You are on page 1of 9

Ould el Moctar

e-mail: ould.el-moctar@gl-group.com

Wave Load and Structural Analysis for a Jack-Up Platform in Freak Waves
This paper analyzed the effects of freak waves on a mobile jack-up drilling platform stationed in exposed waters of the North Sea. Under freak wave conditions, highly nonlinear effects, such as wave run-up on platform legs and impact-related wave loads on the hull, had to be considered. Traditional methods based on the Morison formula needed to be critically examined to accurately predict these loads. Our analysis was based on the use of advanced computational uid dynamics techniques. The code used here solves the Reynolds-averaged NavierStokes equations and relies on the interface-capturing technique of the volume-of-uid type. It computed the two-phase ow of water and air to describe the physics associated with complex free-surface shapes with breaking waves and air trapping, hydrodynamic phenomena that had to be considered to yield reliable predictions. Lastly, the nite element method was used to apply the wave-induced loads onto a comprehensive nite element structural model of the platform, yielding deformations and stresses. DOI: 10.1115/1.2948952 Keywords: freak waves, jack-up platform, CFD technique, FEM modeling, wave loads, deformations, stresses

Thomas E. Schellin1
e-mail: thomas.schellin@gl-group.com

Thomas Jahnke
e-mail: thomas.jahnke@gl-group.com Germanischer Lloyd, Hamburg 20459, Germany

Milovan Peric
CD-adapco Nrnberg, Nrnberg 90402, Germany e-mail: milovan.peric@de.cd-adapco.com

Introduction
Last September, hurricane Katrina scoured the Gulf of Mexico, wreaking havoc among mobile drilling rigs and production platforms. Since then, it has been clear that the energy industry must make offshore structures more hurricane resistant. The key challenge in preventing the failures of marine structures is understanding the storms forces that must be taken into account during design. When structures were rst situated in shallow waters, they were designed for a 25 year return period. Today, offshore structures must withstand the forces generated by a 100 year storm. So-called freak waves are extreme waves that marine structures need to be designed for if those waves occur on site. These design conditions not only account for the characteristics of the threatening wave, but also for highly nonlinear run-up on the legs and impact-related forces on the hull. Consequently, accurate assessment of such loads is essential for the design of marine structures. Of late, modern computational uid dynamic CFD tools, validated against experimental measurements, successfully predicted impact-related slamming loads on ships 1 . The application of these methods needs to be incorporated in the analysis of wave forces on marine structures. Wave impact is a strongly threedimensional nonlinear phenomenon that is sensitive to the relative motion between the structure and the water surface. Furthermore, the inuence of the compressibility of water and air pockets may have to be accounted for. Methods that directly solve the Reynolds-averaged NavierStokes RANS equations, including the two-phase ow of water and air, are better able to describe the physics associated with impact-related wave loads. Today, methods based on codes that implement interface-capturing techniques of the volume-of-uid VOF type promise to be most suitable for computing complex wave-structure interactions. In this paper, we investigated a typical self-elevating jack-up
Corresponding author. Contributed by the Ocean Offshore and Arctic Engineering Division of ASME for publication in the JOURNAL OF OFFSHORE MECHANICS AND ARCTIC ENGINEERING. Manuscript received July 2, 2007; nal manuscript received March 3, 2008; published online March 12, 2009. Assoc. Editor: Ge George Wang. Paper presented at The 26th International Conference on Offshore Mechanics and Arctic Engineering OMAE2007 , San Diego, CA, June 1015, 2007.
1

drilling unit subject to freak waves of varying height. Traditional methods, such as the IMO Code for Offshore Drilling Units 2 and the SNAME Guidelines for Mobile Jack-Up Units 3 , are based on the Morison formula to determine wave loads. Here, wave loads were analyzed using CFD techniques. We employed code COMET 4 as it proved to be most suitable to describe the physics associated with complex free-surface ow shapes such as leg run-up, breaking waves, and air trapping. The use of advanced CFD tools can be relied on only if the results have been validated. No experimental data were available for the platform under consideration; however, impact-related wave slamming loads on ships predicted by the employed numerical method were validated against experimental model test measurements 1,5 . As its application for the prediction of wave loads is not restricted to ships, this numerical technique is suitable also for the wave load analysis of offshore platforms. Of interest here are pressure dominated loads obtained by integration of local pressures over structural plate elds, and not local peak pressures acting over small control volume face areas. To determine the structures stress level, we used the nite element method FEM to predict the effects of the computed waveinduced loads on the structures strength. We started with a comprehensive nite element structural model of the platform Fig. 1 and deployed Version 10.0A1 of the commercial code ANSYS 6 to perform a transient nonlinear nite element analysis of the rigs structure subject to the considered wave conditions. By comparing results with the structures rule based design capability 7 , we determined the reserve strength capacity still available under freak wave conditions 8 .

Jack-Up Rig
The investigated three-legged self-elevating mobile drilling unit is of triangular construction, containing tubular steel legs at each corner that can be jacked up or down by electrohydraulic machinery. The elevating mechanism is of the rack-and-pinion type, featuring an automatic position lock on power failure. The units three legs are spaced 39.0 m apart, forming an equilateral triangle. The FEM model in Fig. 1 shows the overall conguration of the MAY 2009, Vol. 131 / 021602-1

Journal of Offshore Mechanics and Arctic Engineering Copyright 2009 by ASME

Downloaded 07 Apr 2009 to 202.118.71.160. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Fig. 1 Global structural FEM model of the jack-up rig

rig in its elevated drilling position. The short front wall of the hull, situated ahead of the foremost leg, designated the units bow. Table 1 lists principal particulars. Our investigation considered the rig operating in the North Sea on a 33.5 m water depth location. Realistic loading conditions comprised gravity loading together with relevant environmental loading. The units Operating Manual specied a corresponding signicant wave height of 6.24 m, a current velocity of 0.51 m / s, and a wind speed of 58 knots. Current and wind loads were assumed acting collinearly with the waves. The wind force of 4348 kN was determined according to the Rules of Germanischer Lloyd 7 . This force acted at a height of 23.3 m above calm water level.

Wave Load Prediction


In high and steep waves, the ow around the unit is associated with wave run-up along the legs and impact-related loads on the underside of the hull. Both of these highly nonlinear phenomena may signicantly increase local wave loads that, in turn, lead to decreased global stability of the unit as well as increased local damage to the hull. The code COMET solves the RANS equations and relies on the interface-capturing method of the VOF type to determine the inTable 1 Principal particulars of the jack-up rig

Molded hull length Molded hull breadth Molded hull depth Elevated height above calm water level Leg diameter Overall leg length Unsupported leg length Gross tonnage Net tonnage

46.0 m 47.6 m 5.5 m 13.5 m 3.66 m 64.0 m 48.3 m 4033 tons 3209 tons

terface between water and air. This technique accounts for highly nonlinear wave effects in that it computes the two-phase ow of water and air to describe the physics associated with complex free-surface shapes with breaking waves and air trapping. Conservation equations for mass and momentum in their integral form serve as the starting point. The solution domain is subdivided into a nite number of control volumes that may be of arbitrary shape. The integrals are numerically approximated using the midpoint rule. The mass ux through the cell face is taken from the previous iteration, following a simple Picard iteration approach. The remaining unknown variables at the center of the cell face are determined by combining a central difference scheme CDS with an upwind differencing scheme UDS for the convective terms. The diffusive terms are discretized using CDS. The CDS employs a correction to ensure second-order accuracy for an arbitrary cell. A second-order CDS can lead to unrealistic oscillations if the Peclet number exceeds 2.0 and large gradients are involved. On the other hand, an UDS is unconditionally stable, but leads to higher numerical diffusion. To obtain a good compromise between accuracy and stability, the schemes are blended. In the neighborhood of the platform, the blending factor is chosen between 0.8 and 0.9. The Euler implicit method is used to integrate in time. This rst-order fully implicit approximation is stable. Pressure and velocity are coupled by a variant of the SIMPLE algorithm 9 . The system of equations is under-relaxed to dampen changes between iterations. All equations except the pressure correction equations are under-relaxed using a relaxation factor 0.8. The pressure correction equations are under-relaxed using a relaxation factor between 0.2 and 0.4 for unsteady simulations, nding in each case a suitable compromise between stability and convergence speed. The two-uid ow is modeled by a two-phase formulation of the governing equations 10 . No explicit free surface is dened during the computations, and overturning breaking waves as well as buoyancy effects of trapped air are accounted for. The spatial distribution of each of the two uids is obtained by solving an additional transport equation for the volume fraction of one of the uids. To accurately simulate the convective transport of the two immiscible uids, the discretization must be nearly free of numerical diffusion and must not violate the boundedness criteria. For this purpose, the high resolution interface capturing HRIC scheme is used 11 . The scheme is a nonlinear blend of upwind and downwind discretizations, and the blending is a function of the distribution of the volume fraction and the local Courant number. The free surface is typically smeared over one to two control volumes. Fluid structure interaction effects are presently not accounted for, i.e., the body is assumed to be rigid, and the uid is assumed to be viscous and incompressible. The Courant number was almost always less than 0.25 for all simulations. For the grid density and time step used here, the numerical diffusion turned out to be sufciently small. To assure numerical convergence, at every time step the computations were performed until the normalized residuals of all equations to be solved were reduced by three to four powers of ten. Typically, up to 12 outer iterations for each time step were necessary for convergence. To assure time convergence, the ow was simulated until a periodic solution was attained. We started the analysis by investigating the effect of the design wave specied according to the criterion for the rig under the survival condition. A deterministic long-crested wave of 11.6 m height and 13.0 s period, propagating towards the port bow in the direction of 60 deg to the longitudinal axis of the hull, described this condition. Our VOF computations were not deterministic. Therefore, the wave prole represented a regular wave only at the inlet boundaries of the computational grid, and the propagating wave reached its intended height and period when its crest arrived at the rigs legs. An analysis based on the SNAME 3 guidelines showed that these wave parameters resulted in the highest base shear on the structure. We then considered a series of three epiTransactions of the ASME

021602-2 / Vol. 131, MAY 2009

Downloaded 07 Apr 2009 to 202.118.71.160. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Fig. 4 Screen shot of rig in 19.9 m wave

Fig. 2 Numerical grid on legs and hull of the rig and on the ocean bottom

sodic freak waves having the same period and direction. Their heights were 15.8 m, 19.9 m, and 23.7 m. Design guidelines designate the 23.7 m wave height to be the limiting wave height before breaking.

Flow Simulation
An unstructured grid, comprising about 2.5 106 hexahedral control volumes, surrounded the drilling rig, as shown in Fig. 2. Denser H-grids dened control volumes around the legs, in the vicinity of the hull, and in the neighborhood of the free surface, as shown in Fig. 3. To avoid ow disturbances at the outer grid boundaries of the numerical domain, domain boundaries extended 6.1 hull lengths upstream, 15.6 hull lengths downstream, and 7.6 hull lengths to each side of the platform. The upper boundary was located 2.8 hull lengths above the ocean bottom; the lower boundary was the ocean bottom itself. Grid density was higher in the ow region ahead of the structure to resolve the incident waves, whereas aft of the structure the grid became coarse to dampen the waves. No signicant wave reection occurred at the outlet boundary. We specied our numerical grids on the basis of past experience with similar computations performed earlier 1,5,12,13 . Consequently, we were able to narrow down the discretization errors. In the longitudinal direction, 70 control volumes extending over one wavelength dened our grid. In the vertical direction, 20 control volumes extending over the wave height dened our grid. In the vertical direction, this grid was locally rened near the free surface Fig. 3 . The rigs orientation relative to the incident waves represented the direction of wave propagation investigated here. Waves were generated at the front of the computational domain, specied as the inlet boundary of known velocities and known volume fraction distributions dening water and air regions. The back of the

computational domain was specied as a pressure boundary. Velocities that initialized the ow eld arose from the superposition of wave particle velocities and current speed. Second-order Stokes wave theory described the initial wave particle velocities. The entire ow eld was initialized by the hydrostatic pressure. On the rig surfaces, a no-slip condition was enforced on uid velocities and on the turbulent kinetic energy. The wake ow boundary was specied as a zero-gradient pressure boundary hydrostatic pressure . On average, the time step size was chosen such that about 1000 time steps dened each wave period. The momentum equations were discretized using 85% central differences and 15% upwind differences. About 50 h of CPU time on six standard processors was required for one wave period. We used turbulence models with wall functions, but the ow was pressure dominated because here we dealt with wave-induced loads. Based on the assumed current velocity past the platforms legs, the ow was characterized by a Reynolds number of about 2 106 and a Froude number of about 0.08. However, the inuence of friction and turbulence on wave loads was small and, consequently, did not signicantly contribute to overall loading. Of course, the form drag caused by the ow past the platforms legs was accounted for. The boundary conditions of this eld method inuenced the solutions. Wave kinematics at the inlet boundary did not signicantly affect the wave formation. The resulting wave kinematics accounted for the nite water depth and caused wave crests to be higher than wave troughs. This is seen in Fig. 4, showing a sample screen shot of the simulated 19.9 m wave advancing from left to right past the structure. The screen shots in Figs. 5 and 6 show the prole of the 23.7 m wave passing by the structure just as the wave breaks, demonstrating that the VOF method used was capable of simulating this situation. Wave run-up on legs extended up to the underside of the hull deck, but the undisturbed wave crest did not. Wave run-up caused impact-related slamming pressures to act on the legs and the underside of the hull. This was the case for all waves considered. For the 19.9 m wave, Fig. 7 shows a pressure distribution on the rig corresponding to the time of maximum base shear and overturning moment. Only the two highest waves caused wave run-up to reach the hull. Time histories in

Fig. 5 Screen shot of rig in 23.7 m wave

Fig. 3 Higher local grid density for wave-structure interaction

Fig. 6 Simulated wave prole of 23.7 m wave

Journal of Offshore Mechanics and Arctic Engineering

MAY 2009, Vol. 131 / 021602-3

Downloaded 07 Apr 2009 to 202.118.71.160. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Fig. 7 Pressure distribution on rig in 19.9 m wave

Fig. 8 show local slamming pressures as they occurred at the top of the two legs facing the wave crest at locations directly underneath the hull deck. The symbol HW in this and other gures stands for wave height. Although the resulting peak pressures were high for all wave heights, attaining maxima exceeding 200 kPa, hull structural plate scantlings of this rig are more than adequate to withstand these pressures. Grid size as well as time step size did not signicantly affect the resulting wave-induced forces and moments, because these values were obtained by integrating the computed pressures. Some of these pressures were characterized by high pressure peaks. However, these pressure peaks acted only locally over the small control volume face area. To obtain impact-related wave forces, we integrated the computed pressures over a plate eld with an area that was much larger than the control volume face area. We did not perform a sensitivity study specically for this platform, because previous investigations 1,5,14 demonstrated the validity of this procedure. After each time step, pressures acting on the structure were integrated to yield forces caused by the incident waves and the superimposed current. The total horizontal force on all parts of the rig is known as the base shear. Summing moments of these forces about a horizontal axis that runs normal to the direction of wave propagation and through the center of the tip of the most highly

loaded starboard aft leg resulted in the overturning moment. For the four wave heights considered, Figs. 9 and 10 show time histories of the structures base shear and overturning moment, respectively, and Fig. 11 shows time histories of the base shear of one of the two legs facing the wave crest. Negative values of base shear represent horizontal forces acting in the direction of wave propagation; positive values, forces acting against the direction of

Fig. 9 Time histories of base shear

Fig. 8 Slamming pressures on underside of hull deck

Fig. 10 Time histories of overturning moment

021602-4 / Vol. 131, MAY 2009

Transactions of the ASME

Downloaded 07 Apr 2009 to 202.118.71.160. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Fig. 13 Comparative results of base shear for the 11.6 m wave Fig. 11 Time histories of base shear on a single leg

wave propagation. The corresponding values of overturning moment are based on moment arms measured positively upward from the ocean bottom. Base shear and overturning moment are the dominant safety criteria against sliding and capsizing of the rig, respectively. From Figs. 9 and 10, we see that only for the 11.6 m wave height are the computed time histories almost sinusoidal with negative nonzero mean values. For higher waves, the functions are characterized by peaks in the negative direction, corresponding to wave crests attacking the structure. Peak absolute values increase nonlinearly with wave height, as seen by the plots of maximum values of base shear and overturning moment versus wave height shown in Fig. 12. However, this nonlinearity is less pronounced between the two highest waves, indicating that the increase in the wetted areas of the legs decreased when the highest wave started to break. Positive peak values of these time histories are nearly equal for all four wave heights, indicating that the wave troughs were nearly the same for all four waves. Proles of waves when reaching the legs of the rig no longer represented regular waves, because the VOF computations were not deterministic. Only at inlet boundaries did wave proles represent regular waves. Peak values of base shear and overturning moment occurred at the same instants of time. This was expected because the overturning moment was a direct result of multiplying the horizontal loads with their respective moment arms. Wave loading on the two legs simultaneously facing the wave crest was split nearly equally between each of these legs, as seen by the time histories of base shear on one of these legs Fig. 11 . This result was expected because we simulated long-crested waves. We then calculated wave loads using the Morison formula, with coefcients as recommended in the SNAME design guidelines for mobile jack-up units 3 . For the four wave heights considered,

Figs. 1320 show time series of the resulting base shear and overturning moment together with comparative results from our RANS computations. Regarding base shear, both methods predicted nearly equal peak values for the three lower wave heights of 11.6 m, 15.8 m, and 19.9 m Figs. 1315 . Only for the highest breaking wave 23.7 m did peak values from the Morison formula exceed peak values from RANS by about 15% Fig. 16 . Regarding overturning moment, both methods yielded nearly equal peak values only for the 19.9 m wave Fig. 19 . For the 11.6 m wave, peak values from the Morison formula are about 15% greater than peak values from RANS Fig. 17 , whereas for the 15.8 m and 23.7 m waves, the peak values from the Morison formula are about 25% greater than values from the RANS Figs. 18 and 20, respectively . The basics of the two approaches resulted in these differences. Only the RANS technique was able to account for wave kinematics associated with breaking waves and, therefore, predicted wave loads more realistically, especially for the two highest waves considered. The deviation between the time histories from the two approaches at higher waves conrmed this.

Fig. 14 Comparative results of base shear for the 15.8 m wave

Fig. 12 Maxima of base shear and overturning moment versus wave height

Fig. 15 Comparative results of base shear for the 19.9 m wave

Journal of Offshore Mechanics and Arctic Engineering

MAY 2009, Vol. 131 / 021602-5

Downloaded 07 Apr 2009 to 202.118.71.160. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Fig. 16 Comparative results of base shear for the 23.7 m wave

Fig. 19 Comparative results of overturning moment for the 19.9 m wave

Morison results were steady as they repeated themselves from cycle to cycle, indicating that incident Morison waves were regular. This was not the case for RANS results, where the differences in peaks from wave to wave of up to 30% were observed. The variation of predicted impact pressures caused these differences. The VOF computations, simulating breaking waves splashing against a solid structure, were not deterministic. Our RANS method did not resolve small-scale uctuations that turbulence models consider; however, our transient simulations did resolve uctuations that took place on a larger time scale. Only if the free surface would have remained smooth, would our RANS simulations have produced a perfectly periodic behavior. In principle, we solved the unsteady Reynolds-averaged NavierStokes URANS equations. This procedure was justied because the frequencies of the resolved unsteadiness, here caused by wave breaking and splashing, were sufciently far away from the higher frequencies of turbulent uctuations. Generally, Morison peak values turned out to be larger than peak values from RANS computations. This was because the less

accurate Morison formulation had to rely on predetermined hydrodynamic force coefcients and harmonic wave proles.

Strength Analysis
To investigate the structures strength, we performed a transient nite element analysis of the rigs structure subject to wave conditions, current, and wind forces. The analysis accounted for nonlinear large structural deformations, but the material properties were assumed to be linear. For this strength analysis, we considered only the three lower wave heights and not the 23.7 m wave 8 . This was done because it turned out that the 19.9 m high wave was already a breaking wave. The global FEM comprised stiffened plates with eight-node shell elements and three-node beam elements to idealize hull, jack houses, living quarters, and legs Fig. 1 . Modeling of the structure with stiffened shells and beams simulated the complete stiffness distribution of the rig. Concentrated masses modeled the helideck and the cantilevered drill oor. Shell elements modeled plane plating of hull bottom, side shells, transoms, decks, and internal structural bulkheads. Shell elements also modeled the oors. Beam elements idealized webframes, deck girders, and pillars. Beam elements also modeled the longitudinal framing system of decks, bottom, bulkheads, and shell elements. Proles were grouped together to limit the mesh size. Shell elements modeled the cylindrical shells of the legs and the spud wells. For the ring stiffeners inside the legs, we used beam elements. Plane shell elements with a mean height of the tooth prole idealized the elevator racks. For the legs, we used a nonlinear material model because high stresses were expected to occur at the positions of the lower leg guides. Pin supports idealized the boundary condition of the leg tips on the mudline. Nonlinear spring elements surrounding the legs were located at the levels of the leg guides, namely, at the hull bottom and at the top of the jack houses. These spring ele-

Fig. 17 Comparative results of overturning moment for the 11.6 m wave

Fig. 18 Comparative results of overturning moment for the 15.8 m wave

Fig. 20 Comparative results of overturning moment for the 23.7 m wave

021602-6 / Vol. 131, MAY 2009

Transactions of the ASME

Downloaded 07 Apr 2009 to 202.118.71.160. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Fig. 23 Bending stresses at Point 2 for 15.8 m and 19.9 m waves

Fig. 21 Stress distribution at Points 1 and 2 for 15.8 m wave

ments transmitted mainly compressive loads. Truss elements, extending from the top of the legs to the three pinions on each side of the legs, served as vertical supports between hull and legs. When transforming the wave and current induced loads onto the structure, we determined spatial mean wave-induced loads by integrating computed local wave-induced pressures over the areas of the nite elements that idealized the exposed surfaces of the structures legs and hull. The resulting loads were transformed to equivalent nodal forces on the FEM of the structure. To minimize transient effects, a sinusoidal ramp function was used to gradually apply wave- and current-induced loads for the rst half of a wave period 6.5 s . A consequence of using this ramp function was that the time histories of stresses in Figs. 22 and 23 no longer corresponded to the time histories of base shear in Fig. 9. We used three criteria to assess the global safety of the unit: rst, the resulting stress level in the shell underneath the hull for the most loaded leg; second, the vertical holding capacity of the jacking gears; third, the rigs overturning stability. The rigs safety margin or reserve strength was based on safety factors for the survival condition according to Ref. 3 . For allowable stresses, the safety factor of equivalent stress against yield was 1.1; against axial/bending stress, 1.25. For the required leg holding capacity of the jacking system, the safety factor against maximum leg design load was 1.3. The safety factor against overturning was 1.3. To analyze the resulting stresses in the most loaded starboard aft leg, we selected two points, both situated on the outer shell of this leg. As shown in Fig. 21, Point 1 was located on the outer shell directly underneath the hull; Point 2 was located 1.5 m be-

low Point 1. At point 1, equivalent von Mises stresses were superimposed on local stresses caused by the action of the leg guides. At this point, the overloaded structure was most likely to experience plastic deformation. At Point 2, the predicted stresses represented the undisturbed global bending stresses in the leg. Figures 22 and 23 show the resulting time histories of stresses in the most loaded aft starboard leg. It is seen that the higher 19.9 m wave caused disproportionately larger stresses that extended into the plastic range. However, even for the highest wave, the resulting stresses were within allowable limits according to classication society rules 7 . The total required holding capacity for a selected load was evaluated by summing the internal forces of all truss elements representing the jacking system of one leg. Again, the most loaded leg turned out to be the aft starboard leg. As a representative sample, Fig. 24 shows the computed time series of the jacking gear forces for the 19.9 m wave height. In this gure, symbols PS-Aft, SB-Aft, and CL-Fwd designate the rigs portside aft leg, starboard side aft leg, and centerline forward leg, respectively. We investigated overturning in the sense of rule based requirements, that is, when the reaction force of at least one leg loses its downward support loads on the mudline. As a representative sample, Fig. 25 shows the resulting time series of vertical leg tip reaction forces of all three legs for the 19.9 m wave height. In this gure, symbols PS-Aft, SB-Aft, and CL-Fwd again designate the units portside aft leg, starboard side aft leg, and centerline forward leg, respectively. The severe environmental loads investigated here caused sizable global deformations of the rig. To illustrate, Fig. 26 shows the computed time series of deformations for the center bottom node of the hull. The maximum deection in the horizontal x-direction turned out to be about 1.2 m as the 19.9 m wave went by the rig. Symbols x, y, and z in this gure refer to the axes of the

Fig. 22 Equivalent stresses at Point 1

Fig. 24 Jacking system forces for 19.9 m wave

Journal of Offshore Mechanics and Arctic Engineering

MAY 2009, Vol. 131 / 021602-7

Downloaded 07 Apr 2009 to 202.118.71.160. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Fig. 25 Vertical leg tip reaction forces for 19.9 m wave

rig-bound right-handed coordinate system, with the x-axis pointing towards the bow, the y-axis to port, and the z-axis vertically upward. At the outer shell directly underneath the hull of the most loaded starboard aft leg Point 1 , the material deformed plastically. The computed time series reect this behavior, considering the rig had a natural frequency of about 5.0 s. Because the resulting deformations were large, it would have been appropriate to account for these deformations when applying loads to the FEM. Although plastic deformation occurred in the most loaded leg for the 19.9 m wave height Fig. 22 , the overall structural integrity was still assured as far as leg strength is concerned. If this case were classied as an accident, plastic deformation would have been acceptable. The jacking system of the rig was designed for a maximum holding capacity of 32,000 kN per leg. This value was exceeded already for the 19.9 m wave Fig. 24 . With the rule based safety factor of 1.3 and the maximum design leg load for the survival condition 11.6 m wave height , the required holding capacity of the jacking system is 1.3 21,345 kN= 27,750 kN for this rig. This means that the jacking system turned out to be the weak link in the design of this rig. With a mass of 5438 tons and a lever of about 21 m, the uprighting moment was about 1120 MN m. The maximum overturning moment, equal to the sum of the maximum wave- and currentinduced moment for the 23.7 m wave 330 MN m, see Fig. 10 and the wind-induced moment 250 MN m , was 580 MN m. Hence, the safety factor against overturning turned out to be 1120 MN m / 580 MN m = 1.9, well in excess of the rule required safety factor of 1.3.

waves, we demonstrated that this rig possesses sufcient reserve strength to withstand freak wave conditions in excess of the required design wave specied for its survival condition. Our analysis, however, considered a considerably higher hull elevation than the required elevation based on the rule based minimum airgap. We specically selected this higher hull elevation for our analysis because most operating assignments for this kind of rig call for a high hull elevation. Our comparative results of base shear and overturning moment of the platform subject to freak waves revealed that predictions based on the use of the Morison formula differed by not more than 25% from predictions obtained from CFD techniques. Peak values of overturning moment differed more than peak values of base shear. This was brought about by the more accurate distribution of CFD based wave forces acting on the platform legs, especially for the higher breaking waves. These comparative results demonstrated the general usefulness of the Morison formula approach to assess strength related safety aspects although only for cases of high hull elevation. At lower hull elevations with waves attacking the hull directly, it would have been necessary to rely on methods such as this CFD technique to predict wave loads. This CFD approach is just as effective to reliably predict aerodynamic loads. Because of large moment arms, wind loads acting on the above-water parts of the rig can be critical. They can cause wind-induced moments that are of the same order of magnitude as wave-induced moments. However, as wind forces were separately determined, we did not employ this method to compute these wind forces. The complexity of modeling wave loads and their interaction with a nite element structural model forced us to make simplifying assumptions. For instance, we considered waves to be longcrested although this was not necessary as we employed this CFD technique routinely to simulate sea loads on many ships. Also, we relied on linear strain displacement relationships. However, the structural nite element analysis accounted for a bilinear material behavior and compression only spring elements. Only the treatment of long-crested waves was a conservative assumption. The structural analysis resulted in large horizontal hull deformations that would have justied accounting for these deformations when applying loads to the nite element structural model.

Conclusions
The combined use of CFD and FEM techniques proved to be effective to assess the structural integrity of the investigated jack-up mobile offshore drilling platform. By analyzing freak

References
1 el Moctar, O., Brehm, A., and Schellin, T. E., 2004, Prediction of Slamming Loads for Ship Structural Design Using Potential Flow and RANSE Codes, Proceedings of the 25th Symposium on Naval Hydrodynamics, St. Johns. 2 International Maritime Organization IMO , 1989, Code for the Construction and Equipment of Mobile Offshore Drilling Units, IMO MODU Code. 3 Society of Naval Architects and Marine Engineers SNAME , 2002, Guidelines for Site Specic Assessment of Mobile Jack-Up Units, Technical & Research Bulletin 5-5A, Jersey City, 1st ed., Rev. 2. 4 CD-adapco, 2002, User Manual COMET, Version 2.0, Nrnberg. 5 Schellin, T. E., and el Moctar, O., 2006, Numerical Prediction of ImpactRelated Wave Loads on Ships, ASME J. Offshore Mech. Arct. Eng., 129, pp. 3947. 6 Mller, G., and Groth, C., 2002, Practical Application of FEM Code ANSYS, Vol. 1: Basics, 7th ed., Expert Verlag, Renningen, in German. 7 Germansicher Lloyd, 2007, Rules for Classication and Construction, IV Industrial Services, 6 Offshore Technology, Hamburg. 8 Schellin, T. E., Jahnke, T., and Knzel, J., 2007, Consideration of Freak Waves for Design of a Jack-Up Structure, Proceedings of Offshore Technology Conference, Houston, TX, Paper No. OTC-18465. 9 Ferziger, J., and Peric, M., 1996, Computational Methods for Fluid Dynamics, Springer-Verlag, Berlin. 10 Lafaurie, B., Nardone, C., Scardovelli, R., Zaleski, S., and Zanetti, G., 1994,

Fig. 26 Rigid body hull deformation for 19.9 m wave

021602-8 / Vol. 131, MAY 2009

Transactions of the ASME

Downloaded 07 Apr 2009 to 202.118.71.160. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Modeling Merging and Fragmentation in Multiphase Flows With SURFER, J. Comput. Phys., 113, pp. 134147. 11 Muzaferija, S., and Peric, M., 1998, Computation of Free Surface Flows Using Interface-Tracking and Interface-Capturing Methods, Nonlinear Water Wave Interaction, O. Marenholtz and M. Markiewicz, eds., Computational Mechanics, Southampton, Chap. 3. 12 el Moctar, O., Schellin, T. E., and Priebe, T., 2006, CFD and FE Methods to

Predict Wave Loads and Ship Structural Response, Proceedings of the 26th Symposium on Naval Hydrodynamics, Rome, Italy. 13 el Moctar, O., and Bertram, V., 2001, RANSE Simulations for High-Fn, High-Re Free Surface Flows, Proceedings of the Fourth Numerical Towing Tank Symposium, Hamburg, Germany. 14 Klemt, M., 2004, Motion Simulation of Floating Bodies in Viscous Flow, Ph.D. thesis, Hamburg University, Hamburg.

Journal of Offshore Mechanics and Arctic Engineering

MAY 2009, Vol. 131 / 021602-9

Downloaded 07 Apr 2009 to 202.118.71.160. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

You might also like