achieve closure, which means that the number of equations and unknowns is the same, they require modelling with additional equations of known quantities. The equations employed to close the system define the type of turbulence model. Eddy viscosity models are based on the Boussinesq assumption that the action of the Reynolds stresses is analogous to the action of the viscous stresses. Hence, the effective viscosity is increased to viscosity (Lewis, 2011). , where is the eddy

Two-equation models are the most common,

presenting a good compromise between computational effort and numerical accuracy. The velocity and length scales are solved using separate transport equations. The and the two-equations models relate the Reynolds

stresses to the mean velocity gradients and the turbulent viscosity through the gradient diffusion hypothesis, where the turbulent viscosity is modelled as the product of the turbulent velocity and the turbulent length scales. The turbulence velocity scale is calculated from the turbulent kinetic energy, which is provided from the solution of its transport equation. The turbulence length scale is model

estimated from two properties of the turbulence field. For the

(Launder and Spalding, 1974, cited in ANSYS, 2009)), these are the turbulent kinetic energy ( ) and its dissipation rate ( ), which is obtained from the solution of its transport equation. In the model (Wilcox, 1994, cited in

ANSYS, 2009), the transport equations are solved for the turbulent kinetic energy and the turbulent frequency ( viscosity as . The ), which combined yield the eddy approach performs well close to walls in

boundary layer flows; however, it is sensitive to free stream conditions, which


73

can be exemplified by the distance of the first node from the wall (ANSYS, 2009). It is therefore blended with a function in CFX to produce the Shear Stress Transport turbulence model, which is mainly used for highly accurate predictions of the onset and amount of flow separation under adverse pressure gradients. Reynolds stress models are the most complex classical turbulence models. They are based on the transport equations for all components of the Reynolds stress tensor and the dissipation rate. No use is made of the eddy viscosity, but they rather solve an equation for the transport of Reynolds stresses in the fluid. The Reynolds stress model transport equations are solved for each individual stress component in the fluid. Therefore, they are believed to be more accurate, also in the analysis of similar problems (Lewis, 2011). The only model of this type employed in the preliminary stage of the project was the Speziale, Sarkar and Gatski (1991, cited in ANSYS, 2009), which is based on the available. Greater details can be found both in Homsy, et al. (2007) and especially the ANSYS (2009) theory guide, which includes numerous references and all mentioned equations, which have not been reported, as they go beyond the scope of this report. equation

11.5- Additional considerations on the boundary conditions


Boundary conditions are required within a CFD computation to define the behaviour of the fluid at the edges of the domain. The employed geometry, described in section 3.2 and shown in Appendix III, defines also the surface of the immersed rigid body as an external boundary. The boundary conditions can be divided into two main categories in CFD analyses: the Dirichlet boundary condition, which specifies the values that the fluid has; the Neumann boundary condition sets the gradient of a fluid value at the boundary. In the current simulations, only three boundary conditions have been used, as well as the rigid body feature, described in sections 3.2 and 10.1. No-slip walls mean that the tangential velocity at the wall is zero, which results in the formation of a (turbulent) boundary layer (ANSYS, 2009). Since the normal gradient in the flow variables

become large as the distance from the wall tends to zero within the boundary layer, a
74

large number of elements are required to resolve the flow (Lewis, 2011). Hence, the turbulence model can only be used in regions that are fully turbulent (Lewis, 2011), while it cannot be employed in the laminar sub-layer, where viscous effects are dominant as it can be understood from the Reynolds number. This laminar sub-layer is the one determined by the , as described in the section 11.3. A wall-function

approach has been proposed by Launder and Spalding (1974, cited in ANSYS, 2009) to solve this issue. In Ansys CFX, using the wall-function approach, the viscous sub-layer is linked by employing empirical formulae to provide near-wall boundary conditions for the mean flow and turbulent transport equations. Through these formulae the wall conditions are related to the dependent variables at the near-wall mesh node, which is assumed to lie in the outer layer (ANSYS, 2009). Nevertheless, it should be noted that the predictions are very sensitive to near-wall meshing, as it determines the location of the near-wall node (ANSYS, 2009). In addition, refining the mesh does not necessarily result in increased accuracy (Grotjans and Menter, 1998, cited in ANSYS, 2009). For this reason, ANSYS CFX have developed scalable wall functions to be applied in conjunction with the and Reynolds stress SSG turbulence models, which overcome the problem of inconsistencies in the wall function in the case of fine meshes. Furthermore, these allow performing a consistent mesh refinement independent of the Reynolds number of application (ANSYS, 2009). For this reason, although it was tried to have lying within the safe range of too worrying when lower values ( values

(Bressloff, 2011), it was not considered ) were found at the mid-point of the top and

lowest edge of the wedge and ship-like section alike. While the aforementioned wall-functions allow for consistent mesh refinement, they tend to cause errors in the displacement thickness (up to 25%) for low Reynolds numbers, as they neglect the viscous sub-layer from the mass and momentum balance (ANSYS, 2009). Hence, for the turbulence models, a feature is offered, which near-wall formulation for in the

automatically switches from a wall-function to a low-

increasing mesh refinement. This is based on the analytical expression of viscous sub-layer, which enables the blending the wall vale for

between the

logarithmic and the near-wall formulation. This is a further improvement as compared with the scalable wall-function model, as it allows a consistent insensitive mesh

75

refinement, even for very small values (ANSYS, 2009)) (a value of recommended for accurate predictions (ANSYS, 2009)).

is

On the one hand, in the wall-function formulation, the first point is treated as lying outside the edge of the viscous sub-layer. On the other hand, in the lowmode, the

first mesh point is virtually shifted down through the viscous sub-layer as the mesh is refined. Nevertheless, while the first mesh point always lies at the wall, the first mesh point is in fact treated as if it were away from the wall. This causes a reduction in

the displacement thickness, which represents the error associated with the wall-function. No equations are provided for the scalable wall-function and the automatic model due to their length and complexity, since their scope is beyond this individual project. Nevertheless, the theory guide of ANSYS (2009) includes many references and all necessary equations. If more specific knowledge is desired, especially for particular cases, the online forum available at <http://www.cfd-online.com/> [accessed 19/12/2011] should be tried. A free-slip wall represents a wall boundary condition, but no boundary layer is present on its surface. The velocity normal to the boundary is zero, as is the shear stress value in the wall, but the tangential velocity may have a non-zero value. This boundary condition has been favoured on an opening due to its greater simplicity. In addition, it was feared that an opening would cause instabilities in the solver, as both an outflow, caused by the air moved by the fall of the rigid body, and an inflow, caused by the suction area left behind itself by the body, would be present. For the same reason, the presence of a wall that closes the domain causes both positive and negative air pressures in the domain that must be accounted for. On the other hand, although the side walls should have been wall boundary conditions, openings have been used instead. This was not done to avoid returning waves, as the simulation time is very small, but rather in order to specify the hydrostatic pressure gradient at the sides of the domain as explained in section 3.5. Indeed, it was found that without this option, even though a hydrostatic pressure was set in the initial conditions, CFX could not find the equation and would then choose a random point and obtain the hydrostatic pressure itself using the buoyancy equations in reference to this node. Hence, negative static pressure values were originally present from the very first in the domain. The opening boundary condition solved this problem, as a static pressure value
76

could be input, with a zero gradient turbulence model. In order to impede inflows and outflows, CFX automatically sited a free-slip wall along the whole portion of the openings (ANSYS, 2009). This procedure is typically not recommended due to its unreliability (ANSYS, 2009); however, in this case it was exploited, since it produced a very suitable boundary condition for the problem under study. The resulting hydrostatic pressure was found to be correct and no outflow or inflow was observed. Symmetry boundary planes were set at the front and back of the domain in order to simulate a two-dimensional flow. They are a mixture of the Dirichlet and Neumann boundary conditions (Lewis, 2011). The normal velocities and turbulence quantities are set to zero, as well as the shear stress, resulting in no flow and scalar flux crossing the planes (Lewis, 2011). In addition, the flow parameters at all the points just outside the numerical domain equal those at the nearest node in the mesh (Lewis, 2011), so that both geometrical and flow symmetry are obtained. It should be noted that the no-slip wall that represents the rigid body is set to move in accordance with the equations of motion of the rigid body, so that mesh deformation is possible, which results in the required pressures.

11.6- Additional considerations on multiphase flow


The domain of the fluid surrounding the rigid body is in fact represented by two separate fluids: air and water, which can interact via interphase transfer terms (ANSYS, 2009) (indeed, the phenomenon of air-trapping has been recognised as very influential in slamming problems (Faltinsen, 2000)). Two Eulerian fluids have been employed in the current simulations. The pressure field is shared by all fluids. However, either each fluid may have their own flow field if an inhomogeneous model is used or they may all share a common flow field by employing a homogenous model. In the guide to CFX (ANSYS, 2009) the inhomogeneous model is recommended for free-surface problems, especially those with air entrapment, and has thus been selected. However, better results have apparently been obtained by Lewis (2011) with the homogeneous model. Furthermore, the inhomogeneous model is computationally more expensive. The standard free surface model has been included, with an interphase compression layer of 2, as a distinct interphase is present between the two fluids (ANSYS, 2009). The interphase between the two fluids was modelled in preliminary studies using both
77

the mixture (1 mm thickness) and the free surface models in order to assess their suitability for the current problem, where the difference lies in the equations employed to calculate the interfacial area per unit volume between the two phases, fundamental in computing the momentum, heat and mass transfers. The mixture model is very simple, since it treats both phases symmetrically. Nevertheless, it can provide a good first approximation. The free-surface model is applicable to free surface flows, even if there is entrainment of one phase into the other. Therefore, it has been preferred throughout the project. The water depth and hydrostatic pressure were set manually with the use of equations. A step function was used to set the volume fraction of water to be 1 up to the water surface. The relative hydrostatic pressure was set by employing the same step function, as for such a small height of the air volume, there is no significant change in the hydrostatic pressure of the gas. The equations employed can be seen in the Appendix VIII.

11.7- Additional considerations on error analysis


An accurate error analysis should account for different potential sources of error (ANSYS, 2009; Lewis, 2011): Discretisation errors are generated by localised sources and propagated throughout the domain (Lewis, 2011). While localised errors derive from the high-order terms that are excluded from the discrete approximations of terms in the modelled equations, error propagation depends on the form of the terms that are included in the discrete approximations (Lewis, 2011). In order to minimize these errors, reducing the magnitude of the terms excluded from the discrete approximations is essential. This can be done by increasing their order-accuracy and reducing the mesh spacing in regions of rapid flow variations (Lewis, 2011). Hence, high resolution advection schemes have been used in the solver as well as the anisotropic mesh described in section 3.3, which also limit the error amplification (Lewis, 2011). Modelling errors result from the practical need to approximate physical phenomena with empirical models (ANSYS, 2009), e.g. the wall-function as described in sections 3.4 and 11.4.
78

User errors result from user inexperience and can be minimized by reading best practice guidelines (Lewis, 2011). Application uncertainties do not in fact regard the current simulation. Software errors are similarly assumed not to occur in CFX, as it is a validated and established CFD package (ANSYS, 2009; Lewis, 2011). Nevertheless, the problems encountered with Lewis et al.s (2010) wedge show that there are serious issues with the rigid body solution.

79

12.0- Appendix III-CFX Geometry description


The equations of motion for free-fall, ignoring viscosity effects, are those of a constant acceleration motion, where the acceleration is given by the gravitational constant. For zero initial velocity, which is the case of the wedge at its starting position, they can be simplified and rearranged to yield (Bedford & Fowler, 2007): . In equations (22) and (23), m and , (22) (23) m are the initial experimental

wedge drop height and is the initial wedge drop height of the geometry file, which can be deduced from table 1 in section 3.2. The results from equations (22) and (23) are also contained in table 1 in section 3.2. Table 14 presents the location of the six points on the wedge where pressure gauges have been applied to by Lewis, et al. (2010). The actual position of the monitor points employed in CFX differs slightly, as mentioned in section 5.0. The reference system used for their coordinates has its origin coinciding with the lowest point of the wedge and the orientation of the axes can be deduced from figure 56. Table 16 presents the coordinates of the monitor points used in the wedge by Aarsnes (1996). The system of reference is the same as per the other wedge. Nevertheless, it should be noted that a rotation of 20.3 is required to obtain the final position of the body. Table 16 contains the coordinates of the points belonging to one of the two halves of the ship-like section surface (Aarsnes, 1996). In addition, the four monitor points where pressure gauges had been applied to are highlighted as well. The origin of the reference system and the axes orientations can be seen from figure 58 in this case.

80

TABLE 14: LOCATION OF THE SIX MONITOR POINTS FOR THE WEDGE BY LEWIS, ET AL. (2010) Point P1 P2 P3 P4 P5 P6 x (m) 0.0453154 0.0906308 0.1359462 0.1812616 0.2265769 0.2718923 y (m) 0.0211309 0.0422618 0.0633927 0.0845237 0.1056546 0.1267855 z (m) 0 0 0 0 0 0

TABLE 15: LOCATION OF THE FIVE MONITOR POINTS FOR THE WEDGE BY AARSNES (1996) PRIOR TO ROTATION. Monitor point P1 P2 P3 P4 P5 x (m) 0.0259808 0.0519615 0.0779423 0.103923 0.1299038 y (m) 0.015 0.03 0.045 0.06 0.075 z (m) 0 0 0 0 0

TABLE 16: COORDINATE POINTS EMPLOYED TO PRODUCE THE SURFACE OF THE BOW SECTION. THE LOCATION OF THE FOUR MONITOR POINTS IS ALSO HIGHLIGHTED. x (m) 0 0.005 0.016 0.0233 0.0289 0.0327 0.0356 0.0412 0.0495 0.0613 0.075 0.0941 0.1142 0.1397 0.1561 0.16 y (m) -0.24 -0.24 -0.2279 -0.2106 -0.1923 -0.1704 -0.1503 -0.1302 -0.1138 -0.1021 -0.0895 -0.0769 -0.0644 -0.0483 -0.0366 0 z (m) 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 P4 P3 P2 P1

Table 17 presents the position of the monitor points in CFX that describe the lower portion (up to 10% of the depth) of the bow section. The reference system is the same
81

as per table 16 and figure 58. They have been used to produce table 18, which contains the deadrise angle and length of each segment the two-dimensional ship-like section surface has been discretised into in order to calculate the vertical impact force per unit length acting against the lower portion of the section adopting Stavovy and Chuangs (1976) approach. Table 18 has been incorporated within table 6 in section 4.0.
TABLE 17: LOCATION OF THE MONITOR POINTS EMPLOYED BY CFX AND THE EMPIRICAL PROCEDURE BY STAVOVY AND CHUANG (1976) TO CALCULATE THE FORCE PER UNIT LENGTH AGAINST THE LOWER PORTION OF THE BOW SECTION. Monitor Point 1 2 3 4 5 6 7 x (m) 0.0000 0.0050 0.0114 0.0151 0.0182 0.0212 0.0102 y (m) -0.2400 -0.2400 -0.2360 -0.2300 -0.2230 -0.2160 -0.2376 z (m) 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000

TABLE 18: LENGTH AND DEADRISE ANGLE OF THE SEGMENTS THE LOWER PART OF THE BOW SECTION HAS BEEN DISCRETISED INTO. () Segment 1-2 Segment 2-3 Segment 3-4 Segment 4-5 Segment 5-6 Segment 1-7 0 32.16705 58.06334 66.1135 66.80141 13.24052 Length (m) 0.005 0.007513 0.00707 0.007656 0.007616 0.012044

Figures 53 to 55 present the section view of the wedge adopted by Lewis, et al. (2011), the one employed by Aarsnes (1996) and the bow section (Aarsnes, 1996). dimensions as well as the location of the monitor points are clearly evident. The

82

FIGURE 53: SECTION VIEW OF THE WEDGE BY LEWIS, ET AL. (2010).

FIGURE 54: SECTION VIEW OF THE WEDGE BY AARSNES (1996).

FIGURE 55: SECTION VIEW OF THE BOW SECTION.

83

Figures 56 to 58 present the domain dimensions for the wedge adopted by Lewis, et al. (2010), the one used by Aarsnes (1996) and the bow section (Aarsnes, 1996). For both wedges, the tank dimensions have been matched. However, the second wedge presents a greater domain height in accordance with good CFD practice (Bressloff, 2011). Indeed, it is believed that the small domain size for the first wedge can be one of the causes of its problems, as mentioned in section 6.1. The size of the domain for the bow section is purely based on good CFD practice (Bressloff, 2011).

FIGURE 56: DOMAIN DIMENSIONS FOR THE WEDGE BY LEWIS, ET AL. (2010).

FIGURE 57: DOMAIN DIMENSIONS FOR THE WEDGE BY AARSNES (1996).

84

FIGURE 58: DOMAIN DIMENSIONS FOR THE BOW SECTION. ONLY THE CASE WITH NO HEEL ANGLE IS PRESENTED.

85

13.0- Appendix IV-Further discussion of the empirical approaches


Von Karman (1929, cited in Bertram, 2000) considered a two-dimensional wedge entry problem onto the calm-water surface assuming very small elevation and negligible gravity effects for impulsive impacts. Thus, the added mass can be approximated as , where calculated from is the half width of the wet area of the wedge implicitly , where is the final velocity of the wedge (Bertram,

2000). Momentum conservation yields the equality of the momentum before the impact ( , where
is the initial flow velocity) and the sum of the wedge momentum (

and added mass momentum (

) (Bertram, 2000). Hence, the impact pressure (since

a two-dimensional problem is considered) can be calculated as (Bertram, 2000):

(24)

Von Karmans approach is known as momentum impact, since it is based on momentum conservation. As it neglects the water surface elevation, the added mass and impact load are underestimated, particularly for small deadrise angles (Bertram, 2000). The slamming peak impact pressures for the wedge calculated through this method are reported in table 19. Due to their very small magnitude as compared to the other empirical, numerical and experimental results, Von Karmans (1929, cited in Bertram, 2000) method has been neglected.
TABLE 19: PEAK PRESSURE MAGNITUDE CALCULATED THROUGH THE THEORY BY VON KARMAN (1929, CITED IN BERTRAM, 2000) FOR THE WEDGE FOR BOTH IMPACT SPEEDS. Low velocity Peak Impact Pressure (Pa) 2.07 High Velocity 3.07

It is interesting to note that von Karmans (1929, cited in Bertram, 2000) is the only method that employs the rigid body mass. Only the lower of the two masses (23.4 kg) has been employed, as the results from section 4.0 are also compared against this case only in section 5.5.

86

According to Wagners (1932, cited in Bertram, 2000) approach, the velocity potential and its derivative with repect to on the plate is analytically given by

equations (25) and (26) respectively (Bertram, 2000): { for for for for ; (25)

(26)

The wave surface elevation and the half width of the wetted surface area calculated from the time integral of equation (26). The free-surface elevation to the bottom of the wedge can be thus written as (Faltinsen, 1993):

can be relative

(27)

This needs to be equal to , which is the vertical coordinate of a point on the wedge relative to the bottom of the wedge. Hence, it follows that (Faltinsen, 1993): . (28)

According to Wagner (1932, cited in Faltinsen, 1993), the maximum hydrodynamic pressure coefficient very close to | | is: . Hence, in section 4.0. or ( 29)

, which is also reported as equation (5)

87

14.0- Appendix V-Preliminary studies for the CFX simulations


All simulations in this section have been performed on computers at the University of Southampton having Intel Core-Duo processors, with 4 Gb RAM. The total simulation time is 0.15 s and the time step intervals 0.005 s for all simulations in this section. Only the wedge geometry for the experiments by Lewis, et al. (2010) has been considered. In addition, as it can be deduced from the types of elements contained in tables 20 to 22, during this preliminary stage of the project, a major error has been made during the mesh generation process, in that a three-dimensional rather than two-dimensional flow has been actually simulated, despite the geometry and boundary conditions employed are the same as per section 3.2 and Appendix III. This has been caused by the selection of the wrong meshing approach in CFX-Mesh, namely the Delunay/advancing front rather than the extruded 2-D mesh as required. Nevertheless, the three-dimensional effects in the flow are believed to be of minor importance due to the very small thickness of the domain as compared to its other dimensions. Therefore, the quality of these results is still considered valid. In addition, the input mass in the simulations is actually twice the experimental one (Lewis, et al., 2010) in order to dispose with the oscillations in the solver. Hence, although the magnitude of the experimental pressure should be greater, according to the trend shown in section 5.1 and Appendix VI, the qualitative nature of the data is considered valid for estimating the best models, as all simulations already underestimate the results.

14.1- Results from the turbulence and interphase models analysis


This section contains the preliminary results from investigations in the influence of turbulence and interphase models. Indeed, for the same domain size and mesh density, three turbulence models and two interphase models have been adopted. These can be seen in table 20, where the methods are described in sections 11.4 and 11.6 respectively (Appendix II). It is evident that the simulation names are also based on the turbulence and interphase models.

88

TABLE 20: NUMBER OF ELEMENTS (TETRA. ARE TETRAHEDRA, PYR. PYRAMIDS), TOTAL WALL-CLOCK TIME, TURBULENCE AND INTERPHASE MODELS FOR THE SIMULATIONS PERFORMED TO STUDY THE INFLUENCE OF TURBULENCE AND INTERPHASE MODELS FOR THE WEDGE BY LEWIS, ET AL. (2010). No. nodes 10064 10064 10064 10064 No. Tetra. 22757 22757 22757 22757 No. Pyr. 20 20 20 20 No. Prisms 3260 3260 3260 3260 Tot Elements 26037 26037 26037 26037 Turbulence model k- k- sst sst Interphase model freesurface mixture mixture freesurface freesurface

Simulation

T.w.t. (s)

k-, f.s. k-, mix. sst, mix. sst, f.s.

926 976 961

SSG, f.s.

10064

22757

20

3260

26037

1000

SSG

Table 20 also contains the total wall clock time, which corresponds to the total computation time, in order to assess the computational load exerted by the various turbulence and interphase models. Unfortunately, data for the first case ( , free-

surface) has been lost, but it is believed to be less than 926 s from the trend of the results for the shear stress transport turbulence model. Figures 59 to 61 show the pressure measurements for all cases presented in table 20 at P1 and P4, P2 and P5 and P3 and P6 respectively. The results are compared against experimental measurements for validation, where the experimental data corresponds to the case of low impact speed and mass performed by Lewis, et al. (2010), although the experimental mass is in fact smaller.

89

30000 25000 20000 Pressure (Pa) 15000 10000 5000 0 0 -5000 0.02 0.04 Time (s) 0.06 0.08 P1-k-e, f.s. P1-k-e, mix. P1-sst, mix. P1-sst, f.s. P1-SSG, f.s. P4-k-e, f.s. P4-k-e, mix. P4-sst, mix. P4-sst, f.s. P4-SSG, f.s. P1-Experiment P4-Experiment

FIGURE 59: VARIATION OF PRESSURE AT P1 AND P4 WITH TIME FOR DIFFERENT TURBULENCE AND INTERPHASE MODELS FOR THE SAME MESH AND WEDGE GEOMETRY.

25000 P2-k-e, f.s. 20000 P2-k-e, mix. P2-sst, mix. 15000 Pressure (Pa) P2-sst, f.s. P2-SSG, f.s. 10000 P5-k-e, f.s. P5-k-e, mix. 5000 P5-sst, mix. P5-sst, f.s. 0 0 -5000 0.02 0.04 0.06 0.08 P5-SSG, f.s. P2-Experiment P5-Experiment Time (s)

FIGURE 60: VARIATION OF PRESSURE AT P2 AND P5 WITH TIME FOR DIFFERENT TURBULENCE AND INTERPHASE MODELS FOR THE SAME MESH AND WEDGE GEOMETRY.

90

20000 P3-k-e, f.s. 15000 P3-k-e, mix. P3-sst, mix. P3-sst, f.s. Pressure (Pa) 10000 P3-SSG, f.s. P6-k-e, f.s. 5000 P6-k-e, mix. P6-sst, mix. P6-sst, f.s. 0 0 0.02 0.04 0.06 0.08 P6-SSG, f.s. P3-Experiment P6-Experiment -5000 Time (s)

FIGURE 61: VARIATION OF PRESSURE AT P3 AND P6 WITH TIME FOR DIFFERENT TURBULENCE AND INTERPHASE MODELS FOR THE SAME MESH AND WEDGE GEOMETRY.

14.2- Results from a study in the influence of different mesh stiffness models
Two different mesh stiffness models have also been used in the preliminary phases of the project to assess their suitability. These are namely the Increase near Small Volume (ISV), which is automatic and sets the mesh stiffness to be related to the inverse of the volume of each element through an exponential model, and the 1/volcvol model, which considers the mesh stiffness as inversely proportional to the volume of each cell. The latter method is recommended by ANSYS (2009) for a similar problem (Tutorial 32). The mesh parameters and total computation time are presented in table 21 for two simulations performed with the same domain size, mesh density, turbulence and interphase models (shear stress transport and mixture respectively).

91

TABLE 21: NUMBER OF ELEMENTS (TETRA. ARE TETRAHEDRA, PYR. PYRAMIDS), TOTAL WALL-CLOCK TIME AND MESH STIFFNESS MODEL FOR THE WEDGE BY LEWIS, ET AL. (2010) FOR AN ANALYSIS OF THE MESH STIFFNESS MODELS. No. nodes 5904 5904 No. Tetra. 15792 15792 No. Pyr. 430 430 No. Prisms 20 20 Tot Elements 16242 16242 T.w.c. t. (s) 725 724 Max. E.F. 797 797

Simulation

Mesh stiffness Increase near Small Volumes 1/volcvol

ISV volcvol

Figures 62 to 64 show the pressure measurements for the cases presented in table 21 at all monitor points. The very low magnitude of the pressure is caused by the absence of the inflated boundary layers next to the top and bottom surfaces of the domain.

30000 25000 20000 Relative Pressure (Pa) P1-inflated, ISV 15000 10000 5000 0 0 -5000 0.01 0.02 0.03 Time (s) 0.04 0.05 0.06 P1-Inflated, volcvol P4-Inflated, ISV P4-Inlfated, volcvol P1-Experimental P4-Experiment

FIGURE 62: VARIATION OF PRESSURE AT P1 AND P4 WITH TIME FOR DIFFERENT MESH STIFFNESS MODELS FOR THE SAME MESH AND WEDGE GEOMETRY.

92

25000

20000

Relative Pressure (Pa)

15000

P2-Inflated, ISV P2-Inflated, volcvol

10000

P5-Inflated, ISV P5-Inflated, volcvol

5000

P2-Experiment P5-Experiment

0 0 -5000 0.01 0.02 0.03 0.04 0.05 0.06

Time (s)

FIGURE 63: VARIATION OF PRESSURE AT P2AND P5 WITH TIME FOR DIFFERENT MESH STIFFNESS MODELS FOR THE SAME MESH AND WEDGE GEOMETRY.

20000

15000

P3-Inflated, ISV P3-Inflated, volcvol P6-Inflated, ISV

Relative Pressure (Pa)

10000

5000

P6-Inflated, volcvol P3Experiment

0 0 0.02 0.04 0.06 0.08

P6Experiment

-5000

Time (s)

FIGURE 64: VARIATION OF PRESSURE AT P3 AND P6 WITH TIME FOR DIFFERENT MESH STIFFNESS MODELS FOR THE SAME MESH AND WEDGE GEOMETRY.

93

14.3- Results from a study in domain size


Table 22 contains the number of elements, simulation time and total computation time for three simulation files. The mesh density tends to be constant, except for the last file, which presents a coarser mesh to improve mesh deformation. The main difference lies in the domain size and initial height of the wedge above the free surface, which is also presented in table 22. The initial input downward velocity is calculated using the equations of free fall motion (22) and (23) in Appendix III. The water depth is the same as per figure 56 in Appendix III for all cases, but a minimum distance of 25 cm is left between the top of the wedge and that of the domain for simulation files 0.00033 and 0.30. Simulation 0.10 presents the same domain size as figure 56 in Appendix III. In addition, it is the same simulation as the , free-surface case presented in table 20

and figures 59 to 61 in section 14.1. In fact, all simulations in this section have the same turbulence, interphase and mesh stiffness models, namely the and 1/volcvol.
TABLE 22: NUMBER OF ELEMENTS (TETRA. ARE TETRAHEDRA, PYR. PYRAMIDS), TOTAL WALL-CLOCK TIME, INITIAL DROP HEIGHT AND SPEED FOR A STUDY IN DOMAIN SIZE AND INITIAL HEIGHT ABOVE THE WATER SURFACE OF THE WEDGE BY LEWIS, ET AL. (2010). No. nodes 9534 10064 7659 No. Tetra. 20528 22757 12147 No. Pyr. 50 20 20 No. Prisms 3114 3260 3195 Tot Elements 23692 26037 15362 Simulation time (s) 0.1 0.2 0.25 T.w.c. t. (s) 667 1831 In. H. (cm) 0.033 10 30 In. sp. (m/s) -1.562 -1.401 -1.213

, free surface

Simulation 0.00033 0.10 0.30

Figures 65 to 67 present the pressure variation with time at all monitor points for the simulations presented in table 22. The graphs include the experimental measurements for validation despite the difference in mass. The plots associated with the simulation 0.30 needed cutting at 0.05 s after the time of impact because of the start of a serious oscillatory behaviour.

94

30000 25000 20000 Pressure (Pa) 15000 10000 5000 0 0 -5000 0.01 0.02 0.03 Time (s) 0.04 0.05 0.06 P1-0.00033 P1-0.10 P1-0.30 P4-0.00033 P4-0.10 P4-0.30 P1-Experiment P4-Experiment

FIGURE 65: VARIATION OF PRESSURE AT P1 AND P4 WITH TIME FOR DIFFERENT DROP HEIGHTS.

25000

20000 P2-0.00033 15000 Pressure (Pa) P2-0.10 P2-0.30 10000 P5-0.00033 P5-0.10 5000 P5-0.30 P2-Experiment 0 0 -5000 0.02 0.04 0.06 0.08 P5-Experiment

Time (s)

FIGURE 66: VARIATIONS OF PRESSURE AT P2 AND P5 WITH TIME FOR DIFFERENT DROP HEIGHTS.

95

20000

15000 P3-0.00033 P3-0.10 Pressure (Pa) 10000 P3-0.30 P6-0.00033 5000 P6-0.10 P6-0.30 P3-Experiment 0 0 0.02 0.04 0.06 0.08 P6-Experiment

-5000

Time (s)

FIGURE 67: VARIATION OF PRESSURE AT P3 AND P6 WITH TIME FOR DIFFERENT DROP HEIGHTS

14.4- Discussion of the preliminary simulations


As it can be seen from figures 59 to 61 in section 14.1, all turbulence models results in very similar pressure data at all points, so much that the data for the shear stress transport turbulence model with a free surface is actually shadowed by the curve corresponding to the SSG model (with a maximum pressure at P1 of 15084 against 15060 Pa). Hence, the selection should be purely based on the computational resources they need. Indeed, while so simple a mesh causes no problems, finer meshes could result in enormous wall-clock times if the turbulence models (as well as other parameters) are not selected wisely. Therefore, the turbulence model should be

selected for all simulations in this project, in particular for the three-dimensional ones. Nevertheless, it was discovered that this and the SSG turbulence model would result in overflow errors in the solver, particularly for very fine meshes. These are believed to be caused by the low values (beyond the prescribed limits (section 3.3)) in the mesh

around the rigid body, which are the result of mesh deformation. Indeed, the scalable wall function cannot model the flow accurately in these regions, whereas the shear stress transport, which relies on the automatic wall function, as described in section
96

11.4, causes no solver issues, although it requires a greater number of coefficient loops for convergence. Hence, the claim by ANSYS (2009) of having solved all issues with the viscous sub-layer for the turbulence model through the scalable wall function

is in fact incorrect. For this reason, the shear stress transport has been selected in all subsequent simulations (section 5). The mixture model is found to underestimate the pressure magnitude for all monitor points as compared with the free surface interphase model. In addition, although it is supposed to be simpler, it actually results in longer computation time, as it can be seen in table 20. Therefore, the free surface has been chosen as the standard interphase model for all subsequent simulations (section 5). From figures 62 to 64, there is no appreciable difference in the pressure data for the Increase near Small Volumes and the 1/volcvol mesh stiffness models. Similarly, the computation time requirements are almost identical. It has been decided to employ 1/volcvol mesh stiffness model in all subsequent simulations (section 5) because of ANSYS (2009) recommendations. Nevertheless, from the comparison of figures 59 to 61 and 62 to 64, it is clear that the inflated boundary layer, despite its small thickness, as described in section 3.3, must be applied to the top and bottom surfaces of the domain as well. This has been done in all subsequent simulations (section 5). From the study in the domain size (figures 65 to 67 in section 14.3), it is clear that the larger the initial height of the wedge above the water level, the greater the magnitude in the pressure measurements. It should be noted that obtaining the actual experimental drop height in the simulations was not possible due to mesh folding. Nevertheless, for simulation 0.3 the pressure time distribution is rather different from the experimental measurements and the pressure magnitude at P4, P5 and P6 tends to be overestimated. Nevertheless, care should be taken in that the experimental results refer to a different mass setting than the numerical results, as already mentioned. However, the reason for discarding the file 0.3 is the significant oscillatory behaviour that occurs in the solver, associated with the exceedance of the prescribed maximum expansion factor limit. Hence, an initial file of 0.1 m in the geometry file seems a reasonable compromise. For this reason, it has been selected in all subsequent wedge simulations (section 5.1).

97

15.0- Appendix VI-Experimental results for the wedge by Lewis, et al. (2010)
Figures 68 to 71 are taken from the paper by Lewis, et al. (2010). They present the pressure variation at the six pressure gauges, whose location can be seen in figure 53 in Appendix III, for the low body mass and impact speed, the low wedge mass and high impact speed, the high mass and low impact speed and the high mass and impact speed cases respectively. They are presented as a means of validation for the numerical results presented in section 5.1, where the zero impact time occurs at t=0.07 s for the low impact speed simulations and t=0.051 s for the high impact speed cases. These times correspond to the simulation time, as per figures 5 to 7 in section 5.1.

FIGURE 68: PRESSURE TRACE FOR A 0.50 M DROP, WITH A WEDGE MASS OF 23.4 KG, TAKEN FROM FIGURE 18 IN THE PAPER BY LEWIS, ET AL. (2010).

98

FIGURE 69: PRESSURE TRACE FOR A 0.50 M DROP, WITH A WEDGE MASS OF 33.4 KG, TAKEN FROM FIGURE 19 IN THE PAPER BY LEWIS, ET AL. (2010).

FIGURE 70: PRESSURE TRACE FOR A 0.75 M DROP, WITH A WEDGE MASS OF 23.4 KG, TAKEN FROM FIGURE 20 IN THE PAPER BY LEWIS, ET AL. (2010).

99

FIGURE 71: PRESSURE TRACE FOR A 0.75 M DROP, WITH A WEDGE MASS OF 23.4 KG, TAKEN FROM FIGURE 21 IN THE PAPER BY LEWIS, ET AL. (2010).

Figure 72 shows the visual description of the jet flow around the wedge during its free drop, presented by Lewis (Lewis, 2011) in his thesis. On the left side of the figure, high-frequency images taken during the experiment (Lewis, et al., 2010) are presented.

FIGURE 72: EXPERIMENTAL AND NUMERICAL FLOW DESCRIPTION AROUND THE WEDGE, TAKEN FROM LEWIS (2011).

100

16.0- Appendix VII- Visual Results


16.1- Wedge by Lewis, et al. (2010)
Figures 73 to 75 show the contours of the water volume fraction and relative pressure for t=0 s for the simulations on Lewis, et al.s (2010) wedge. Figure 73 is particularly interesting for understanding the meaning of mesh density thanks to the enabled wireframe display that shows the position of each node. Figure 73A corresponds to file a (very coarse grid) and figure 73B to file c (fine grid), as per section 5.1. Figure 75 is an enlargement of figure 73A. From this as well as figure 73, it is evident that coarser meshes present compression layer of greater thickness (a comrpession layer of two was selected in order to allow mixture of the two fluids near the interphase), since this is selected automatically by CFX based on the size of the elements lying on the interphase between the two fluids. This is considered the main cause for the smaller pressure magnitude obtained for the coarser meshes in section 5 and it can be grouped into the discretization errors mentioned in section 11.7. In figure 74, it can be seen that the opening boundary layers selected for the two side walls have resulted in the correct hydrostatic pressure distribution.

FIGURE 73 A,B: WATER VOLUME FRACTION FOR THE WEDGE BY LEWIS, ET AL. (2010) AT T=0S FOR TWO DIFFERENT MESH DENSITIES.

101

FIGURE 74: HYDROSTATIC PRESSURE DISTRIBUTION AROUND THE WEDGE BY LEWIS, ET AL. (2010).

FIGURE 75: NODE DENSITY AROUND THE WEDGE BY LEWIS, ET AL. (2010) FOR A COARSE MESH.

Figures 76 to 78 show the pressure and water volume fraction around Lewis, et al.s (2010) wedge for low mass and low initial velocity at simulation time 0.065 s, corresponding to the time of impact, 0.08, and 0.1 s. The grid used has medium node density (file c_0.001 in section 5.1). The extent of the compression layer becomes significant away from the body, especially if compared with section 16.2, 16.3 and 16.4. This is believed to be caused by the smaller magnitude of the domain size for this simulation (wedge by Lewis, et al. (2010)) and in particular the shorter domain height.
102

FIGURE 76 A,B: PRESSURE AND WATER VOLUME FRACTION CONTOURS AROUND THE WEDGE AT T=0.065 S.

FIGURE 77 A,B: PRESSURE AND WATER VOLUME FRACTION CONTOURS AROUND THE WEDGE AT T=0.080 S.

103

FIGURE 78 A,B: PRESSURE AND WATER VOLUME FRACTION CONTOURS AROUND THE WEDGE AT T=0.100 S.

Figures 79 and 80 show the pressure distribution around the wedge at 0.115 and 0.116 s simulation time. It is very interesting to note the negative pressure that is formed under the wedge in figure 80, which is one of the causes for the low pressure magnitude and oscillations in the result in section 5.1. In figure 80B, the flow at t=0.116 s is displayed.

FIGURE 79: PRESSURE DISTRIBUTION AROUND THE WEDGE AT T=0.115 S.

104

FIGURE 80 A,B: PRESSURE AND WATER VOLUME FRACTION CONTOURS AROUND THE WEDGE AT T=0.116 S.

Figures 81 and 82 present the pressure distribution around the wedge at t=0.149 and 0.150 s respectively. In both figures, the same pressure limits are used for the countour. While in the former case the pressure below the wedge is very high and the pressure above it is atmospheric, in the second case this situation is reversed. In fact, in figure 82 there is a very high suction pressure below the rigid body. These are examples of the oscillatory behaviour of the rigid body solution, described in section 3.2 and 11.2, and confirm the unreliability of the numerical results of section 5.1, as explained in section 6.1. Figure 81B shows the flow around the wedge at t=0.149 s.

105

FIGURE 81 A,B: PRESSURE AND WATER VOLUME FRACTION CONTOURS AROUND THE WEDGE AT T=0.149 S.

FIGURE 82: PRESSURE DISTRIBUTION AROUND THE WEDGE AT T=0.150 S.

106

16.2- Wedge by Aarsnes (1996)


In figure 83, the node density and water surface can be seen at simulation time t=0s for the Aarsnes (1996) wedge. In figure 84, the impact instant is shown, which

corresponds to t=0.02 s. Figures 85 and 86 present the water volume fraction and the pressure contours at t=0.025 s, where air entrapment is evident under the lower side of the wedge. This causes hydroelastic effects against the side of the wedge (Faltinsen, 2000), which are one of the causes for the discrepancy between the experimental and numerical impact pressure and force data, as discussed in section 6.2. For simulating hydroelastic effects, the CFD solver should be coupled with a FEA solver (Maki, et al., 2011). However, this approach has been considered too complex by the student and has thus not been adopted in the current project.

FIGURE 83: NODE DENSITY AND WATER VOLUME FRACTION AROUND THE WEDGE BY AARSNES (1996) AT T=0 S.

FIGURE 84: WATER VOLUME FRACTION FOR THE WEDGE BY AARSNES (1996) AT IMPACT TIME.

107

FIGURE 85: AIR ENTRAPMENT UNDER THE WEDGE BY AARSNES (1996) AT T=0.025 S.

FIGURE 86: PRESSURE DISTRIBUTION AROUND THE WEDGE BY AARSNES (1996) AT T=0.025 S.

The flow in figure 87 corresponds to a time of t=0.05 s. It is interesting to note the two areas where an air gap has been left near the hull, namely next to the upper wedge side and against the bulwark of the lower side (the part above the chine). The air gaps could result in a possible capsize of the ship if rotations were enabled. This is particularly important for planing hulls, which present similar sections. However, more simulations and especially experiments are required to validate this behaviour.

108

FIGURE 87: FLOW AROUND THE WEDGE BY AARSNES (1996) AT T=0.100 S.

16.3- Two-dimensional ship-like section


Figures 88 and 89 show the relative pressure distribution and the water and air density fractions at t=0 s and t=0.175 s respectively for the straight ship-like section and low impact speed.

FIGURE 88 A,B: HYDROSTATIC PRESSURE AND WATER VOLUME FRACTION CONTOURS FOR THE BOW SECTION WITH NO HEEL ANGLE AND LOW IMPACT SPEED AT T=0 S.

109

FIGURE 89 A,B: PRESSURE AND WATER VOLUME FRACTION CONTOURS FOR THE BOW SECTION WITH NO HEEL ANGLE AND LOW IMPACT SPEED AT T=0.175 S.

Figures 90 to 94 present the relative pressure contours and the flow around the bow section for a mesh of medium density and the low impact speed, for t=0.012 (actual time of impact), 0.05, 0.1, 0.125 s and 0.15 s.

110

FIGURE 90 A,B: PRESSURE AND WATER VOLUME FRACTION CONTOURS FOR THE BOW SECTION WITH NO HEEL ANGLE AND LOW IMPACT SPEED AT IMPACT TIME (T=0.012S).

FIGURE 91 A,B: PRESSURE AND WATER VOLUME FRACTION CONTOURS FOR THE BOW SECTION WITH NO HEEL ANGLE AND LOW IMPACT SPEED AT T=0.050 S.

111

FIGURE 92 A,B: PRESSURE AND WATER VOLUME FRACTION CONTOURS FOR THE BOW SECTION WITH NO HEEL ANGLE AND LOW IMPACT SPEED AT T=0.100 S.

FIGURE 93 A,B: PRESSURE AND WATER VOLUME FRACTION CONTOURS FOR THE BOW SECTION WITH NO HEEL ANGLE AND LOW IMPACT SPEED AT T=0.125 S.

112

FIGURE 94 A,B: PRESSURE AND WATER VOLUME FRACTION CONTOURS FOR THE BOW SECTION WITH NO HEEL ANGLE AND LOW IMPACT SPEED AT T=0.150 S.

The extent of mesh deformation can be understood from figure 95, where the wireframe is displayed for both t=0 and t=0.175 s.

FIGURE 95 A,B: NODE POSITION AND FLOW AROUND THE BOW SECTION WITH NO HEEL ANGLE AND LOW IMPACT SPEED AT T=0 AND T=0.175 S.

113

16.4- Heeled ship-like section


Figures 96 to 101 show the flow and pressure around the bow section with a 9.8 heel angle and low impact speed at simulation time t=0.1, 0.15 and 0.175 s. The extent of the compression layer is very evident for these simulations.

FIGURE 96: FLOW AROUND THE BOW SECTION AT T=0.1 S WITH A 9.8 HEEL ANGLE AND LOW IMPACT SPEED.

FIGURE 97: PRESSURE DISTRIBUTION AROUND THE BOW SECTION AT T=0.1 S FOR A 9.8 HEEL ANGLE AND LOW IMPACT SPEED.

114

FIGURE 98: FLOW AROUND THE BOW SECTION AT T=0.15 S WITH A 9.8 HEEL ANGLE AND LOW IMPACT SPEED.

FIGURE 99: PRESSURE DISTRIBUTION AROUND THE BOW SECTION AT T=0.15 S FOR A 9.8 HEEL ANGLE AND LOW IMPACT SPEED.

115

FIGURE 100: FLOW AROUND THE BOW SECTION AT T=0.175 S WITH A 9.8 HEEL ANGLE AND LOW IMPACT SPEED.

FIGURE 101: PRESSURE DISTRIBUTION AROUND THE BOW SECTION AT T=0.175 S FOR A 9.8 HEEL ANGLE AND LOW IMPACT SPEED.

Figures 102 to 109 show the flow and pressure around the bow section for a 28.3 heel angle at simulation time t=0.07, 0.1, 0.125, 0.15 s. The clear void next to the upper

116

surface caused by the suction pressure compares well with the visual results of section 16.2.

FIGURE 102: FLOW AROUND THE BOW SECTION AT T=0.07 S WITH A 28.3 HEEL ANGLE AND LOW IMPACT SPEED.

FIGURE 103: PRESSURE DISTRIBUTION AROUND THE BOW SECTION AT T=0.07 S FOR A 28.3 HEEL ANGLE AND LOW IMPACT SPEED.

117

FIGURE 104: FLOW AROUND THE BOW SECTION AT T=0.1 S WITH A 28.3 HEEL ANGLE AND LOW IMPACT SPEED.

FIGURE 105: PRESSURE DISTRIBUTION AROUND THE BOW SECTION AT T=0.1 S FOR A 28.3 HEEL ANGLE AND LOW IMPACT SPEED.

118

FIGURE 106: FLOW AROUND THE BOW SECTION AT T=0.125 S WITH A 28.3 HEEL ANGLE AND LOW IMPACT SPEED.

FIGURE 107: PRESSURE DISTRIBUTION AROUND THE BOW SECTION AT T=0.125 S FOR A 28.3 HEEL ANGLE AND LOW IMPACT SPEED.

119

FIGURE 108: FLOW AROUND THE BOW SECTION AT T=0.15 S WITH A 28.3 HEEL ANGLE AND LOW IMPACT SPEED.

FIGURE 109: PRESSURE DISTRIBUTION AROUND THE BOW SECTION AT T=0.15 S FOR A 28.3 HEEL ANGLE AND LOW IMPACT SPEED.

120

16.5- Three-dimensional ship-like section


Figure 110 shows the section view of the node position around the bow section for the three-dimensional simulation for the low impact speed at t=0.175 s, where the greatest mesh deformation has occurred. The node motion has also caused a deformation of the flow around the section, as it is clearer in figure 111, which is an enlargement of figure 110 without the wireframe displayed. These are the visual consequences of the

maximum expansion factor exceedance, presented in table 12 in section 5.4.

FIGURE 110: NODE POSITION IN THE SECTION VIEW AROUND THE THREE-DIMENSIONAL BOW SECTION AT T=0.175 S.

FIGURE 111: ENLARGEMENT OF THE FLOW AROUND THE TIP OF THE THREE-DIMENSIONAL BOW SECTION AT T=0.175 S.

121

Figures 112 and 113 present the profile view of the pressure and water volume fraction contours at t=0.175 s for the simulation x10, presented in section 5.4.

FIGURE 112: PRESSURE DISTRIBUTION AROUND THE THREE-DIMENSIONAL WEDGE AT T=0.175 S IN THE PROFILE VIEW OF FILE X10.

FIGURE 113: FLOW AROUND THE THREE-DIMENSIONAL WEDGE AT T=0.175 S IN THE PROFILE VIEW OF FILE X10.

122

Figures 114 and 115 present the profile view of the pressure and water volume fraction contours at t=0.175 s for the simulation x1.5, presented in section 5.4.

FIGURE 114: PRESSURE DISTRIBUTION AROUND THE THREE-DIMENSIONAL WEDGE AT T=0.175 S IN THE PROFILE VIEW OF FILE X1.5.

FIGURE 115: WATER VOLUME FRACTION CONTOUR AROUND THE THREE-DIMENSIONAL WEDGE AT T=0.175 S IN THE PROFILE VIEW OF FILE X1.5.

123

17.0- Appendix VIII- CEL equations


The following lines are an extract from the .ccl equation files used for the simulations involving the wedge by Lewis, et al. (2010). These equations are included within the report to show the appearance of the CEL language. In order for equations to be imported into CFX, this format is necessary.
LIBRARY: CEL: EXPRESSIONS: AirVolFra=1.0-WatVolFra denAir=1.185 [kg m^-3] denWat= 997.0 [kg m^-3] hwat=-0.16 [m] hystaticP=(denWat-denAir)*-g*WatVolFra*(y-hwat) WatVolFra=step((hwat-y)/1.0 [m]) END END END COMMAND FILE: Version = 12.0.1 END

The variable hwat represents the level of the water in respect to the origin of the reference system. As it can be seen from figures 56, 57 and 58 in Appendix III, this is negative and it actually depends on the domain size and system of reference (the wedge by Aarsnes (1996) and the bow section will have different values for hwat). However, a positive value could be selected instead in any other geometry file if the system of reference were inverted, without affecting the validity of the above equations. Nevertheless, in this case the hydrostatic pressure would require a change in sign. The use of the step functions in the determination of the water volume fraction was necessary. In fact, another approach was used initially, where the volume fraction of
124

the water was expressed as the ratio of the volume displaced by the water over the total volume of the tank minus that of the wedge. However, this was found to result in each element in the cell having the same volume fraction of air and water, causing a mixture of the two fluids within the whole domain. Use of the step functions is recommended in the ANSYS (2009) modelling guide. The values of the water and air densities have been copied from the ANSYS materials library (2009) from the fluids Air at 250C and Water at 250C.

125

You might also like