You are on page 1of 75

FINAL TECHNICAL REPORT

Confidential

CONTRACT N : PROJECT N : TITLE :

BRPR-CT97-0513 BE97-4375 Design for Structural Safety under Extreme Loads (DEXTREMEL)

PROJECT CO-ORDINATOR : PARTNERS :

Germanischer Lloyd Technical University of Denmark Maritime Research Institute Netherlands National Technical University of Athens Sirehna University of Newcastle upon Tyne Astilleros Espaoles

REFERENCE PERIOD : STARTING DATE : 1 Dec. 1997

FROM 1 Dec. 1997 TO 30 Nov. 2000 DURATION : 36 Months 23 January 2001 Project funded by the European Community either under the IMT/SMT Programmes (1994-1998)

Date of issue of this report :

DEXTREMEL (BE97-4375) Final technical report page 2 / 75

Table of Contents
1 2 3 EXECUTIVE SUMMARY OBJECTIVES OF THE PROJECT MARKET AND INTRODUCTION TO RISK ASSESSMENT 3.1 MARKET REVIEW 3.2 REFERENCE SHIP 3.3 RISK ASSESSMENT 4 SCIENTIFIC AND TECHNICAL DESCRIPTION OF THE RESULTS 4.1 COLLISION DAMAGE 4.2 GROUNDING DAMAGE 4.3 BOW DOOR DAMAGE 4.4 GREEN WATER DAMAGE 4.5 STRUCTURAL INTEGRITY 5 LIST OF DELIVERABLES 3 4 5 5 6 7 9 9 21 27 39 53 63 64 64 65 66 67 68 68 68 69 70 71 72 75

6 COMPARISON OF INITIALLY PLANNED ACTIVITIES AND WORK ACTUALLY ACCOMPLISHED. 6.1 6.2 6.3 6.4 7 7.1 7.2 7.3 8 9 10 TASK 1: COLLISION AND GROUNDING LOADS TASK 2: BOW DOOR LOADS TASK 3: GREEN WATER LOADS TASK 4: RESIDUAL STRUCTURAL STRENGTH OF DAMAGED SHIP STRUCTURES PERFORMANCE OF CONSORTIUM ORGANIZATION AND COMMUNICATION LIST OF CONTACT PERSONS FOR FOLLOW UP

MANAGEMENT AND CO-ORDINATION ASPECTS

RESULTS AND CONCLUSIONS ACKNOWLEDGEMENTS REFERENCES

APPENDIX

DEXTREMEL (BE97-4375) Final technical report page 3 / 75

Executive Summary

Results of the European research project DEXTREMEL are presented. The objective of the research work was to meet the industrial need for design methods to ensure structural safety and subsequently passenger safety of ships in adverse environments. Four extreme load scenarios were identified and they were investigated in detail for a fast conventional RoRo passenger ship on route from Cadiz to the Canary Islands. A new method was developed to predict collision probability and damage distributions taking into account actual traffic data. A second method was developed to predict grounding damage distributions for variable structural arrangements of the double bottom. Extreme loads on and possible failure of the bow door were investigated by means of model tests and a new computational method to predict wave loads on bow doors. Tests comprised a model with elastically mounted bow segment to measure the forces and an innovative approach to visualize wave contours at the bow using a laser augmented system. To study effects of green water, wave heights at the fore deck and loads on the front bulkhead of the super structure were measured. In addition, the laser visualization method was used to collect information on wave contours and wave propagation over the fore deck. Finally, the residual structural strength of the damaged ship in waves was numerically and experimentally investigated. All results were integrated into risk analyses that were carried out for collision and grounding scenarios, bow door failure and failure of the super structure due to green water. It was shown that risk analyses are feasible for the selected events. However, computational effort was large and some key elements for the estimation of consequences were missing. Research needs were identified to fully exploit this performance based approach.

DEXTREMEL (BE97-4375) Final technical report page 4 / 75

Objectives of the project

The future of European shipyards depends on their ability to produce new generation ships at competitive costs. With ever increasing traffic density, safety is increasingly becoming key for competitiveness. DEXTREMEL was submitted to respond to the industrial need for design methods to ensure structural safety and subsequently passenger safety of ships in adverse environments. Recent disasters of passenger RoRo ferries highlighted the demand for a rational performance-based approach to safeguard passengers and the environment in case of accidents. The safety of RoRo ferries was identified as a key priority by the European Commission and the Maritime Industry R&D Masterplan. Main objective of DEXTREMEL was to provide numerical methods to predict the residual structural strength of damaged RoRo passenger ferries in extreme conditions. Based on the simulation of a ship's behavior under realistic operating conditions, these methods aim at the rapid evaluation of new concepts and configurations. The results of DEXTREMEL will enable the quantification of risks associated with (total) loss of structural integrity due to extreme and / or accidental loads, such as collision and grounding loads, wave loads acting on bow doors, and loads on deck structures caused by green water shipping. In addition, the formulation of prenormative guidelines will contribute to new rules concerning RoRo passenger ferry safety in extreme conditions. To meet the outlined objectives, research activities were divided into four tasks. Task 1 focused on the development of software to predict damages due to collision and grounding taking into account actual traffic situations. Loads acting on bow doors were experimentally investigated in task 2. Numerical prediction was enabled by newly developed software. Task 3 concentrated on loads on deck structures due to shipping of green water. Model test were performed to validate numerical predictions. Extreme load scenarios were used in task 4 to study loss of structural integrity. Again, model test results were used to verify newly developed software. The establishment of prenormative guidelines consolidated the main findings of the project.

DEXTREMEL (BE97-4375) Final technical report page 5 / 75

Market and introduction to risk assessment

3.1 Market review


The prospects for the sea transportation, specifically the vehicle and passenger ferry segment, appear promising if they are exploited in a proper way, i.e., the inefficiencies in the transportation chain are tackled to have a competitive integrated segment. From the side of the industry, in this case the shipbuilders, the RoPax segment contains high added value products, see Fig. 3.1.1. Therefore a good position in this market segment is of paramount importance, specially for European shipbuilders with higher labor costs. The achievement of this objective should be considered a strategic task. As an example, a RoPax having a length of 188 m, a service speed of 27.5 knot, and a capacity to transport 2300 passengers, has an estimated price of Euro 115m.

Ships Added Value Curve


Pr/t
Naval Ships Cruisers & Ro-Pax Ro-Ro/Car Carriers Portacontainers LNG/LPG Tankers

Chemical Tanker/ Reefers

Bulk-Carriers

Lightship W eight, LSW (t)

Figure 3.1.1: Ship's relative value The RoRo and RoPax market is characterized by shortsea trade (more of the 85 % of the fleet in the European case), progressive aging of the fleet (the average age has increased to 19.7 years in 1997), and inefficiency in their exploitation (a shortsea trader in its present design spends 30% to 40% of its life in port). In addition, sea transport is the cheapest part in the transport chain and the weakest because of the long transit time. The European market for RoRo and RoPax vessels is (still) considered to have potential because shortsea shipping is an efficient remedy for traffic breakdowns and environmental damage, sea transportation is a highly economical mean and is readily available, minimum energy consumption per t / km ensures environment friendliness, and factors as regularity, punctuality and quality can be offered by shortsea shipping. If the general set-up for shortsea shipping can be improved, it will be possible to shift trade currents from the road to the sea. Figure 3.1.2 provides the prognosis of transportation in Europe up to the year 2010, considering the four surface transportation segments: Rail, Inland Water Transport (IWT), Sea and Road, versus their corresponding volume in Millions of tonnes. The added sea trade market potential is contained in the difference between the road and sea curves. To exploit the European sea transport market potential, the maritime transportation chain has to be improved, e.g. reducing the transport time through hinterland transport to the port of departure, transshipment in the port of departure, transshipment in the port of arrival, and hinterland transport at the port of arrival. Further improvements are needed to reduce costs and to increase flexibility and reliability of the transport.

DEXTREMEL (BE97-4375) Final technical report page 6 / 75 As a result, a new RoPax based transport design should focus on the following aspects: incorporate systems for fast cargo handling operation, potential for higher velocity, and systems in general to cope with a more dense traffic. In addition, safety must be a major driving factor in designing a new RoPax vessel. Also, increasing attention to environmental aspects should be paid. The method should focus to enhance the maritime safety including protection of life, health, the maritime environment and property by using risk assessments.

Prognosis of Transportation
Volume up to the year 2010

600 500 400 300 200 100 0


2010 2005 2000

Road Sea IWT*


1995

Rail
1990 1985

Figure 3.1.2: Market potential of transportation for different categories

3.2 Reference ship


The selected reference ship, the MS Dextra, represents a new class of vessel: a generic platform that fulfils the demands of the market potential above described, summarized regarding ship design in higher speed and safety. Main particulars are summarized in Table 3.2.1. A side view of the reference vessel is presented in Fig. 3.2.1, and the sections for two variants are given in Fig. 3.2.2. Bow geometry above the design waterline was modified to study effects of flare angle on loads. Table 3.2.1: Main particulars of reference ship Length between PP [m] Breadth [m] Draft [m] Service speed [Knots] Displacement [t] Deadweight [t] 173.0 26.0 6.5 28 16800 4500

DEXTREMEL (BE97-4375) Final technical report page 7 / 75

Figure 2.3.1: Side view of reference ship MS Dextra

Figure 3.2.2: Sections of reference ship. Original (right) and modified form (left). Two routes in European waters were selected as typical for the reference vessel. Particulars are given in Table 3.2.2. Route lengths are approximate values. Voyage times were calculated taking into account a sustained service speed of 28 knots. However, during the project only the route from Cadiz to the Canary Islands was considered. Wave climates along the routes are described by joint probabilities of occurrence P of significant wave height HS and zero upcrossing wave period T0. A convenient source of wave scatter diagrams is Global Wave Statistics (1990). For this investigation, only annual probabilities for all wave directions were selected. The wave scatter diagram is given in Appendix A. Table 3.2.2: Reference routes Reference Routes Cadiz - Canary Islands Ancona - Patras Code CC AP Length (nm) 710 500 Voyage time (h) 25 18

3.3 Risk assessment


Risk assessment in the maritime industry (slowly) follows the routes that other industries have gone before, e.g., the nuclear industry, the hazardous industry and the offshore industry. In shipping, formal safety assessment started 1993 with the United Kingdom proposal to IMO. It is necessary to distinguish between formal safety assessment (FSA) that is related to the rule making process, and risk assessment (RA) that is intended to be used in design and operation. Few risk assessments were published lately and it seems that more experience must be gathered before the usefulness of risk assessment will become obvious. The objective of the project DEXTREMEL was to contribute to this base of experience.

DEXTREMEL (BE97-4375) Final technical report page 8 / 75 Research activities were monitored by the concerted action FSEA (1996) and the effort is ongoing within the thematic network THEMES (2000). Risk assessment usually comprises five steps: 1) hazard identification, 2) determination of risk that is the product of probability and consequence, 3) identification of risk management options, 4) cost-benefit analysis, and 5), recommendations. Within the basic research project DEXTREMEL, only steps 1 and 2 were investigated. Furthermore, some identified hazards such as fire were not further evaluated because of the limited scope of the project. However, a cost-benefit analysis requires values to compare against the risk. These so called risk acceptance criteria are difficult to fix. Current practice dictate that risks should be reduced to As Low As Reasonably Practicable. A possible measure of cost effectiveness is the implied cost of averting a fatality (ICAF) that is the ratio of the increase in cost and the reduction in risk, Skjong and Ronold (1998). Within the project DEXTREMEL four hazards were identified: 1) collision, 2) grounding, 3) bow door failure, and 4) front bulkhead of superstructure failure. For three of the above hazards we predicted the annual risk. However, for the failure of the front bulkhead of superstructure our analysis did not fully succeed and, therefore, we do not report results here. Other hazards like fire were also not investigated, nor were fire, loss of human life and environmental pollution taking into account as possible consequences. The following sections present the results of the research project. The sections address collision, grounding, bow doors, green water and structural integrity. In each section we start with the presentation of new computational methods and model test results to validate the new tools. A risk analysis concludes each of the first four sections.

DEXTREMEL (BE97-4375) Final technical report page 9 / 75

Scientific and technical description of the results

4.1 Collision damage


It is a major challenge for the maritime community to develop probability-based procedures for design against collision and grounding events. To quantify the risks involved in ship traffic in specific geographic areas implies that probabilities as well as inherent consequences of various collision and grounding events have to be analyzed and assessed. In DEXTREMEL such a rational procedure for evaluation of the probabilistic distribution of damages caused by collisions against other ships for a specific ship on a specific route has been described and applied to the reference ship MS Dextra on the route between Cadiz and the Canary Islands.

4.1.1 Collision probability analysis


The first step in a risk assessment procedure for ship collisions is to apply a procedure for determination of the estimated frequency for collision events. Here historical data cannot be expected to give a sufficiently reliable prediction, see Pedersen (1995). In recent years there has been a rapid development of new navigational systems. A growing number of VTS systems are established around the world. Extensive trials have been carried out with sole lookout during night on ship bridges. IMO has introduced requirements for new ships to fulfill particular maneuverability criteria [IMO Resolution A 751]. AIS will be mandatory on future ships, and a new generation of large fast ferries has emerged. It is generally agreed that all these activities have considerable influence on the probability of ship accidents in the form of collisions and grounding. But so far no rational analysis tools to quantify the effect of these changes have been available. Instead nearly all research on ship accidents has been devoted to analysis of consequences of given accident scenarios. It is with this background that the work in this project on a rational model for determination of the probability for ship accidents has been carried out. The main principle behind the most commonly used risk models is to determine the number of possible ship accidents Na, i.e. the number of collisions if no aversive maneuvers are made. This number Na of possible accidents is then multiplied by a causation probability Pc in order to find the actual accident frequency. The causation probability Pc is the fraction of the accident candidates that result in an accident. For calculation of the number of possible ship collision candidates Na the shipping route and intensity of the ship traffic on this route must be known and grouped into a number of different ship classes according to vessel type, dead weight tonnage or length, loaded or ballasted, with or without bulbous bow etc. With this information at hand a procedure described by Pedersen (1995) was used to calculate the number Na of possible events where two ships will collide if no aversive maneuvers are made. The causation probability Pc can be estimated on the basis of available accident data collected at various locations and then transformed to the area of interest. Another approach is to analyze the cause leading to human inaction or external failures and set up a fault-tree procedure. In the DEXTREMEL project work we have applied a new method based on a Bayesian Network procedure for calculation of Pc presented in Hansen and Pedersen (1998). The methodology for the developed procedure is based on the assumption that the ship and crew characteristics and the navigational environment mainly determine the collision probability. That is, technical failures such as engine failure and rudder failure play a minor

DEXTREMEL (BE97-4375) Final technical report page 10 / 75 role. The most important ship and crew characteristics are taken to be: ship speed, ship maneuverability, the layout of the navigational bridge, the radar system, the number and the training of navigators, the presence of a look out etc. The main parameters affecting the navigational environment are ship traffic density, probability distributions of wind speeds, visibility, rain and snow. With knowledge of the ship characteristics and a study of the human failure probability, i.e. a study of the navigator's role in resolving critical situations, a set of causation factors Pc has been derived.By application of the Bayesian Network procedure for estimation of the causation probability, it is possible through future research to examine the beneficial effect of new bridge procedures, of improved maneuvering capability, of having a pilot on board, of introducing a VTS system in certain geographical areas, and of introducing AIS. The expected number of ship-ship collisions is then determined as

N ship - ship = Pc N a

(4.1.1)

Based on the mathematical model for estimation of collision probabilities described above a computer program named COLLIDE has been written for calculation of collision probabilities in specific waterways where the ship traffic distribution is known. As an example on the application of the procedure we shall consider the reference RoRo passenger vessel making round trips on the route between Cadiz in mainland Spain and Las Palmas and Tenerife on the Canary Islands. The main traffic routes in this area are shown in Fig 4.1.1. The reference vessel is assumed to make 245 voyages a year on this route.

Figure 4.1.1: Ship Traffic between Cadiz and the Canary Islands. On this route there is further a given distribution of other types of vessels such that three in principle different types of collisions can occur. One type of collision is a head-on induced collision due to two-way traffic in the straight waterway segments. Another type of collision on the Ro-Ro waterway occurs at bends where two straight route segments intersects. At such an intersection a ship can become a collision candidate if the course is not changed at

DEXTREMEL (BE97-4375) Final technical report page 11 / 75 the intersection. Finally, the model calculates the probability for collisions in areas with crossing shipping routes. For each waterway segment the number of collision candidates related to head-on, intersections where the shipping route bends, and crossing situations is calculated for each vessel type. The probabilities of collision is then estimated by multiplication with the causation factors verified by the Bayesian Network procedure P[head on] P[bends] P[crossing] = 4.910-5 = 1.310-4 = 1.310-4

The result of the numerical calculations is a complete probabilistic distribution of collision scenarios at all the geographic locations on the shipping route. The integrated result is that the annual probability of collision involving MS Dextra is .042. That is, a collision is expected to take place with a return period of 24 years. The calculations also show that the probability MS Dextra is the struck vessel in a collision is .021. It is of course not possible to directly verify these calculated collision frequencies. But the annual probability, 0.042, for a collision involving MS Dextra can be compared to the historical collision probability for RoRo vessels in Northern Europe which is 0.031. That is a figure not far from the one calculated for MS Dextra, especially considering that the sailing time for MS Dextra is high compared to other RoRo vessels.

4.1.2 Consequences given a collision


To determine the consequences given a collision has taken place the next step is to determine the energy released for crushing of the involved structures and the impact impulse of the collision by analyzing the rigid body motions of the colliding ships taking into account the effect of the surrounding water. Solution of this external mechanics problem is accomplished uncoupled from the solution of the internal deformation and is based on an analytical method developed in Pedersen and Zhang (1998). The energy loss for dissipation by structural deformation is here expressed in closed-form expressions. The procedure is based on a rigid body mechanism, where it is assumed that there is negligible strain energy for deformation outside the contact region and that the contact region is local and small. The collision is considered to be instantaneous as each body is assumed to exert an impulsive force on the other at the point of contact. The model includes friction between the impacting surfaces so that situations with glancing blows can be identified. Both the striking and the struck ships are assumed to have three degrees of freedom: surge, sway and yaw. The interaction between the ships and the surrounding water is approximated by simple added mass terms and energy released for crushing is determined in two directions, perpendicular and parallel to the side of the struck ship. The lost energy perpendicular to the side is used to determine the penetration depth. The bow of the striking vessel is assumed rigid so that all the energy must be absorbed in the structure of the struck vessel. Both right and oblique angle collisions are handled. As typical results for MS Dextra sailing on the route between Cadiz and Tenerife, Fig. 4.1.2 shows the probability density function for the energy released for crushing of the structure of MS Dextra given a collision has taken place. Also more detailed information on the energy distribution along the length of the ship can be determined. See Fig. 4.1.3. When the energy loss to be dissipated by destroying the side structure is known, the subsequent damage to the struck vessel has to be calculated. During the project a simplified method for calculating the collision force and the resulting hole in the ship structures have

DEXTREMEL (BE97-4375) Final technical report page 12 / 75 been applied. At the same time two comprehensive finite element models were established to verify the simplified method.

Figure 4.1.2 Probability density function for energy released for crushing given a collision on the route between Cadiz and Tenerife.

Figure 4.1.3 Conditional released energy distribution as function of longitudinal position along the hull.

DEXTREMEL (BE97-4375) Final technical report page 13 / 75 The side structure of most ships is very complex. The possible deformation and crushing modes are also very complex. Therefore, some simplifications need to be introduced for fast evaluation of the crushing damage. To do this we utilize that observations from full-scale ship accidents and model experiments reveal that the primary energy absorbing mechanisms of the side structure are Membrane deformation of shell plating and attached stiffeners Folding and crushing of transverse frames and longitudinal stringers Folding, cutting and crushing of horizontal decks Cutting or crushing of ship bottoms Crushing of bulkheads

By analyzing the crushing resistance of each of these basic structural elements and adding their contributions together, the total collision resistance and dissipated energy can be determined by a relatively simple and above all fast evaluation procedure. To describe the procedure further we shall assume that a collision position is located in the middle between two transverse frames, see the finite element model in Fig. 4.1.4. In the initial phases of the collision, the shell plating of the struck ship is subjected to tension. With increasing penetration, the striking bow comes into contact with frames, stringers and horizontal decks. The frames, the stringers and the decks are then subsequently crushed. It is assumed that frames, stringers and decks are not deformed and crushed until the striking bow touches them directly. By calculation of the resistance of the deformed shell plating, frames and decks etc., the collision resistance and the absorbed energy are obtained. When the calculated absorbed energy is equal to the energy loss, determined from the outer analysis procedure, the calculation stops. After the maximum penetration has been determined, the size of a hole in the shell plating created by the striking bow is also calculated by the simplified method, see Pedersen and Zhang (1999)

Figure 4.1.4. Finite element model (LS-DYNA) for calculation of crushing resistance for the case where an MS Dextra bow penetrates the side. The simplified procedure is based on standard bow forms on the striking vessels. The analytically based simplified procedure has been verified by comparison of calculated results

DEXTREMEL (BE97-4375) Final technical report page 14 / 75 with results from small scale experiments and from numerical calculations using explicit finite element codes. One verification example related to calculated forces on the bulbous bow and energy absorption as function of penetration is shown in Figs. 4.1.5 and 4.1.6. Here finite element results calculated by use of LS-DYNA (by DTU) and results obtained by the simplified procedure are compared for a collision situation similar to the one depicted in Fig. 4.1.4. The only difference is that the shape of the bow is changed to the generic shape used throughout the damage calculations in this project. Figures 4.1.5 also shows results calculated taking into account material hardening by use of the program MSC/DYTRAN by NTUA, see Servis and Samuelidis (1999), using a different meshing and the actual bow geometry shown in Fig. 4.1.4. Figure 4.1.6 shows that the energy absorption results generated by the simplified analytical procedure are quite close to the results determined by the much more laborious LS-DYNA finite element results for this case. The experience from this comparison is that when the geometries are similar then the simplified procedure normally gives good results. But as can be seen from the MSC/DYTRAN results in Fig. 4.1.5 then a small variation in the applied bow shape can result in quite drastic changes in the crushing forces depending on contact with different structural strength members.
3 ,0 E + 0 7

B u lb - fo r c e

2 ,5 E + 0 7

L S - D Y N A - s im p lif ie d b o w A n a ly t ic a l S im p lific a tio n - s im p lifie d b o w M S C /D Y T R A N

2 ,0 E + 0 7

Force [N]

1 ,5 E + 0 7

1 ,0 E + 0 7

5 ,0 E + 0 6

0 ,0 E + 0 0 0 ,0 0 ,5 1 ,0 1 ,5 2 ,0 2 ,5 3 ,0 3 ,5 4 ,0 4 ,5 5 ,0

D is p la c e m e n t [m ]

Figure 4.1.5. Force - penetration curve for a bulbous bow impact amidships.
3,5E+07 3,0E+07 2,5E+07 LS-DYNA Analytical Simplification

Bulb - Energy

Energy [J]

2,0E+07 1,5E+07 1,0E+07 5,0E+06 0,0E+00 0,0 0,5 1,0 1,5 2,0 2,5 3,0 3,5 4,0 4,5 5,0

Displacement [m]

Figure 4.1.6. Energy absorption as function of penetration for bulbous bow impact.

DEXTREMEL (BE97-4375) Final technical report page 15 / 75 The objective of additional finite element collision simulations of MS Dextra using the explicit code MSC/DYTRAN was to investigate the level of energy, which can be absorbed by the ships hull up to the time that critical situations arise for the safety of the ship, Servis et al (1999). Further, the numerical results were compared with the analytical results of collision simulation. It must be noted that the analytical and numerical simulations were not based on the same assumptions. It was investigated whether it is possible to reproduce failure modes, which have been observed during large-scale impact tests and to numerically predict measurements, which were taken during small-scale impact tests (Samuelides, 1984). Figures 4.1.7 and 4.1.8 present some results of these studies.

Figure 4.1.7: Penetration of a stiffened panel by a bulb

Figure 4.1.8: Stress results for small-scale collision simulation The FE simulation of a collision encompasses a number of problems, which should be given appropriate attention. These are a) the selection of a mesh, which should be fine enough for the results to converge and to reproduce the failure modes and coarse enough for executing the code in acceptable time, b) the inclusion of the effect of the surrounding water, c) the extent of the ship, which would be modeled as a deformable body, d) the modeling of the material, including the material failure criterion, which appears to be a weak point in the collision analysis, either this is performed with simplified analytical techniques or numerically. The preliminary runs, the simulation of impact on structural parts of ships and of small-scale collision tests lead to the following conclusions, which provide guidance for collision simulations and which were followed for the collision simulation of MS-Dextra. The mesh density used can be tested for convergence using F- curves. Conservation of energy should not be violated at any point of the simulation. Shell elements used by explicit codes tend to exhibit spurious energy modes that shatter the

DEXTREMEL (BE97-4375) Final technical report page 16 / 75 energy equilibrium. No results may be considered reliable if energy equilibrium does not hold at that point of the analysis. It appears that it is sufficient to model the structure of the impacted area and the adjacent compartments. For the simulation of central and right angle collisions, it is possible to model the rest of the hull with beam elements with appropriate stiffness and mass, which accounts also for the effect of the surrounding water (Servis, 1997). However such a simulation is not validated in the case of oblique and/or eccentric collisions. The hull is to be modeled across its breadth. Account should be taken for simulating contact of the bow with outer as well as inner structural elements. The contact definition is such that it accounts for the changes of the contact areas as elements fail. The inclusion of material hardening in the material modeling is of importance.

For the collision investigation of MS Dextra it has be assumed that she collided with a ship having a rigid bulbous bow similar to hers, see Fig. 4.1.9 for the modeling of the struck ship. The results of the analysis were used to predict the shape of the damage, Fig. 4.1.10, to determine the loads vs. penetration exerted by the bow and the bulb, Fig. 4.1.11, the energies absorbed by individual structural elements, to identify areas which although away from the contact exhibited high stresses and finally to determine the energies absorbed by the structure until critical situations occur. Two critical conditions were identified: a) when the upper part of the bow strikes the inner bulkhead, and b) when the bulb penetrates the outer shell. In the first case the energy absorbed was 24.93 MJ and in the latter 56.62 MJ.

Figure 4.1.9: finite element model of midship part.


Force exerted by the bulb
3.E+07 3.E+07 2.E+07 2.E+07 Force [N] 1.E+07 5.E+06 0.E+00 -5.E+06 -1.E+07 -2.E+07 Penetration [m] 0 0.25 0.5 0.75 1 1.25 1.5 1.75 2 2.25 2.5 2.75 3 3.25 3.5 3.75 4 4.25 x-comp y-comp z-comp

Figure 4.1.10: side damage, effective stress, approx. at 0.5 s after impact.
Force exerted by the bow
3.00E+07 2.50E+07 2.00E+07 1.50E+07 Force [N] 1.00E+07 5.00E+06 0.00E+00 -5.00E+06 -1.00E+07 -1.50E+07 Penetration [m]
0 0.25 0.5 0.75 1 1.25 1.5 1.75 2 2.25 2.5 2.75 3 3.25 3.5 3.75 4 4.25 4.5 4.75 5 5.25 5.5 5.75 6 6.25

x-comp y-comp z-comp

Figure 4.1.11: force results for MS Dextra force exerted by the bulb and bow respectively vs penetration from contact of bulb and bow respectively

DEXTREMEL (BE97-4375) Final technical report page 17 / 75

4.1.3 Collision damage distributions


Given that a collision has taken place the corresponding damage can be assesses by the described procedure. The IMO regulation A.265 (VIII): Regulations on Subdivision and Stability of Passenger Ships as an Equivalent to Part B of Chapter II of the International Convention for the safety of Life at Sea, 1960 gives such distributions given a collision. The basis of these damage distributions is 296 recorded damage cases. This may seem as a reasonable database, but due to the large variation in ship types and the development in ship design since 1940 to 1960, including the introduction of bulbous bows on the striking ships and the emergence of the RoRo concept, the statistical basis for the probabilistic modeling must be considered as somewhat outdated. Another driving force for the present development of a rational procedure for calculation of collision damages is that only by such methods is it possible to determine the influence of new more crashworthy side structures. As indicated in Fig. 4.1.4 then the model generally operates with the occurrence of two holes, an upper hole and a lower hole. Although the calculated geometry of these two holes in general is rather complicated, we have in the DEXTREMEL project restricted ourselves to present only the outer vertical and horizontal limits of these holes. For collisions having large collision energies the two holes will overlap. So, for practical purposes we have in the present study restricted ourselves to consider only four parameters. These are: The maximum length (or breadth) of upper and lower hole The maximum indentation depth The maximum height of the combine upper and lower hole The vertical upper limit of the hole.

As an example on the application of the derived procedure for calculation of collision probabilities and the associated damage distributions we consider the results given in Fig. 4.1.12 that shows the probability density function for the damage depth conditioned on the MS Dextra being the struck vessel. Similarly, Fig. 4.1.13 shows the calculated probability density function for the longitudinal length (or breadth) of the holes given MS Dextra is subjected to a side collision on the route between Cadiz and Tenerife.

Figure 4.1.12 Probability density function for damage depths given a side collision.

DEXTREMEL (BE97-4375) Final technical report page 18 / 75

Figure 4.1.13. Probability density function for damage breadth given a side collision.

4.1.4 Risk analysis


To estimate the significance of possible damages two criteria were introduced. If a collision occurs the ships hull and internal structural elements are ruptured. If the damage is located below the waterline and if the striking ship penetrates the hull of the struck ship water flows into several compartments of the ship. The consequence of such a flooding scenario is a reduction of ship stability. In the worst case the ship will capsize. An international convention with respect to the ship stability was issued by the International Maritime Organization IMO. This convention (SOLAS 1974) defines a maximum damage length dLmax and a maximum damage depth dDmax for ships. The value calculated with (4.1.2) is valid for "one compartment vessels". Considering "two compartment vessels" like MS Dextra the damage length has to be doubled. For the MS Dextra the maximum damage size is dLmax = 16.38 [m] and dDmax = 5.2 [m]. If the damage exceeds both maximum values the ship was considered as total loss.

dLmax = min { 3 + 0.03 L [m] ; 11 [m] } dDmax = B/5 [m]

(4.1.2) (4.1.3)

For the cost estimation of consequences two cases have to be distinguished. If the damage size exceeds the values of the stability criterion the ship capsizes and sinks. Then the consequence is the total loss of the ship. The costs of such a collision are the price of the ship. For the MS Dextra a price of Ccap = 3.00E+08 [EUR/collision] is assumed. If the damage do not exceed the value dLmax and dDmax the costs result from repair of the damage. The repairing costs depend on different aspects (damage location, damage size, repair yard) and can be estimated using the amount of (ruptured) steel that has to be repaired or the repair time that is needed. In this report the amount of ruptured steel is used for the costs estimation. The price of a typical repair job is assumed as c2 = 6.00E+03 [EUR] per ton of steel. This price includes the full repair process i.e., cutting-building-fitting of the damaged area. With respect to the voyage to the shipyard costs of c3 = 5.00E+04 [EUR/repair] are assumed. These costs include fuel consumption, crew payment and possible support by tugs.

DEXTREMEL (BE97-4375) Final technical report page 19 / 75 Using the simplified methods presented in chapter 4.1.3, the density function of the damage length DL, height DH and depth DD can be calculated. The density function for the damage length is shown in Fig. 4.1.13 for MS Dextra sailing from Cadiz to the Canary Islands. A three-dimensional joint probability density function Dcoll that includes all three damage parameters can be calculated as follows.

D coll = f(dL , dD , dL ) = DL DD DH

(4.1.4) (4.1.5)

Pcap = Pstruck

dLmax dDmax 0

coll

(dL , dD , dH ) ddL ddD ddH

The integration of the joint density function (4.1.5) delivers the probability that MS Dextra capsizes if it sails between Cadiz and Tenerife, Pcap = 4.19E-02. The probability that the damage can be repaired is Prep = 1-Pcap =9.58E-01. If the damage is smaller than the stability limit, dLmax and ddmax, the repair costs are computed from the expected mass of ruptured steel. This mass is calculated from the damage size and the related probability as well as the density of steel Fe = 7.85 [t / m3 ] and the ratio RV = 2.23E-02 of the ship steel volume Vsteel and ship volume Vship, i.e.

Vship = L B (T c B1 + (H T) c B2 )

(4.1.6)

The parameters in (4.1.6) are the ship length L, the ship breadth B, the ship draft T, the ship Height H, the block coefficient of the submerged ship CB1 and the block coefficient of the ship above the waterline CB2. Considering the MS Dextra the ship volume is Vship = 4.97E+04 [m3] using L = 173 [m], B = 26 [m], T = 6.5 [m], H = 15.7 [m], cB1 = 0.56 and cB2 = 0.8. The ship steel volume VSteel is the product of the light ship weight mSteel (excluding non-load carrying parts, e.g., machinery, propeller, etc.) and the steel density Fe . The ship steel volume of MS Dextra is Vsteel = 1.105E+03 [m3]. The joint probability density function Dcoll relates the damage size and the frequency of certain damages. The application of (4.1.7) gives the expected mass of steel that will be ruptured per collision in the case that the damage not exceeds the damage criteria. For the MS Dextra eq. ( gives E[mrup] = 1.65E+00 [t] based on an expected damage volume of 9.46E+00 [m3].

E[mrup ]=Fe R V

dLmax dDmax

d d d D(d ,d ,d
l d h l d 0 0 0

)ddl ddd ddh

(4.1.7)

Multiplying E[mrup] with the repair costs per ton steel c2 = 6000 [EUR/t] and adding the costs c3 = 5.00E+04 for the voyage to the ship repair yard yields the costs Crep of repair due to collision damages: MS Dextra: Crep = 5.99E+04 [EUR/collision]. The annual risk Qcoll calculated with the following equation is Qcoll = 1.92E+05 [EUR/year]).

Qcoll =fcoll (Pcap Ccap +Prep Crep )

(4.1.8)

Resulting annual risk due to collision is presented in Table 4.1.1. The main part of the risk is due to the capsizing of the ship that is characterized by the exceedence of damage criteria. A more reliable approach is needed in the future to accurately predict consequences, i.e., if the vessel will capsize or not following a collision.

DEXTREMEL (BE97-4375) Final technical report page 20 / 75 Table 4.1.1: annual risk of MS Dextra due to collision Annual collision frequency fcoll =0.042 [1/year] Probability of capsize Pcap = 3.92E-02 New building price of ship Ccap = 1.15E+08 [EUR/Collision] Probability of repair Prep = 9.58E-01 Expected repair costs Crep = 5.99E+04 [EUR/Collision]

Annual risk due to collision Q coll = f coll Pcap C cap + Prep C rep = 1.92E + 05 EUR/year

DEXTREMEL (BE97-4375) Final technical report page 21 / 75

4.2 Grounding damage

4.2.1 Grounding damage distributions


A methodology for the assessment of the grounding behavior of a vessel was developed by NTUA. The input required for the analysis are the particulars of the vessel main particulars and scantlings of the bottom/double bottom structure, and the output is the relationship between the causes of the grounding (kinetic energy of the ship prior to incident, morphology of seabed) and the consequences of the accident (for example, dimensions of damage, reduction in cross-sectional area). A similar approach for the behavior of the assessment of tankers involved in a collision has been developed by Samuelides (1999). The determination of the relationship between the causes and consequences of a grounding accident is performed in three steps: step 1. Definition of possible damage cases Grounding damages are assumed to have an orthogonal shape and each damage case is defined by five parameters: The longitudinal and transverse location of the fore end of the damage and its length, width and height. Depending on the method, which is used for the investigation of the internal mechanics the width of the damage may be a dependent parameter. A large number of damage cases is defined by varying the damage parameters with a predefined step, within a predefined range and predefined range. step 2. Determination of relationship of causes and consequences for each damage case. step 2.a. Definition of grounding scenario One or more grounding scenario(s) is (are) defined for each damage case defined in step 1. The grounding scenario is in more cases defined by the morphology of the rock on which the ship grounds. step 2.b. Calculation of absorbed energy The energy absorbed by the damaged structure is calculated for each damage case and grounding scenario. step 2.c. Determination of the consequences of damage The consequences of the damage are determined for each damage case. The consequences may be expressed as the length or volume of damaged material, the reduction of the cross sectional area, the reduction of ultimate strength of the hull, ... step 3. Determination of relationship of causes and consequences, in case of grounding. All cases analyzed in step 2, are grouped according to the energy absorbed and the grounding scenario. With appropriate compilation of the results the outcome of this step is the relationship of the causes and consequences of grounding for the given ship. In order to implement the proposed methodology for the assessment of the grounding behavior the computer code Grounding 2.0 was developed. The calculation of the dissipated energy, i.e. step 2.b, is based at present on four methods, those reported by Paik (1994, 1997), Simonsen (1997a, 1997b, 1997c), Minorsky (1959) and Wang et al (1995,1997). Wang et al and Paik assume grounding on a wedge shaped rock, while Simonsen considers grounding on a rock, which resembles to a cone. At present the consequences of the grounding are expressed either by the reduction of the cross-sectional area or in terms of the length and area of the damage. The methodology is adaptable, so that it will be possible to evaluate the grounding damage when the ship grounds at a given area, provided that the morphology of the seabed of that area is known. Grounding 2.0 is an application built in Visual Basic 6.0. Taking advantage of the new features in Visual Basic 6.0, regarding data access, it has been decided to store the input

DEXTREMEL (BE97-4375) Final technical report page 22 / 75 and output of the program in a database. This resulted in an efficient way of manipulating the data, input and output. Input data such as the shape of the waterline at the double bottom level and the structural configuration of the double bottom are entered into online forms. The user also decide which method(s) is (are) to be employed for the calculation of the energy dissipation and the way that he wishes to express the consequences of the damage. After completion of the calculations the user may view the results, relevant to each damage case considered, as these results have been stored in a database. These individual results are properly manipulated by the code, to produce the relationship of the causes versus the consequences of the grounding. Figures 4.2.1 to 4.2.3 show the relationship between absorbed energy and the damage average length and average reduced cross-sectional area for a grounding of MS-Dextra. The charts show, for each range of absorbed energy (x-axis): the average damage length and its variation the number of damage cases, where the absorbed energy was in the same range the corresponding velocity of the vessel, on the basis that the kinetic energy of the vessel is absorbed by the deformation of the metal structure and for displacement (including the effect of the added-mass).

Figure 4.2.1: Absorbed energy vs. Damage Length according to Wang et al.

Figure 4.2.2: Absorbed energy vs. Damage Length according to Simonsen

DEXTREMEL (BE97-4375) Final technical report page 23 / 75

Figure 4.2.3: Absorbed energy vs. Damage Length according to Simonsen. It should be noted that in the case of Fig. 4.2.2 the results were generated on the basis that the energy is absorbed solely by the fracture of the outer plate, as in the case of Fig. 4.2.1. Higher resistance is exhibited if fracture of all members is taken into account, Fig. 4.2.3). Parametric study conducted using the code revealed that a more dense arrangement of the transverse webs is more effective, in terms of energy absorption during grounding, than an increase of the plate thickness. The program has been used to estimate the grounding behavior of a single skin VLCC, which was grounded on Buffalo Reef on January 6, 1975. The particulars of the vessel were given in (Simonsen, 1997). Figure 4.2.4 shows the damage length obtained by the program and the measured damage of the actual grounding.

Figure 4.2.4: Absorbed Energy vs. Damage Length according to Simonsen. The actual grounding event is marked as , as she rode over the rock at a speed of 12 knots resulting in a damage 180 m long.

4.2.2 Risk analysis


This section focuses on a method to calculate the risk due to hard grounding events. Hard grounding describes a ship that grounds on hard materials like rocks or coral reefs. Soft grounding means that ships ground on a sea floor consisting of soft material like sand or clay which is not considered here. In the first part of the risk calculation for grounding scenarios

DEXTREMEL (BE97-4375) Final technical report page 24 / 75 the frequency of a possible grounding event will be calculated. In the second part the consequences are expressed as repair or new-building costs. The grounding scenario is shown in Fig. 4.3.5 that shows two distributions reflecting probable routes of the MS Dextra. It must be noted that the scenario is not related to the reference route from Cadiz to the Canary Islands. The probable routes were assumed as normal distributed with a mean value of 900 m and a standard deviation of 900 m. In a distance of 1800 m from the center of the traffic lane an obstacle with a diameter of 300 m is located. The integration of those parts of the route distributions that hit the obstacle delivers the probability per voyage that MS Dextra grounds if no maneuvers will be done to avoid the grounding. In the considered example these values are PG1 =1.32E-01 and PG2 = 1.82E-02. The probability Pcg = 1.59E-04 that the crew notices the obstacle and changes the course and avoids grounding is provided in Fujii and Mizuki (1998). Assuming that the MS Dextra makes nv = 240 voyages per year yields the annual frequency of grounding fG = 5.78E-03.

fG =nv Pcg (PG1 +PG2 )

(4.2.1)

R o u te D is tr i b u tio n 2 300m 900m


O b s ta c le

C e n t e r o f th e T ra ffic L a n e

1800m R o u t e D is t r ib u t io n 1

Figure 4.2.5: example of a possible grounding scenario. A beta distributed probability density function of the ship speed was assumed to reflect the probabilistic character of the ship speed, see Fig. 4.2.6 for operation on open sea. Assuming ship operations in coastal waters the modal value of the probability density function would be located around 9 knots.
0 .1 0 .0 9 0 .0 8 0 .0 7 frequency 0 .0 6 0 .0 5 0 .0 4 0 .0 3 0 .0 2 0 .0 1 0 0 5 10 15 v e lo c it y [k n o ts ] 20 25 30

Fig. 4.2.6: Probability density function of ship velocity of MS Dextra

DEXTREMEL (BE97-4375) Final technical report page 25 / 75 The computer program Grounding 2.0 was used to calculate the damage size of a grounding event related to the ship speed or kinetic energy. The damage size varies depending on the calculation method, the location of the damage and the geometry of the obstacle. Figure 4.2.7 shows the average damage sizes of the MS Dextra grounded on an obstacle of 300 m diameter. The top of the obstacle is located 0.5 m above still water level. With respect to a ship draft of T = 6.5 m there is a possible damage height of 7 m. The combination of the velocity distribution and the functions given in Fig. 4.2.7 yields the density function of the damage parameters DL, DD and DH.
120.0
damage length damage depth damage height

100.0

size of damage [m]

80.0

60.0

40.0

20.0

0.0 0.0 5.0 10.0 15.0 velocity [knots] 20.0 25.0 30.0

Figure 4.2.7: damages due to hard grounding of MS Dextra Further calculations are similar to the collision approach yielding a joint density function DGr of grounding damage, see (4.2.2). A damage stability criterion was introduced to rate the significance of the damages. If the damage exceeds certain limits the ship is considered to capsize. The "International Convention for the Prevention of Pollution from Ships" MARPOL defines damage limitations that are valid for tankers. For RoRo vessels no rules or guidelines exist for grounding. MARPOL damage criteria were applied to the MS Dextra. Equations (4.2.3), (4.2.4) and (4.2.5) show the dependency of the damage criteria on the ship dimensions.
DGr = f(dL , dD , dL ) =

DD DH

dHmax

dLmax = 0.4 L [m] dDmax = B/3 [m] = height of the double bottom [m]

(4.2.2) (4.2.3) (4.2.4) (4.2.5)

For MS Dextra, the criteria yield dLmax = 69.2 [m], dDmax = 2.73 m and dHmax = 1.48 m. Considering one grounding event the probability of capsizing can be calculated with (4.2.6).

Pcap =

dLmax dDmax dHmax

coll (dL , dD , dH ) ddL

ddD ddH

(4.2.6)

For MS Dextra (4.2.6) gives a value of Pcap = 6.70E-01. The probability that the ship does not sink in case of grounding is only 3.30E-01. The damage criteria will exceeded at a velocity of around 21 knot]. The costs of the sunken ship was assumed as Ccap = 1.15E+08 Euro. The

DEXTREMEL (BE97-4375) Final technical report page 26 / 75 application of (4.2.6) gives the expected mass of steel that ruptured per grounding event if the ship does not capsize.
dLmax dDmax d H max

E[mrup ] = Fe R V

d d
l 0 0 0

dh D(dl , dd , dh ) ddl ddd ddh

(4.2.7)

For MS Dextra (4.3.7) gives E[mrup] = 2.08E+02 t. Multiplying E[mrup] with the repairing costs per ton steel c2 = 6.00E+03 Euro and adding the costs c3 = 5.00E+04 Euro/repair for the voyage to the ship yard yields the costs Crep of repairing efforts due to grounding damages (MS Dextra: Crep = 1.30E+06 Euro/grounding). The annual risk QGr = 0.45E+06 Euro/year was calculated with (4.2.8).

QGr = fG (PcapCcap +PrepCrep )

(4.2.8)

Table 4.2.2 presents results of risk analysis for grounding. The main part of the risk is due to the capsizing of the ship that is characterized by the exceedence of damage criteria. A more reliable approach is needed in the future to accurately predict consequences, i.e., if the vessel will capsize or not following grounding. Table 4.2.2: annual risk of MS Dextra due to grounding Annual frequency of grounding events fG=5.78E-03 Probability of capsize Pcap = 6.70E-01 New building price of ship Ccap = 1.15E+08 Euro Probability of repair Prep = 1- Pcap = 3.30E-01 Repair costs dependent on damage size Crep = 1.30E+06 Euro

Annual risk due to grounding

Q Gr = f G (Pcap C cap + Prep C rep ) = 0.45E + 06 EUR/year

DEXTREMEL (BE97-4375) Final technical report page 27 / 75

4.3 Bow door damage

4.3.1 Introduction
Expert discussions after the "Estonia" disaster in 1994 focused on technical and human failures (Final report 1997). Among the technical failures, strength of bow door locks subject to extreme wave loads was considered a most critical issue. Loads acting on bow doors are highly nonlinear for a variety of reasons, for example, bow flare shape, disturbed wave system at the bow, inclusion of air pockets, etc. Two approaches to determine bow door loads were presented in the literature. Unfortunately, no comparison between those approaches was made or published. Kvlsvold, Svensen and Hovem (1996) used a nonlinear strip theory method and computed impact forces on 2D sections normal to the stem contour according to simple wedge/circle impact pressure formulas. They accounted for swell-up due to forward speed and systematically varied both bow flare angle and bow stem angle. They performed a systematic comparison of numerical predictions with model tests and with classification society's rules. stergaard, Rathje and Sames (1996) presented a time domain method to predict loads on bow doors. They used a linear panel method to predict normal velocities. Impact pressure coefficients were determined based on an empirical approach. Impact pressures were then determined in the time domain combining relative normal velocities and impact coefficients. A single loadcase was investigated that yielded loads comparable to classification society's rule loads. Recently ship motions and loads were predicted with a Navier-Stokes solver using a hybrid grid system to couple the free surface flow with ship motions, see Kinoshita, Kagemoto and Fujino (1999). Although no extreme load cases were presented, the applied method has the potential to accurately compute bow door loads in extreme wave conditions.

4.3.2 Model tests


An experimental investigation of bow door loads on a model of MS Dextra has been performed at MARIN. The objectives were to measure (simultaneously) local and global loads on the bow door and to measure wave profiles just ahead of the bow, during the bow water entry. The model was constructed by MARIN at a scale of 1/40 and equipped with 9 pressure sensors at the bow segment that was in itself attached to a 6-component force gauge. In addition, ship motions and accelerations of the bow segment were recorded. The model test campaign consisted of tests in regular and irregular waves at several ships speeds. To identify effects of flare angle two bow geometries having different flare angles were tested. Figure 4.3.1 shows the model with emphasis on the installations at bow.

Figure 4.3.1: Model of MS Dextra for measurements of bow door loads.

DEXTREMEL (BE97-4375) Final technical report page 28 / 75 A new non intrusive experimental technique based on free surface visualization by an underwater laser sheet has been adopted and employed to measure wave contours. A vertical laser sheet has been generated from the model bulbous bow in the median plane of the model, just in front of the bow. The intersection of the laser sheet with the free surface materializes a cut of the free surface, which is video recorded and then extracted by means of image digitalization and processing techniques. In this purpose, the model has been additionally equipped with optical equipment. The laser sheet was generated by a laser beam, transported inside the model from the laser source on the carriage by a fiber optic, and a cylindrical lens, see Fig. 4.3.2. The bulb was separated from the model by a waterproof window. The lens was mounted close to the window, on the model side, so that the laser sheet crossed the window. The laser sheet was then reflected upward by a mirror mounted inside the bulb, and crossed a second window mounted on the top of the bulb. Images were recorded by a video camera, mounted in a waterproof container, rigidly fixed on the starboard side of the model, Fig. 4.3.3. Thus, images are obtained in a reference frame linked to the model.

laser sheet

bow window s fibre optic cylindrical lens

bulbous bow mirror

Figure 4.3.2: Generation of a vertical laser sheet from the model bulbous bow

Figure 4.3.3 - Video camera mounted on the model During experiments, the images were stored on a video tape, and were simultaneously digitized, for duration on the order of 10 s to 20 s, on a dedicated PC computer. A flash light, triggered by an analog signal logged by MARIN, was used to allow further time

DEXTREMEL (BE97-4375) Final technical report page 29 / 75 synchronization between MARINs measurements and images. In addition, high speed video recordings (500 frame/s) have been taken for some tests. Examples of images obtained and of wave profiles extracted from images are presented in Fig. 4.3.4. No major difficulty has been encountered regarding the equipment and its integration in the model. However, analyzing images was difficult due to the physics of the phenomena in the severe operating conditions selected (large forward speed).
350

330

310

290

270 Z ( mm )

250

230

210

190

170

150 650 600 550 X ( mm ) 500 450

Figure 4.3.4 Example of image obtained (with extracted free surface profile in red) and of series of profiles obtained for one run Firstly, when the bulb enters the free surface, a strong wake is created with separation on the bulb sides and creation of an eddy with air pockets on the top of the bulb. Jets are then created, which go towards the bow. Due to this flow pattern, the window on the top of the bulb is covered with a very disturbed air-water interface, which causes the laser sheet to diverge completely. To reduce this effect, the initial bulb has been replaced by a narrow one. Although reduced, the wake effect still existed. Secondly, the oncoming water was projected forward by the bow, creating a water curtain masking the field of view of the camera. This effect was reduced at the lower speed, and with the second bow shape.

4.4.3 Numerical analysis


The present investigation was initiated to predict time varying and highly nonlinear loads on a bow door with a new method based on results originating from linear seakeeping tools. The new tool is referred to as time domain postprocessor named GLTiS. The approach avoids time consuming time domain seakeeping predictions but aims at retaining the advantages of such tools, see Sames, Muzaferija and Rathje (1998). Results from a linear seakeeping code were used as input to the time domain postprocessor. The authors used the program GLPANEL, see stergaard and Schellin (1995), to predict response amplitude operators of hydrodynamic pressures and relative velocities normal to the panel surfaces. A finite volume method was used to predict impact pressure coefficients. Although three-dimensional effects were expected to play an important role, only two-dimensional simulations were performed and impact pressure coefficients were extracted from the resulting time traces of pressure. Finally, three-dimensional impact pressure coefficients were generated from sectionwise distributions. Resulting two dimensional impact pressure coefficients are shown in Fig. 4.3.5 for two sections. The original and the modified bow were compared with each other. Main difference occurred at section 20.5 where the modified bow geometry which exhibited a flatter (less convex) shape, see Fig. 3.2.2, resulted in higher impact pressure coefficients than the original bow shape. The postprocessor combines the results of the linear seakeeping prediction and the impact pressure coefficients in the time domain for any specified wave condition. RAO of pressures are used to determine wave heights and the wetting of the hull. Impact pressure coefficients

DEXTREMEL (BE97-4375) Final technical report page 30 / 75 are used to predict slamming loads if a specified threshold velocity is exceeded by the normal relative velocity. Bow door forces are determined by integrating the instantaneous pressure loads. Results are transferred to the visualization tool GLASiS.

Figure 4.3.5: Comparison of two dimensional impact pressure coefficients for original (solid line) and modified bow geometry (dashed line). Section 19.5 (left) and section 20.5 (right). To compare with model tests a part of the bow structure was assumed to represent a bow door. The bow door extended from section 19.5 forward and from WL 9 upwards. Figure 4.3.6 shows the discretized bow segment. During the simulation, the pressures acting on the bow segment were integrated to yield the three-component time dependent vector of bow force. A pressure distribution at bow for a single instant of time is shown in Fig. 4.3.7.

Figure 4.3.6: Panelization of bow segment. Figure 4.3.7: Typical pressure distribution Comparison of measured and predicted bow door forces generally showed good agreement. An example is presented in Fig. 4.3.8 for a test case with 5 m wave height. Peak value of force was well predicted. However, the duration of the force was overpredicted. Predicted bow forces for both bow shapes are compared with measurements in Fig. 4.3.9. Results were compiled for a ship speed of 26 knots and for a constant wave height. As expected, predicted forces for the modified bow were smaller than for the original bow. Maximum forces occurred for wave frequency of =0.55 rad/s. Predicted forces for smaller wave lengths were smaller than measured forces for both geometries. To identify the effect of impact pressure coefficients, computations were repeated neglecting the impact pressure coefficients. As expected, forces were lower than before. At higher wave frequencies, predicted vertical force

DEXTREMEL (BE97-4375) Final technical report page 31 / 75 was smaller than measured. This indicated that impact pressure coefficients were only partly responsible for the unfavorable agreement with measurements. In addition, measurements of local pressures and computed relative normal velocities were used to identify "correct" impact pressure coefficients. Relative normal impact velocity was similar for both bow geometries. Results indicated that "correct" impact pressure coefficients were between k=3.5 and k=5 for locations along the stem. Higher impact pressure coefficients were deduced for locations further aft but a clear tendency was not observed.

Figure 4.3.8: Typical time traces of vertical force on bow segment. Comparison of model test results and prediction.
Vertical Force [kN]
With pressure coeff. 16000 14000 12000 10000 8000 6000 4000 2000 0 0,40 0,50 0,55 Wave frequency [rad/s] 0,60 0,65 without pressure coeff. Experiment
16000 14000 12000 10000 8000 6000 4000 2000 0 0,40 0,50 0,55 Wave frequency [rad/s] 0,60 0,65

Vertical Force [kN]


With pressure coeff. Without pressure coeff. Experiment

Fig. 4.3.9: Comparison of predicted and measured vertical forces for original bow (left) and modified bow (right). The wave contour ahead of the bow was measured using a laser sheet that was directed perpendicular to the undisturbed free surface. Measurements displayed in Fig. 4.3.10 were shifted in time to more clearly match the predictions. Note that an arbitrary shift in time was valid for the purpose of this comparison. Overall agreement of predicted bow wave heights with measurements was favorable. The resulting slope (vertical wave velocity) agreed for the other cases with experiments.

DEXTREMEL (BE97-4375) Final technical report page 32 / 75

Figure 4.3.10: Comparison of measured and predicted bow wave contours. The clam doors form a weather-tight part of the ship's hull. Each door is hinged with a swinging arm that extends from the side structure of the hull to the door. Double-acting hydraulic cylinders swing the doors outwards in a parallel motion. This motion is ensured by arc-shaped rails at the upper part of the hull and smaller arms, attached to the hull and doors. The doors are equipped with individual locking mechanisms. Four cleats arranged around each door secure it at the closed position. A wedge-cleat is located at the aftermost part of the door and locks onto the underlying deck and another at the fore part of the hull and locks onto the door. Two push-cleats are located at intermediate positions on the hull. The two doors are locked together by three hooks. For stopping the motion of the door, three trapezoidal stops between the door and hull are located at its upper part. Three additional stops are located at the lower part of the door. Along the outline of the door, a sealing interface is applied in order to provide the required weather-tightness. The finite element modeling of the bow door, Fig. 4.3.11,and bow structure, Fig. 4.3.12, based on the lines of the platform delivered by AESA and on scantlings, which were determined by NTUA. The material is mild steel and hardening is taken into account.

stop

cleat hinge

stop
Figure 4.3.11: The FE model of the bow door.

DEXTREMEL (BE97-4375) Final technical report page 33 / 75 The aim of the finite element analysis was to determine the stresses induced to the primary members of the bow doors, such as webs and stringers. Four node shell elements, with 6 degrees of freedom per node were used to model the shell, plate elements and webs of the bow, whereas all stiffeners of the decks and bulkheads, as well as the flanges of the web frames, are modeled using beam elements. Thus the model may simulate a) the bending and the torsion response of the webs, and b) the response of the plating, attributed to the inplane and out-of-plane deflections, and rotations of the nodes connecting the webs to the plating. Such a response may induce on one hand axial stresses due to the in-plane deflection of the nodes of the element, and bending stresses as a result of the curved shape of the shell element bent in two directions. The present analysis did not account for the bending response of the plating due to the distributed pressure acting between the webs (tertiary stresses). The arm, hinges and a hydraulic cylinder, which is attached to the hull and keeps the door in closed position, are modeled using multi-point constraints (MPC's). These are special purpose rigid bodies that generate dependency equations of the degrees of freedom of certain nodes to independent degrees of freedom of a single node. The cleat arrangement, which is made of high tensile steel, is considered rigid in the finite element model. Thus the four cleats around the door are modeled using MPC's. In this case, certain nodes of either the bow door or the underlying deck are constrained to the nodes of the cleat supports. The three stops at the upper part of the door, which are modeled with shell elements, are trapezoidal and prevent the doors from moving inwards, when at the closed position. Gap elements are used to simulate the contact between the stops and the hull closing stringer. These elements when compressed, i.e. the gap closes, are very stiff and when the gap is opening they do not exhibit any resistance. The gaps allowed between the two structures are artificial, since they do not actually exist in the real structures, but very small (about 2-3 cm). The three stops at the lower part of the bow door are modeled using MPC's.

Figure 4.3.12: Half model of the bow section and door (centreline bulkhead not shown) Gap elements are also used to model the sealing interface between the doors and each door and the hull. According to the drawings of the interface consists of packing corners and flatbars that keep a glued rubber-like material in place (Mc Gregor Ltd). According to the modeling the doors are allowed to move away from each other and from the hull, as far as the locks allow for such movements, which is also the case in the actual arrangement. The three hooks that act between the two doors keep the two doors together (Mc Gregor Ltd) are modeled using MPC's between the respective supporting elements on both doors.

DEXTREMEL (BE97-4375) Final technical report page 34 / 75 The FE model of the bow section extends from frame 204 the collision bulkhead is located at frame 206 forward and along the whole depth. The forward bulkhead above the bulkhead deck is made continuous throughout the whole depth of the ship. A large opening is also cut out of the bulkhead where a weather-tight door should be fitted. This door is not included in the FE model. A longitudinal bulkhead extending from the collision bulkhead forward and an intermediate deck between 10mBL and 15.7mBL was added to provide support to the closing stringer and cleats. Additional fittings and solid floors were also used for this purpose. For the simulation of the bow door slamming, the pressures were applied over the bow and bow door area. Three different pressure patterns were applied: a) uniform pressure, b) nonuniform pressure, which is symmetric with respect to the longitudinal plane of the vessel, and c) non-uniform and asymmetric pressure pattern to account for the case that the vessel travels in oblique seas. Stresses were used to identify the most critical areas and to investigate the level of the stresses on the primary structural members. The obtained results could be used to carry out an analysis of the plating between the webs, as well as of hot spots, i.e. stress concentration points, in particular in areas adjacent to the connecting elements. In one loading case the stresses obtained were compared to the allowable stresses for the primary members according to Germanischer Lloyd (2000). It is noted that the allowable stresses in the plating are higher when tertiary stresses are included in the analysis. The nodal forces on the connecting elements were used to calculate the forces acting on these elements for a specific pressure level. On the basis of this force it was possible to perform a detailed analysis of these elements and subsequently to carry out an adequate design. Further investigation may refine the idealization used in this project for the modeling of these elements. Such a methodology for the calculation of the forces acting on securing elements has not been retrieved in the existing literature. Figure 4.3.13 presents the von Mises stresses, for a pressure level of 140.71 kPa, which corresponds to the design pressure for restricted international service according to GL. The allowable primary von Mises stress also according to GL equals 150 Mpa.

Figure 4.3.13: : v equivalent (Von Mises) stresses: total results and bounded by the permitted values It can be seen that though the structure itself does not suffer large stresses due to the pressure loading, there were areas where the stresses are above 150MPa. These areas were the side shell above the closing stringer and especially the area where the upper cleat secures the door, the stops at the lower part of the door and the lower part of the hinge. These stresses were below the yield stress, except for two localized spots: the hard corners

DEXTREMEL (BE97-4375) Final technical report page 35 / 75 on the lower mid and aft stops. The relative high stress areas under this loading appeared to be local, and a local finite element analysis with a model including the local stiffening elements, which strengthens the areas of the bow doors adjacent the connecting elements, would provide a more accurate picture of the stress field. Such a modeling should be based on actual drawings for the particular components and account for the geometry and material properties of the equipment delivered by the manufacturing company. Relative higher forces were also exerted in the plating adjacent to the lower stops rather than to upper, a factor which should be taken into account in the design of the door.

4.3.4 Risk analysis


A statistical approach was employed to evaluate risk of bow door (failure). Prediction of probability of occurrence is based on standard long term statistical analysis using the wave climate data for the reference route. Consequences were estimated using current IACS rules for bow door loads. The long-term statistics presented here were computed by a modified method presented in Rathje et al. (2000) based on the assumption that the ship can be considered as a linear system in conjunction with (nonlinear) pseudo-transfer functions (PSF) defined for different wave heights and ship velocities. As pseudo-transfer function Y the bow door force F divided by the wave amplitude H/2 is used:

Y(H, T, , v) =

F(H, T, ,v) H/2

(4.3.1)

Pseudo-transfer functions are related to regular waves that are defined by the wave height H, the wave period T, and the angle of encounter . In addition to the procedure presented in Rathje et al. (2000), a dependency between the pseudo-transfer function and the ship velocity v was introduced. Based on the assumption that the ship sails with different velocities dependent on different sea states a distribution of ship speeds was assumed taking into account the velocity reduction due to waves, propeller racing and the speed reduction due to the crew. Each sea state described in the wave scatter diagram is assigned a ship speed. The probability distribution P[Hs,Tz] of sea states is in the wave scatter diagram, see Appendix. The distribution of ship speed is given in Table 4.4.1. Application of (4.3.2) yielded the expected ship velocity vexp = 12.95 [m/s] which was very close to the service speed.

v exp =

P[H , T ] v(H , T )
s z s z Hs Tz

(4.3.2)

Using the above method, an assumption is needed about the relation between the wave height H used to calculate the pseudo-transfer functions and the significant wave height Hs. The ratio H/Hs determines the relation between the height of a regular wave H and the significant height of a sea state Hs. Using the assumption that the seaway contains as many waves with heights greater than H as waves with heights lower than H yield to the ratio H/Hs = 0.59, see Rathje et al. (2000). We used results of model tests to determine this ratio: test 104 with Hs=5 m and Tp =10 s, and test 105 with Hs=7 m and Tp =12 s. Model tests were conducted with irregular head waves for about 800 s of real time. Measured peaks of bow door force were divided into classes of certain load intervals. The classification yielded the histogram shown in Fig. 4.3.14. Table 4.3.1: Ship speed distributions as a function of wave height
v [m/s] 3.5 4.5 5.5 6.5 Zero upcrossing period Tz [s] 7.5 8.5 9.5 10.5 11.5 12.5 13.5

DEXTREMEL (BE97-4375) Final technical report page 36 / 75


14.5 13.5 12.5 11.5 10.5 9.5 8.5 7.5 6.5 5.5 4.5 3.5 2.5 1.5 0.5 2.50 2.50 2.50 2.50 2.50 3.10 4.40 5.60 6.95 9.35 11.25 12.50 13.50 13.50 13.50
30 25 20 frequency 15 10 5 0
8 11 2 3 4 5 6 7 9 12 15 16 19 20 =1 0 14 1< = 2< = 3< = 4< = 5< = 6< = 7< = 8< = 10 <= 12 <= 11 <= 14 <= 15 <= 16 <= 18

Sig. Wave Height Hs [m]

2.50 2.50 2.50 2.50 2.50 3.10 4.40 5.60 6.95 9.35 11.25 12.50 13.50 13.50 13.50

2.50 2.50 2.50 2.50 2.50 3.10 4.40 5.60 6.95 9.35 11.25 12.50 13.50 13.50 13.50

2.50 2.50 2.50 2.50 2.50 3.10 4.40 5.60 6.95 9.35 11.25 12.50 13.50 13.50 13.50

2.50 2.50 2.50 2.50 2.50 3.10 4.40 5.60 6.95 9.35 11.25 12.50 13.50 13.50 13.50

2.50 2.50 2.50 2.50 2.50 3.10 4.40 5.60 6.95 9.35 11.25 12.50 13.50 13.50 13.50

2.50 2.50 2.50 2.50 2.50 3.10 4.40 5.60 6.95 9.35 11.25 12.50 13.50 13.50 13.50

2.50 2.50 2.50 2.50 2.50 3.10 4.40 5.60 6.95 9.35 11.25 12.50 13.50 13.50 13.50

2.50 2.50 2.50 2.50 2.50 3.10 4.40 5.60 6.95 9.35 11.25 12.50 13.50 13.50 13.50

2.50 2.50 2.50 2.50 2.50 3.10 4.40 5.60 6.95 9.35 11.25 12.50 13.50 13.50 13.50

2.50 2.50 2.50 2.50 2.50 3.10 4.40 5.60 6.95 9.35 11.25 12.50 13.50 13.50 13.50

test No 104 test No 104

vertical bow door force Fz x103 [KN]

Figure 4.3.14: Histogram of measured amplitudes of bow door force Both histograms of amplitudes showed a similar pattern with a second maximum which was probably induced by impulsive loads while the global maximum of frequency is caused by non-impulsive (wave) loads. Due to the two maxima of the frequency distribution of amplitudes, a Rayleigh distribution can not fit the distribution. Therefore, it is questionable to use standard long-term and short-term statistics to the stochastic process of bow door loads. To assess the response of the ship structure to stochastic loads, the number of exceedance of certain bow door load levels were considered. Measured functions were not smooth due the limited number of measured amplitudes. However, these functions were used to make a comparison with short-term statistics and the measured values from the model tests. To calibrate our statistical method different ratios of H/Hs were considered and compared with model test results, see Fig. 4.3.15. Iteratively, the best ratios H/Hs were found that fitted the test data roughly. The result that two different ratios H/Hs were found to match the model test data raised the question which ratio should be taken in long-term statistics were all sea states are considered. This could mean that it is not possible to use non-linear pseudo-transfer functions for the long-term analysis of vertical loads of bow doors.

20 <=

19 <=

13 <=

17 <=

18 <=

9<

21

13

17

DEXTREMEL (BE97-4375) Final technical report page 37 / 75


1.4E+04 1.2E+04 vertical bow force [KN] 1.0E+04 8.0E+03 6.0E+03 4.0E+03 2.0E+03 0.0E+00 1 10 100 number of exceedance 1000

vertical bow force [KN]

model tests No 104 short-term statistics, H/Hs=1.9 short-term statistics, H/Hs=1.0 short-term statistics, H/Hs=0.59

2.5E+04 2.0E+04 1.5E+04 1.0E+04 5.0E+03 0.0E+00 1

model test No 105 short-term statistics, H/Hs=1.1 short-term statistics, H/Hs=1.0 short-term statistics, H/Hs=0.59

10 100 number of exceedance

1000

Fig. 4.3.15: Comparison of model tests with short-term statistics Pseudo-transfer functions of bow door force were systematically computed for 11 wave amplitudes, 7 angles of encounter, and 29 wave periods using the approach outlined above. The maximum force for each case was considered as reference force of the considered wave condition. Figure 4.3.16 shows typical pseudo-transfer functions for two ship speeds and two bow shapes. Vertical forces increased with the ship speed for both bow shapes, and vertical forces were higher for the original bow shape than for the modified bow shape.
MS-Dextra, vertical force on bow door wave amplitude 4 [m], angle of encounter 180 [] 8.00E+03 MS-Dextra 2, v=27 [knots] vertical force/ wave amplitude [kN/m] 7.00E+03 6.00E+03 5.00E+03 4.00E+03 3.00E+03 2.00E+03 1.00E+03 0.00E+00 0 5 10 15 wave period [s] 20 25 30 MS-Dextra 1, v=27 [knots] MS-Dextra 2, v=18 [knots] MS-Dextra 1, v=18 [knots]

Figure 4.3.16: Pseudo-transfer functions for different velocities and bow shapes An operation time has to be assumed in long-term statistical analysis. The chosen operation time is based on the yearly number of voyages of the ship. The distance between Cadiz - Tenerife is estimated as 710 [nm]. Assuming 240 voyages per year, a total distance of 170400 [nm] per year has to be covered. The total travel time of the ship is 263 [days/year] for a ship speed of 27 [knots]. The assumption of velocities depending on the sea state, see Table 4.3.1, yielded an operational time of 281 [days/year]. According to IACS rules (1999), the design pressure and vertical bow door force are calculated as: 2 (4.3.3) p e = 2.75 c H [0.22 + 0.15 tan()] 0.4v sin() + 0.6 L (4.3.4) Fz = p A

with the coefficient describing the area where the ship is intended to be operated. We used =1. The coeffcient cH is set equal 1 for MS-Dextra. The angles and are influenced by the bow shape. Table 4.3.2 summarizes design values for the bow door force. Table 4.3.2: Vertical design forces given by IACS-Rules

DEXTREMEL (BE97-4375) Final technical report page 38 / 75 MS-Dextra hull shape 1 2 flare angle [] 45.2 43.7 Entry angle [] 23.6 21.3 design pressure [kN/m] 152.3 139.3 Projected area Az [m2] 161.3 125.6 Vertical force Fez [kN] 24580 17510

Resulting long term predictions are shown in Fig. 4.3.17 for both bow shapes. We assumed a ration of H/HS=0.85 that resulted in a long term - exceeded once in 20 years - value for vertical bow door force which equals the IACS rule based value for bow shape 1. The same ratio was then used for bow shape 2. This approach resulted in a higher probability for bow shape 2 to exceed the IACS rule value.
MS-Dextra 6.E+04 5.E+04 vertical bow force [kN] 4.E+04 3.E+04 2.E+04 1.E+04 0.E+00 1.E-08 2x IACS Rule, bow shape 1 2x IACS Rule, bow shape 2 IACS-Rule, bow shape 1 IACS-Rule, bow shape 2 bow shape 1 bow shape 2

1.E-06

1.E-04

1.E-02

1.E+00

1.E+02

1.E+04

1.E+06

1.E+08

number of exceedance [1/year]

Figure 4.3.17: Long-term statistics of vertical bow door force The bow door of a RoRo ferry is a very complex construction. Because there is only a positive coupling and no permanent joint between the bow door and the other part of the ship, the bow door reacts very sensible against deformations. If loads exceed the strength, deformations of the locks, scantlings or hinges occur that may yield to partially or total failure of the bow door. A possible way to estimate the consequences of bow door loads was presented above. Nonlinear finite element analysis was used to estimate the deformations and stresses that arise from various load distributions. Such investigations are very time consuming and not practicable to investigate a large number of wave situations. Here we used a simplified way to rate the consequences of bow door loads. The following three categories are assumptions to estimate the consequences of bow door loads: If the vertical bow door force is lower than the IACS vertical bow door force, no damage occurs. The costs of that category are c1 = 0 [MEuro] If the vertical bow door force exceeds the IACS design force but is lower than two times the IACS design force, structural damages occurs. It is assumed that the ship survives and the damage can be repaired. The installation of a new bow door is assumed. The repair costs are estimated as c2 = 1.0 [MEuro]. If the vertical bow door force exceeds two times the design forces, heavy structural damage occurs the door is assumed to be removed/opened and the ship capsize. The resulting costs are the new building price of the vessel, c3 = 115.0 [MEuro]. Frequency of occurrence can be deduced from the long term statistical analysis shown in Fig. 4.3.17. They are listed in table 4.3.3. Multiplying the frequencies fi with their consequences ci yields the risk contributions Qi. Risk contributions are summarised in Table 4.3.4 for all consequence categories and the total annual risk that results from summation of the contributions Qi. It is shown that the main contribution to the total risk is caused by Q.2 that is due to repair costs. The risk Q3 due to total loss of the bow door was significant lower. The annual risk of bow shape 2 was higher than for bow shape 1 despite the fact that bow

DEXTREMEL (BE97-4375) Final technical report page 39 / 75 shape 2 had smaller flare angle. This effect was caused by the different evaluation of bow flare angle in current IACS rules and in the new computational method. It was shown in Fig. 4.3.5 that impact pressure coefficients were larger for bow shape 2 than for bow shape 1 due to the fact that bow shape 2 showed a more flat geometry. Table 4.3.3: Frequencies of bow door loads frequency f1 (no damage) f2 (damage) f3 (total loss) bow shape 1 3.83E+06 4.34E-02 0.00E+00 bow shape 2 3.84E+06 4.22E-01 2.12E-06 Total risk [Euro / year] 4.34E+04 4.22E+05

Table 4.3.4: Total risk of bow door loads Bow shape 1 Bow shape 2 Q1 [Euro / year] Q2 [Euro / year] Q3 [Euro / year] 0.00E+00 4.34E+04 0.00E+00 0.00E+00 4.22E+05 2.44E+02

4.4 Green water damage

4.4.1 Introduction
A ship operating in severe seas may be subject to the shipping of green water, which can result in extreme loads on the deck and deck structures including the superstructure. The term "Green Water" describes the situation when wave heights exceed the freeboard and water starts to flow over the deck, Buchner (1995). No numerical tool for the complete analysis exists as of today and, therefore, present numerical approaches to predict green water loads consider two clearly separated stages. The first step consists in the prediction of the relative wave motion at the bow. The second step comprises the prediction of a shallow water free surface flow over the deck and the possible impact at the front wall of the superstructure. Usually model tests are the only way to gather reliable information on specific designs. A design tool was created in a joint industry project using data from model tests for FPSOs, Buchner (1998). Zhou, De Kat and Buchner (1999) presented a method to predict the flow on deck and resulting loads. They used a time domain strip theory to predict ship motions coupled to shallow water equations to predict the flow over the deck. The problem of accurate prediction of green water events was highlighted in March 1998 by the announcement of the British government to reopen the formal investigation into the sinking of the MS Derbyshire that today is believed to be caused by water ingress after green water damage (UK 1998).

4.4.2 Model tests


The same model was used as in the test campaign for bow door loads, see section 4.3. To facilitate green water events, the model was equipped with the modified bow segment having small flare angle. A superstructure with so called panel force sensors was installed and tests were performed for a vertical and a sloped front bulkhead. In addition, wave gauges were installed along the deck edge, on the centerline of the deck and close to the superstructure. The installation is shown in Fig. 4.4.1. Tests were performed for different ship speeds, wave periods and wave amplitudes. Flexibility of the panel force sensors was systematically changed to study effects of fluid structure interaction.

DEXTREMEL (BE97-4375) Final technical report page 40 / 75

Fig. 4.4.1: Positions of wave height sensors (left) and pressure panel sensors (right). The experimental investigation of green water loads was carried out just after bow door experiments. The objectives for SIREHNA was to measure wave profile over the deck and to derive horizontal impact velocities on the superstructures. The experimental setup is shown in Fig. 4.4.2. As for bow door slamming, an optical set up has been implemented. It consisted in a second vertical laser sheet, generated by another assembly mirror + cylindrical lens + fiber optic, mounted inside the model, below the deck. The laser sheet was located in the longitudinal median plane. It crossed a window installed in the deck. One sheet extremity extended to the superstructure, in order to obtain profiles in the impact region. The other extremity could not reach the tip of the bow, because of the finite aperture of the laser sheet and the limited space available inside the model. In order to obtain measurements on the deck close to the bow tip, the first laser sheet generated from the bulbous bow was simultaneously used. In this purpose, a window has been mounted in the bow. However, due to the very disturbed flow around the bulb and the bow, the laser sheet was very disturbed, due to refraction on the disturbed free surface, and it could not cross the bow. Images were recorded using the same set up as used for bow door slamming. Tests were carried out with the second bow shape, for various speeds and on head sea with various wave characteristics. Besides, two superstructures have been tested : a 45 sloped and a vertical one. No major difficulty has been encountered regarding the equipment and its integration in the model. However, the measurements were again difficult, firstly because of the important three-dimensionality of the flow, mainly characterized by the creation of a jet close to the ship longitudinal axis, and secondly by the very low green water level on the deck encountered for all the test conditions considered. Examples of obtained images are presented in Fig. 4.4.2. Examples of wave front profiles and position at a given height (11m) above the deck are presented in Fig. 4.4.4.

DEXTREMEL (BE97-4375) Final technical report page 41 / 75

Figure 4.4.2: Generation of a vertical laser sheet across the ship deck

Figure 4.4.3: Examples of images obtained for green water experiments


500

600
450

550
400

500 Z ( mm )

350

X (mm)

300

450
250

400
200

350 650 600 550 500 450 400 350 300 250 200 150 100 50 0 X ( mm )

150

100 0 0 ,0 2 0 ,0 4 0 ,0 6 0 ,0 8 0 ,1 t (s ) 0 ,1 2 0 ,1 4 0 ,1 6 0 ,1 8 0 ,2

Figure 4.4.4: Examples of extracted wave front profiles and instantaeneous longitudinal position at 11m above the deck (model scale results).

4.4.3 Numerical Analysis


A standard seakeeping method was used to predict relative wave motions. Only incident waves were considered. As an example, predicted relative wave heights for test 201005

DEXTREMEL (BE97-4375) Final technical report page 42 / 75 (=0.55 rad/s, A=4.21 m, V=10 knots) are presented in Fig. 4.4.5. Predictions favorably agree with measurements for positions RB1, RB2 and RB3. However, at position RB4, computations overpredicted relative wave heights. The same spatial distribution of relative wave heights was visible in all computations performed for a ship speed of 10 knots. Results for a ship speed of 0 knots showed that the maximum relative wave height was located at the most forward part of the deck whereas measurements for a ship speed of 20 knots showed that the maximum was shifted backwards. We normalized relative wave heights with wave amplitude to facilitate the comparison of predicted and measured values for different wave conditions and ship speeds. We took the maximum of measured amplitudes for each selected time window for the comparison with predictions. These trends are summarized in Fig. 4.4.6 which shows normalized relative wave heights as a function of position along the deck edge, compare also Fig. 4.4.1. As usual, section 20 corresponded to the forward perpendicular.

Fig. 4.4.5: Comparison of predicted and measured time signals of relative wave height for test 201005. Computations succeeded to predict the effect of ship speed, i.e., the increase of normalized relative wave height with increasing speed. The decrease of normalized relative wave height towards the aft at lower speeds and the increase of normalized relative wave height towards the aft at high speed (20 knots) was not predicted. Computations showed the same spatial trend for all three ship speeds.

DEXTREMEL (BE97-4375) Final technical report page 43 / 75


Normalized rel. motion, wave freq. 0.55 rad/s, 45 deg structure

Exp. 201005 Exp. 201006

Comp. 201005 Comp. 201006

Exp. 201028

Comp. 201028

2 1,8 1,6 1,4 1,2 1 0,8 0,6 0,4 0,2 0 20 20,5 Section 20,75 21

Fig. 4.4.6: Comparison of measured and predicted normalized relative wave height for test cases with =0.55 rad/s, 45O angle of front bulkhead. Ship speeds: V=0 knots (201028), V=10 knots (201014), V=20 knots (201006). In general, results for wave frequency of =0.60 rad/s showed the same trends as results for wave frequency of =0.55 rad/s. Increasing wave frequency from =0.55 rad/s to =0.60 rad/s resulted in increasing normalized relative wave heights at a ship speed of 0 knots and in decreasing values for a ship speed of 20 knots. Close examination of model test results revealed that measured relative wave heights were smaller for the model having a vertical front bulkhead than for the model with tilted front bulkhead. This trend was strongest at the most forward located wave gauge RB1. Video recordings of the model tests showed that water was flowing backward after the impact at the front bulkhead. For a vertical bulkhead, significant amount of water was directed to the side, whereas for the titled bulkhead the water was directed upwards. This was followed by the water returning to the deck and flowing towards the bow. Thus, the relative wave height measurements at bow were affected by backward flowing water. However, the height of backward flowing water seemed not large enough to fully explain the difference in measurements at bow. To document this feature, we compared results for the model with a vertical front bulkhead with numerical predictions and with results for the model with a tilted front bulkhead. Results for test 207004 (=0.55 rad/s, A=4.29 m, V=10 knots) are presented in Fig. 4.4.7. Relative wave heights were significantly overpredicted at RB1 and RB4. Good agreement was obtained only for RB2. Comparison of results for wave frequency =0.55 rad/s and V=10 knots is given in Fig. 4.4.8. The influence of the angle of the front bulkhead on the measured normalized relative wave heights is significant. Tilting the front bulkhead backwards increased the normalized relative wave heights in particular at RB1. At a wave frequency =0.60 rad/s, difference between results for vertical and tilted bulkhead were smaller. To summarize, agreement of predicted and measured relative wave height was reasonable. The numerical method could not predict spatial variation of relative wave heights. Arrangement of front bulkhead influenced the measurements at bow, i.e., relative wave heights were smaller when front bulkhead was vertical than for backwards tilted front bulkhead. The numerical method was based on linear theory and could not take into account the bow geometry above the waterline. At the present prediction capability, we do not

DEXTREMEL (BE97-4375) Final technical report page 44 / 75 recommend to use computed relative wave heights as input for the computation of free surface flows over the fore deck.

Fig. 4.4.7: Comparison of predicted and measured time signals of relative wave height for test 207004
Normalized rel. motion, wave freq. 0.55 rad/s, speed 10 knots

Exp. 207004

Comp. 207004

Exp. 201014

Comp. 201014 2 1,8 1,6 1,4 1,2 1 0,8 0,6 0,4 0,2 0

20

20,5 Section

20,75

21

Fig. 4.4.8: Comparison of measured and predicted normalized relative wave height for test cases with =0.55 rad/s, A=4.3 m and V=10 knots. Angle of front bulkhead: 90O (207004) and 45O (201014).

DEXTREMEL (BE97-4375) Final technical report page 45 / 75 Two model test runs were selected for detailed investigations of the flow over the deck. Details are summarized in Table 4.4.1. Two angles of the front bulkhead were considered because backward flowing water seemed to influence water heights along the deck edge. Table 4.4.1: particulars of test runs selected for analysis of deck flow MARIN id 207004 201014 Wave amplitude Wave [m] [rad/s] 4.29 0.55 4.31 0.55 freq. Front bulkhead angle [deg] 90 45

Two numerical grids were generated with 18845 (coarse) and 119260 (fine) cells, respectively. Both grids had the same topology and, therefore, we refer to the grid in the following. The grid consisted of 3 blocks: one on top of the fore deck, one before and beside the bow, and one behind the bow. The complete grid system is shown in Fig. 4.4.9. The block on top of the fore deck was aligned with the deck edge to ensure that initial conditions could be accurately specified. Cells were more or less evenly distributed. The outer block was generated to match the inner block to ensure that water flowing through this artificial boundary would encounter no numerical difficulties. The block aftwards of the bow region served as outflow because initial investigations indicated that a outflow boundary condition beside the front bulkhead did not behave robustly. A time step of 0.002 s was used. The fore deck extended from the front bulkhead at frame 18.5 to the most forward point at frame 21 and it covered a longitudinal extension of 21.6 m. We then extended the grid in front of the bow by 10 m to ensure that initial conditions could be specified. To approximately model the bow shape, we added a layer of cells with a height of 5 m below the fore deck level to guarantee that water could flow off the fore deck. Beside the bow, the grid extended 10 m. In vertical direction, we also extended the grid for 10 m. The deck itself was covered with a 48000 cells (fine grid). Maximum cell dimension was approximately 0.4 m in horizontal direction and in vertical direction.

Fig. 4.4.9: Grid system for analysis of flow over deck (coarse grid with 18845 CV). Wall boundary conditions were used the deck and the front bulkhead. For all other domain boundaries, a slip-wall condition was set. At the outer domain edges, the slip wall condition is not correct but we considered this boundary condition a good solution in terms of robustness and simplicity.

DEXTREMEL (BE97-4375) Final technical report page 46 / 75 Initial conditions for the computations were difficult to establish and several attempts were made to find correct initial conditions. In principle, we took the measured relative wave heights at four locations (Rel. Bow 1 to Rel. Bow 4, see Fig. 4.4.1) as initial conditions. However, the information regarding water levels beyond the wave sensors was scarce. Video sequences were focussed on the flow over the deck and, therefore, the evolving mass of water (the green mass) in front of and beside the bow was not clearly visible. We modeled the green mass by extending the water level at the deck edge (as given by measurements) in longitudinal direction, see Fig. 4.4.10. The validity of our initial condition was later checked through a comparison of predicted and measured time traces of relative wave height at locations RB1 to RB4. Initial conditions were numerically realized by an explicit declaration of cell volume fraction in the entire domain. A constant horizontal velocity was imposed for all cells below the free surface. This velocity comprised the ship speed and a component related to the undisturbed wave.

Fig. 4.4.10: Initial condition for green water flow (two perspectives) In the following, we first present a comparison of measured and computed time traces of relative wave height at the deck edge. This corresponds to a check of the initial conditions. Second, wave heights on the deck are analyzed and wave contours are compared with measurements. Third, a comparison of predicted and measures pressures at the front bulkhead is presented. Table 4.4.2 summarizes the water levels that were specified as initial condition at the respective locations. Table 4.4.2: Initial water heights for test 207004 RB 1 2.0 m RB 2 4.5 m RB 3 3.4 m RB 4 1.8 m

Figure 4.4.11 shows time traces of wave heights for locations Rel. Bow 1 to Rel. Bow 4 which were positioned along the deck edge. A favorable agreement of predicted and measured wave height was obtained for locations RB 2 to RB 4. However, predicted wave heights decreased faster than measured wave. Predicted and measured contours of the water (projected on longitudinal centerplane) flowing over the deck are presented in Fig. 4.4.12. The first contour is associated with time t=0 s, i.e., start of the simulation. Time increment for contours was 0.1 s. Only the solution obtained on the fine grid is shown. Contours on deck did not exhibit the tongue-like jet that is associated with green water shipping of full bow forms, e.g., FPSO, see Buchner (1995). We see that shortly after the start of the simulation the wave overturned (indicated by contours crossing each other). Wave height increased at the center line after start of the simulation. This indicated an inflow from the side to the center.

DEXTREMEL (BE97-4375) Final technical report page 47 / 75

Fig. 4.4.11: Comparison of predicted and measured (run 207004) wave heights [m] at locations RB1 (top left), RB 2 (top right), RB 3 (bottom left) and RB 4 (bottom right). Solid and dashed lines: numerical predictions, dotted line: measurement.

Fig. 4.4.12: Contours of wave flowing over the deck. Top: horizontal section at deck level, middle: longitudinal section at center line, bottom: measurement at center line. Wave heights were also measured at two locations on deck and another three locations just before the front bulkhead, see Fig. 4.4.1. Agreement between predicted and measured water

DEXTREMEL (BE97-4375) Final technical report page 48 / 75 heights for location WH5 was good, see Fig. 4.4.13. The instant of time when the water level started to rise was accurately predicted indicating that the initial phase of the flow over the deck was well predicted. At location WH4, rising water levels were hardly noticeable in the model test but computations resulted in a distinct peak of water height. At locations WH2 and WH3, agreement between prediction and measurement was good only for the initial phase (at time t=298 s) when the water reached the front bulkhead. Thereafter, computed water heights increased up to 8 m which was about twice as high as measured water heights. The run up of water at the vertical front bulkhead and the determination of actual water height was considered inaccurate because of possible mixture of water with air and , overturning of the wave. Video recordings showed that during some events the water rose up to the upper edge of the front bulkhead which was 16.8 m above the deck. Therefore, we believe that both the measured and the computed time trace are correct in their own respect but cannot compared with each other.

Fig. 4.4.13: Comparison of predicted and measured (run 207004) water heights [m] at locations WH5 (top left), WH4 (top right), WH3 (bottom left) and WH2 (bottom right). Solid and dashed lines: numerical predictions, dotted line: measurement. When the jet of water hit the vertical front bulkhead, it resulted in a high pressure load. A comparison of predicted and measured pressures at locations F3C and F4C is presented in Fig. 4.4.14. Measurements were carried out with panel force transducers that yield average pressures which are representative for the panel area. For the computations, we also considered an area where pressure were averaged. It must be noted that, for the coarse grid, only 4 numerical cells represented the panel area. Agreement between predicted and measured pressure at location F3C was considered very good. The shape and the maximum value of the pressure pulse was accurately predicted. At location F4C, which is located 4 m off the center, predicted pressure was lower than measured one and the shape of the pressure pulse was differently predicted. When we consider the uncertainties associated with initial conditions and differences between predicted and measured water heights on deck, the good agreement for the predicted pressure load came as a surprise. It was found that relative wave heights at location RB 1 changed significantly when the front bulkhead of the super structure was tilted. Therefore, we investigated test 201014 that matched test 207004 very closely in all other aspects. Initial water heights were larger than for test 207004. They are listed in Table 4.4.3. Time histories of relative wave height at

DEXTREMEL (BE97-4375) Final technical report page 49 / 75 locations RB 1 and RB2 are shown in Fig. 4.4.15. Good agreement between predicted and measured wave heights was obtained. A comparison of predicted and measured pressures is presented in Fig. 4.4.16. Again, a very good agreement between predicted and measured pressures at location F3C was obtained. Agreement for location F4C was also good. The results showed that the numerical method was capable of accurately predicting pressures at the front bulkhead in response to changing initial conditions, see Fig. 4.4.14 and Fig. 4.4.16.

Fig. 4.4.14: Comparison of predicted and measured (run 207004) pressures at locations F3C (left) and F4C (right). Table 4.4.3: Initial water heights for test 201014 RB 1 6.0 m RB 2 6.0 m RB 3 6.0 m RB 4 4.0 m

Fig. 4.4.15: Comparison of predicted and measured (run 201014) wave heights [m] at locations RB 1 (left) and RB 2 (right). Solid and dashed lines: numerical predictions, dotted line: measurement.

Fig. 4.4.16: Comparison of predicted and measured (run 201014) pressures at locations F3C (left) and F4C (right).

DEXTREMEL (BE97-4375) Final technical report page 50 / 75 To summarize, time histories of water levels along the deck edge (initial conditions) were in good agreement with measured relative wave heights. Wave heights on deck were also in good agreement with measurements. However, wave heights at front bulkhead were overpredicted. The initial phase of water run-up was in good agreement with measurements and resulting pressures at center of the bulkhead were in good agreement with selected measurements. Future investigations should focus on the movement of the deck and the sensitivity of the flow to obstacles (winches, break waters, etc.). Computed pressure loads were applied to a finite element model of the fore deck and the superstructure to identify structural parts that may become weak points under extreme loads. The finite element model is presented in Fig. 4.4.17 and the same modeling was applied as for the bow door in section 4.3. Pressure pattern shown in Fig. 4.4.18 corresponded to test 202005 that exhibited the largest pressure loads in the whole campaign.

Fig. 4.4.17: Finite element model of the deck and superstructure (left) and stiffeners at the front bulkhead (right)

Fig. 4.4.18: Computed pressure distribution

DEXTREMEL (BE97-4375) Final technical report page 51 / 75


3.0E+08 Upper shell 2.5E+08 2.0E+08 1.5E+08 1.0E+08 5.0E+07 0.0E+00 0.0E+00 Lower shell 1 Lower shell 2 vm Stress [Pa] Upper stiffener Lower stiffener

5.0E+04

1.0E+05

1.5E+05

2.0E+05

2.5E+05

Max Pressure [Pa]

Fig. 4.4.19: von Mises stress vs. maximum pressure for different locations Maximum pressure loads were computed for the case with vertical front bulkhead. Applying this load (250 kPa) resulted in passing the yield stress at several locations of the FE model, see Fig. 4.4.19. Both plating and stiffeners were affected. The curve for the upper shell is derived for the most loaded element between stiffeners above the intermediate deck. The curves for the lower shell correspond to the most loaded elements between stiffeners at two locations where high pressures occur. The curves for the stiffeners correspond to locations near the flange of the mostly stressed elements above and below the intermediate deck. To verify these results, an additional finer model of the two front panels was developed. Six elements were used between stiffeners and four elements for the web of the stiffeners. Elements and resulting stresses are presented in Fig. 4.4.20. As for the shell, it is evident that the locations and values of plastic stresses are similar. Especially high plastic stresses occur at the stiffener above the intermediate deck and at the location of the crossing of the stiffener and the intermediate deck.

Fig. 4.4.20: Fine elements of front bulkhead (left) and resulting stress (right). It must be noted that the design pressure based on current rules of Germanischer Lloyd (2000) is significantly lower than the maximum pressure that was obtained in the model tests. It must also be noted that a sloped front bulkhead would result in smaller pressure loads than the vertical front bulkhead. However, no provision is made in the rules for sloped front bulkheads. Therefore, the considered case was deemed a worst case scenario.

DEXTREMEL (BE97-4375) Final technical report page 52 / 75

4.4.4 Risk analysis


Analysis of risk due to green water loads could not be completed in the course of the project. Only probability of deck flooding, i.e., the probability that wave heights exceed the deck height could be computed. Effect of bow flare were incorporated through a nonlinear correction due to Kriebel and Dawson (1993), see eqn. (4.4.1). The correction is based on a comparison of linearly predicted and measured results, see Fig. 4.4.21. The corrected distribution of amplitudes used in long-term statistics is then given by equation (4.4.2).

linear predicted response R l

model test response polynominal fit linear equal to model test response

measured response Rm

Fig. 4.4.21: Nonlinear correction of linearly predicted values.

Rl =a0 +a1Rm +a 2 Rm +a3 Rm

(4.4.1) (4.4.2)

P[F>a,v ]=

Fa 2 dRl exp 2m dRm 0

Relative motions were computed for 29 wave periods, for 7 angles of encounter and for 4 ship speeds. Nonlinear response functions were then generated for each wave height. Longterm statistical analyses were performed using H/Hs = 1 based on the analysis of a single model test in irregular waves. The same distribution of ship speeds according to wave height was used as for the analysis of bow door loads, see Table 4.3.1. The Kriebel correction was used to correct the amplitude distribution with respect to the results of the model tests. The results of long-term statistics are shown in Fig. 4.4.22. Wave gauge 1 has the largest deck flooding. With regard to 1 year, there is a height of about 6m Considering the number of deck flooding exceedence within 20 years (5.0E-02 [1/year]) a value 7.9 m for gauge 1 was computed. To determine consequences of the green water event, it is necessary to compute the flow over the deck and predict resulting pressure loads. This approach, however, could not be applied here due to a lack of information regarding the initial conditions. The first part of the initial condition is the water height above the freeboard deck or the wave contour around the deck. In our analysis, we calculated this height for each gauge using long-term statistics. Doing this we lost the link to regular waves. That means, no regular wave is assigned to the maximum water height above freeboard deck. Thus, the maximum water height above the freeboard deck of all gauges can not be assigned to a regular wave. And, therefore, the wave contour cannot unequivocally defined. The second part of the initial condition is the velocity inside the green water. This initial velocity is dependent on ship speed and the incoming waves that will be disturbed by the ship hull. It depends mainly on the bow flare of the ship that could not be taken into account using linear seakeeping method. The third part of the initial condition is the amount of water that is available to flow over the deck that is independent of the contour of green water. Especially this third part is not known because

DEXTREMEL (BE97-4375) Final technical report page 53 / 75 measurements concentrated on the wave contour along the deck edge and video recordings focused on the deck area. The result of the finite volume computation could be the time traces of pressure in the area of deck and the front wall of the superstructure. The probability of the pressure loads for each case under consideration is assumed equal to the probability of the initial conditions, i.e. the wave contour based on the long term statistical approach. Applying these pressure values to finite element models yields to strains and stresses of the ship structure. If the calculated stresses exceed the yield stresses or the ultimate strength, damages occur on the ship structure. The next step of risk assessment is the estimation of costs. This assessment contains the costs due to damage repair or the costs of total losses if the ship has capsized.
long-term statistics of green water above freeboard deck, with Kriebel correction and H/Hs=1.0
deck flooding df [m] 18 16 14 12 10 20 years value 1 year value

Gauge 1 Gauge 2 Gauge 3 Gauge 4

df1 8 df2
6 4 2 0 1.E-08 1.E-07 1.E-06 1.E-05 1.E-04 1.E-03 1.E-02 1.E-01 1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E+05 1.E+06

noe1 noe2

number of exceedance noe [1/year]

Fig. 4.4.22: Long term prediction of green water wave heights

4.5 Structural integrity

4.5.1 Numerical model and model test


A method to predict the large amplitude motions of a intact and damaged ship has been developed using Euler equations of motion of a rigid body in five degrees of freedom. The hydrodynamic and hydrostatic coefficients of the equations of motion of the intact and damaged ship are calculated as function of the instantaneous position of the vessel as it moves in waves. The sectional values of the hydrodynamic coefficients and wave exciting forces at various ship sections are obtained by means of two-dimensional source distribution technique. The hydrostatic forces i.e. buoyancy force and moment of submerged body are calculated by integration of sectional area and moment of submerged section. The dynamic effects of flooding water in a damaged compartment on ship motions are taken into account by adding time dependent mass of flooding water to the mass of the intact vessel. As it is difficult to simulate the free surface of flooding water, the sloshing effects are not considered in the formulations. For simplicity the level of flooding water is assumed to be the same height as that of the incident wave profile. The non-linear motion response equations are solved in time domain using the Runge-Kutta technique. Time-domain motion software based on the technique described above has been developed and numerical computations using this tool have been carried out to predict the dynamic motion responses of intact and damaged RoRo MS Dextra to small and large amplitude regular waves in three different headings. The wave amplitudes are 0.1, 1.5, 2.0 and 2.5m. Stern quartering waves, beam waves and head seas are considered.

DEXTREMEL (BE97-4375) Final technical report page 54 / 75 To validate the new computational method, a series of model experiments were carried out with a 1/125 scale model of MS Dextra, see Fig. 4.5.1. The model and full scale vessel characterstics are summarized in Table 4.5.1. The tests involved the measurements of 6 degrees-of-freedom motion responses as well as global loads of the stationary model in intact and damaged conditions for three different wave heading angles, four different sets of wave amplitudes at nine different regular wave frequencies, see Fig. 4.6.2. Model test results are detailed in Atlar et al. (1999 a,b). The tests also included the measurements of local pressure at a bulkhead and the damaged compartment.

Figure 4.5.1: A 1/25 scale model of MS Dextra

Figure 4.5.2: Force gauge to measure structural loading Typical results showing the roll response amplitude operators of intact MS Dextra and comparisons between the roll response amplitude operators of the intact and damaged vessel are shown in Figures 4.5.3 and 4.5.4. In the second part of this task global hull girder loads of intact and damaged RoRo hull forms have been formulated in time-domain. The hull girder loads, such as shear forces and bending moments arise from various distributions of wave-induced forces and mass inertia forces. The wave-induced forces are those due to wave excitation as well as motion responses while the mass inertia forces are due to the acceleration of the vessel.

DEXTREMEL (BE97-4375) Final technical report page 55 / 75

10.0 9.0 8.0 7.0

Theory 0.1 m Theory 1.5 m Theory 2.0 m Theory 2.5 m Experiment 1.3 m Experiment 2.0 m Experiment 2.5 m Experiment 5.0 m

4 /(k )

6.0 5.0 4.0 3.0 2.0 1.0 0.0 1 2 3

L /g
o

Roll R.A.O. of Intact RoRo ship DEXTRA at = 45 and Different Wave Amplitudes

Figure 4.5.3: Roll response of intact vessel.


6.0 5.0 4.0

Intact Damage

4 /(k )

3.0 2.0 1.0 0.0 1 2


o

L /g

Roll R.A.O. of DEXTRA at = 45 and = 2.5 m

Figure 4.5.4: Roll response of intact and damaged vessel.

The wave-induced loads at any particular cut of the hull are the resultant forces and moments of the inertia forces, hydrodynamic and hydrostatic forces on one side of the cut. These wave loads consists of compression force P1, lateral shear P2, vertical shear P3, torsion moment P4, vertical bending moment P5 and horizontal bending moment P6. Typical results showing the vertical bending moment response amplitude operators of intact M/S Dextra and comparisons between the vertical bending moment response amplitude operators of the intact and damaged vessel are shown in Figures 4.5.5 and 4.5.6. In summary, comparisons between predictions and measurements for dynamic motion responses and global wave loads at midship section of the RoRo ship in intact and damage conditions have been made and show good agreement except in roll resonant region. Nonlinear effects are significant in horizontal modes of motion and resonant roll motion but less remarkable in heave and pitch modes of motion. It is evident that most motion response

DEXTREMEL (BE97-4375) Final technical report page 56 / 75 amplitude values of the damaged ship are less than those of the intact ship except in roll resonant region. On the contrary dynamic wave load amplitudes of the damaged ship are larger than those of the intact ship.
0.025

Theory 0.1 m
0.020

Theory 1.5 m Theory 2.0 m Theory 2.5 m Experiment 1.3 m Experiment 2.0 m Experiment 2.5 m Experiment 5.0 m

P 5 /( g L 2 B )

0.015

0.010

0.005

0.000

L /g

Dynamic Vertical Bending Moment R.A.O. of Intact RoRo ship DEXTRA at = 45o and Different Wave Amplitudes

Figure 4.5.5: Vertical bending moment response of intact vessel.


0.025

0.020

P 5 /( g L B )

0.015

Intact Damage

0.010

0.005

0.000

1.5

2.5

3.5

4.5

L /g

Dynamic Vertical Bending Moment R.A.O. of RoRo ship DEXTRA at = 45o and = 2.5 m

Figure 4.5.6: Vertical bending moment response of intact and damaged vessel.

4.5.2 Ultimate structural strength


In this section the levels of safety of the damaged Ro-Ro ship MS Dextra with various damage extents are assessed. As part of residual strength assessment exercise, damage scenarios resulted from various accident events are considered. Numerical results of structural loads on a RoRo MS Dextra induced by calm water and waves in intact and various damage conditions are calculated based on the methodology described in the previous section. The residual strength of the damaged RoRo MS Dextra is calculated. The

DEXTREMEL (BE97-4375) Final technical report page 57 / 75 effects of combined vertical and horizontal bending moments, damage extents and locations on hull girder strength are discussed. Finally the structural integrity for design assessment is examined by comparison of ultimate strength and extreme structural load and conclusions are drawn. Details of structural integrity calculations are given in Chan and Incecik (1990). To assess the residual strength three damage scenarios are considered. They are engine rooms damage, fore peak flooding, and fore peak damage as described in the following:

The damage in the struck ship is assumed to occur in the Engine Room Area and two compartment flooding scenario is considered. These two adjacent compartments are the main Engine Room and Fuel Treatment Room. The damage is assumed to occur at the starboard side of frame 64 and developed symmetrically with no regards to transverse watertight bulkhead separating the two compartments and any closed location, tanks etc present in the compartments. The possible extent of damage in Engine Rooms depends on the side and mass of striking ship and the impact velocity. In compliance with SOLAS regulations (regulation 8, stability of passenger ships in damage conditions, paragraph 4), the extent of damage is assumed to be 8.61 m longitudinally, 5.2 m transversely and depth from the baseline upwards without limit (up to Sun deck). The unlimited depth of damage represents the most worse condition in which striking ship may be twice the side of struck ship. However, damage depths of 4.9 m and 13.195 m measured from the baseline other than unlimited vertical extent (24.4 m) are possibly caused by less destructive collisions and are also considered in the present study for the assessment of residual strength of damaged RoRo MS Dextra. Fore peak flooding is a scenario resulting from bow door failure or green water damage as the partial flooding of the room between the bow door and the collision bulkhead and the flooding of chain lockers and fore peak ballast tank. It is also assumed that the height of the partial flooding is one meter. Fore peak damage due to collision as a striking bow, grounding or bottom slamming may result in flooding of fore peak up to the outside draught level. The longitudinal extent of damage is assumed up to the collision bulkhead.

The structural loading due to still water loading and wave induced forces are determined and given in Table 4.5.2 in terms of the total bending moments at damaged frame 64 (51.2 m from A.P.) in stern quartering waves. Table 4.5.2: Total bending moments on damaged frame 64 in stern quartering seas
Vertical and Horizontal Bending Moments (kN-m) Intact Condition E. R. Damage Fore Peak Flooding Fore Peak Damage 437471 -259270 496617 468101 770329 -680922 1110155 -910562 739635 -653790 755967 -668227 1207799 197441 0.1635 1223831 -243451 -240769 0.9890 -342401 850886 229532 0.2698 881301 -1169832 1236252 -419360 223253 0.3585 0.1806 -1242727 1256249 -157173 -272246 1.7321 -314359 1224069 209709 0.1713 1241903 -200126 -255729 1.2778 -324727

Still Water BM Wave Vertical BM Total Vertical BM Wave Hori. BM HBM/VBM Total BM

Frame 64 is still intact in intact and fore peak flooding and damage scenarios but is damaged in engine rooms damage condition. At intact frame 64 the worst hogging moment occurs in fore peak flooding condition and the worst sagging moment takes place in intact condition. Moreover, the hogging moment at intact frame 64 is greater than the sagging moment. On the contrary the sagging moment at damaged frame 64 is greater than the hogging moment. Using local strength characteristics of stiffened panels at frame 64, the ultimate longitudinal strength of an intact ship and the residual strength of a damaged vessel can be evaluated

DEXTREMEL (BE97-4375) Final technical report page 58 / 75 based on the incremental moment-curvature method described in Smith (1977). Intact and three damage depths of 4.9, 13.195 and 24.4 m at the starboard ship of frame 64 are considered. Furthermore, one more damage location at the port side of frame 64 having unlimited damage depth (24.4 m) is also considered. It is assumed that damaged structural members are totally ineffective under in-plane loads and neglected in hull section properties. The ultimate strengths of damaged section at frame 64 are summarized in Table 4.5.3. Positive vertical bending moment is hogging and positive horizontal bending moment indicates port side in compression. The total bending moment is the resultant of horizontal and vertical bending moments in coupled bending situations. Positive total bending moment indicates starboard deck structures in tension and port side bottom structures in compression. Table 4.5.3 Ultimate bending moments (BM) on damaged frame 64
Damage Depth (m) Pure Vertical BM (kN-m) Total BM (kN-m) 0.0 (S) 3243016.2 -1866276.2 3196190.7 -1906471.5 4.9 (S) 2913611.4 -1839024.7 3034437.8 -1905198.9 13.195 (S) 2733184.2 -1834138.7 3016973.3 -1750495.8 24.4 (S) 2769145.1 -1392332.9 2549849.9 -1237245.6

Comparing the total structural loads on MS Dextra given in Table 4.5.2 with the ultimate longitudinal strengths in Table 4.6.3, the respective factors of safety for intact and damaged cross-section at frame 64 are summarized in Table 4.5.4. Table 4.5.4 Factors of safety for damaged frame 64
Intact Section 0.0 2.544 5.900 Engine Room Damages 0.0 3.627 1.534 4.9 (S) 3.443 1.533 13.195 (S) 3.423 1.409 24.4 (S) 2.893 0.996

Damage Depth (m) Total Bending Moment in Stern Quartering Seas

Hogging Sagging

The factors of safety against total collapse in intact, fore peak damage and flooding conditions in stern quartering seas are high. In particular, the safety margin in sagging mode for these conditions is twice the value in hogging mode in spite of ultimate sagging moment being smaller than ultimate hogging moment. In engine rooms damage scenario the total sagging moment becomes larger than the total hogging moment and the ultimate bending strength in sagging mode is considerably smaller than that in hogging mode. These two factors make the safety margin against hull collapse in sagging mode half the value in hogging mode. For unlimited damage depth specified by SOLAS regulations, the damaged structure at frame 64 under extreme sagging moment is unsafe, see shaded area in Table 4.5.4. The low values of safety factors against the worst sagging moment suggest that reinforcement of deck structures may be required. Finally shear forces and shear stress distribution in different damage scenarios were determined as shown in Tables 4.5.5 and 4.5.6. The factors of safety values against shear failure are also given in Table 4.5.7. Table 4.5.5: Vertical shear force in frame 64 for damage scenarios in head sea condition Intact Condition 10083.4 20242.2 30325.6 Fore Peak Flooding 11904.2 20696.3 32600.5 Fore Peak Damage 11030.2 20485.2 31515.4 Engine Room Damages -1034.0 20227.5 21261.5

Static V.S.F. (kN) Dynamic V.S.F. (kN) Total V.S.F. (kN)

DEXTREMEL (BE97-4375) Final technical report page 59 / 75 Table 4.5.6: Shear stress distribution in frame 64 for damage scenarios in head sea condition Location Intact Condition Sun deck corner 29.35 Bridge deck corner 60.35 Cabin deck corner 71.61 Public sp. deck 68.18 corner Trailer deck corner 49.38 Critical shear stress = 117.5 N/mm2 Shear Stress (N/mm2) Fore Peak Fore Peak Flooding Damage 31.55 64.88 76.98 73.29 53.08 30.50 62.72 74.42 70.85 51.32

Engine Rooms Damage 44.39 84.05 98.17 91.67 67.55

Table 4.5.7: Factor of safety against shear failure in frame 64 for damage scenarios in head sea condition Intact Condition 1.641 Fore Peak Flooding 1.526 Fore Peak Damage 1.579 Engine Rooms Damage 1.197

Safety factor

Based on the above calculations the following conclusions can be drawn:

Still water hogging bending moment increases substantially due to fore peak flooding. Thus, the fore peak flooding condition is the worst case in hogging mode. Still water hogging moment on the hull with engine rooms damaged is reduced considerably and comparable with still water sagging moment. Engine rooms damage scenario is the worst condition for the most probable extreme amplitude of dynamic wave induced vertical and horizontal bending moments. The worst condition in sagging mode is engine rooms damage scenario. Dynamic wave horizontal bending moment increases with increasing draught Damages in deck structures subjected to in-plane compressive loading reduce the residual strength Damages in starboard deck structures are more dangerous than those in port side for sagging mode in quartering waves coming from starboard side. The structural integrity of MS Dextra of current design with damaged engine rooms complied with IMO regulations may not be strong enough to withstand the worst sagging moment encountered in adverse environment. This low value of safety factor against sagging moment in damage condition suggests that risk and reliability analysis may be required.

Additional finite element analyses were carried out to simulate the bending response of damaged MS Dextra and calculate its ultimate load carrying capability. Towards this scope, the finite element model of MS Dextra, including the damage caused by the collision simulation as featured in section 4.1, was subjected to bending on the vertical plane. The objective was to compare with analytical calculations presented above, and to formulate a methodology for the modeling and analysis of such problems. The finite element model used for the simulations is the same model that has been used for the collision simulations. Since such a task, namely the calculation of the ultimate bending capability of a ship, has not been performed in the past, it was preferred to use different analysis sequences in order to define the optimum methodology that is both stable and yields

DEXTREMEL (BE97-4375) Final technical report page 60 / 75 dependable results. The initial model used for the collision simulations was consequently transferred to ABAQUS/Standard and all element properties and material definitions have been put in a format understandable to ABAQUS. To include the damage that occurred during the collision simulation, all elements that had failed where removed from the original configuration and subsequently all nodes where translated to their position after the collision. The model extended over three compartments of the vessel, namely the engine room and two compartments forward. At the foremost and aftermost sections, so called Multi-Point Constraints (MPCs) were applied, see Fig. 4.5.8, and they were used to tie the degrees of freedom of several nodes to the degrees of freedom of one node. In the case examined here, all the nodes of a boundary section were tied to a node at the center of gravity of the section, forming a rigid body. This way, it was possible to impose correct bending loads to the model: the independent node at the aftermost section was clamped while a moment was applied at the respective node of the foremost section. This moment was oriented in such a way as to impose a vertical bending moment to the ship. It should be noted, that though the boundary sections remained plane after deformation due to the imposed MPCs, this was not the case for the sections at the damaged area, since this was bounded by deformable bulkheads. For the modeling of the side shell, decks, bulkheads and web frames, full integration shell elements with six degrees of freedom were used. The stiffeners were modeled using beam elements. The material used for the analysis is elastoplastic, i.e. strain hardening after yield was accounted for. The length of the three compartments modeled was 39m. The model consisted of approximately 36500 nodes and 64000 shell and beam elements, see Fig. 4.5.7.

Fig. 4.5.7: The FE model of thethree compartments

Fig. 4.5.8: The Multi Point Constraints

DEXTREMEL (BE97-4375) Final technical report page 61 / 75


Moment 50 40 30 Pure VBM x 104 20 10 0 0,00 -10 -20 -30 curvature [m-1] hogging sagging UNEW

-1,00

-0,50

0,50

1,00

1,50

2,00

Fig. 4.5.9: Bending moment vs. curvature plots for hogging and sagging Figure 4.5.9 shows the bending moment vs. curvature and Figures 4.5.10 and 4.5.11 show the elastic and plastic energy absorbed vs. curvature. The overall response compared well with the response predicted by simplified analytical methods. Under the pure vertical bending loading the response was elastic for a curvature range from -0.5m-1 for the sagging condition to 0.5m-1 for the hogging condition. In the sagging condition, up to the point where a knuckle was observed at the bending moment vs. curvature curve, the response of the ship was mainly elastic. The knuckle was caused by the elastic buckling of the upper decks that were compressed. In the hogging condition the bending moment vs. curvature curve was very smooth there was no apparent knuckle.
Energies Hogging 1,6E+07 1,4E+07 1,2E+07 1,0E+07 Ntm 8,0E+06 6,0E+06 4,0E+06 2,0E+06 0,0E+00 0 0,5 1 curvature [m-1] PLASTIC ELASTIC 1,5 2 2,5

Fig. 4.5.10: Elastic and plastic energy curves vs. curvature for the hogging condition

DEXTREMEL (BE97-4375) Final technical report page 62 / 75


Energies Sagging 3,5E+06 3,0E+06 2,5E+06 2,0E+06 Ntm 1,5E+06 1,0E+06 5,0E+05 0,0E+00 -0,7 -0,6 -0,5 -0,4 -0,3 -0,2 -0,1 0 curvature [m-1] PLASTIC ELASTIC

Fig. 4.5.11: Elastic and plastic energy curves vs. curvature for the sagging condition

DEXTREMEL (BE97-4375) Final technical report page 63 / 75

5
Task 1.0 1.1 1.2 1.3 1.4 2.0 2.1

List of deliverables
Deliv.# D101 D111 D121 D131 D132 D141 D201 D211 D212 D213 D221 D222 D223 Title Selection of Design Cases and Operation Scenarios Collision Probability Analysis Structural Analysis of Grounding Damages on MS Dextra Collision Mechanics Assessment of damage due to grounding User Manual for the Program COLLIDE Prediction of collision and grounding damages Definition of Relevant Design Cases and Operation Scenarios Prediction of Extreme Wave Contours and Bow Door Loads Included in D222 Included in D223 Manual of Programs GLTiS and GLASiS Bow Door Load Measurement on a RoRo Ferry Results of model tests for bow door loads : Wave contours Structural response of bow doors under slamming loading Prediction of Bow Door Damages Included in D201 Prediction of Green Water Wave Heights Included in D322 Included in D323 Prediction of Green Water Loads Green Water Load Measurement on a RoRo Ferry Results of model tests : green water flow Finite Element Analysis of Green Water Loading on Deck and Deckhouse Prediction of Green Water Damages Included in D431 Prediction of Large Amplitude Motions and Wave Loads on a RoRo ship in Regular Oblique Waves in Intact and Damage Conditions Update of D411 Manual of Program QNLMWLA Small Scale Tests for the Measurements of Motions with an Intact and Damaged Model Small Scale Tests for the Measurements of Global Loads with an Intact and Damaged Model Residual Strength of a Damaged RoRo Ship Canceled Residual Structural Strength of Damaged Ship Structures Prenormative Guidelines for Analysis of Structural Safety under Extreme Loads for RoRo-passenger ships Document ID DTR-1.0-AESA-03.98 DTR-1.1-DTU-11.98 DTR-1.2.1-DTU-04.99 DTR-1.2.2-DTU-05.99 DTR-1.3-NTUA-10.99 DSD-1.3-DTU-11.99 DTR-1.4-GL-01.00 DTR-2.0-AESA-07.98 DTR-2.1-GL-03.99 Planned 3 12 18 21 21 24 7 12 Actual 4 12 17 18 23 26 8 16

2.2

DSD-2.2-GL-10.99 DMT-2.2-MARIN-05.00 DMT-2.2-SIR-08.99 DTR-2.3-NTUA-11.99 DTR-2.4-GL-03.00 DTR-3.1-GL-04.00 DTR-3.2-GL-03.00 DMT-3.2-MARIN-01.01 DTR-3.2-SIR-11.99 DTR-3.3-NTUA-05.00 DTR-3.4-GL-11.00 DTR-4.1-UNEW-12.98

18 18 18 22 25 8 13 19 19 19 23 26 27 12

23 30 21 24 28 8 29 28 37 24 30 36 29 13

2.3 2.4 3.0 3.1 3.2

D231 D241 D301 D311 3.1.2 3.1.3 D321 D322 D323 D331 D341 D401 D411

3.3 3.4 4.0 4.1

4.2

D412 D413 D421

DSD-4.1-UNEW-01.00 DTR-4.2-UNEW-03.99 DTR-4.2-UNEW-11.99 DTR-4.3-UNEW-04.00 DTR-4.5-UNEW-12.00 DPG-4.5-GL-11.00

18 24 15 21 33 33 36 36

18 26 16 24 29 na 37 36

4.3 4.4 4.5

D431 D441 D451 D452

DEXTREMEL (BE97-4375) Final technical report page 64 / 75

6 Comparison of initially planned activities and work actually accomplished.


6.1 Task 1: collision and grounding loads
Two computer programs and accompanying manuals were developed. Software and reports constitute the main results of this task. Overall performance in terms of initially planned and actually accomplished results Not acceptable
Objectives not fulfilled

Fair

Good

Very good

Excellent
All objectives fulfilled

Some objectives fulfilled / partly fulfilled / not fulfilled

Subtask 1.0

Definition of relevant design scenarios

Objectives fulfilled

Subtask 1.1

Development of probabilistic models for collision Objectives partly fulfilled and grounding event

Probabilistic model for the prediction of grounding events could not be developed. Main reason was not mathematical difficulty but the problem to obtain reliable data on sea bed topology along the route of the vessel. A detailed and accurate description of available water depths together with the widths of the traffic lane is now considered minimum requirement for the prediction of grounding frequency.

Subtask 1.2

Development of models for external ship collision Objectives fulfilled and grounding dynamics and for internal ship structure dynamics

Subtask 1.3

Establishment of damage probability distributions Objectives fulfilled

Subtask 1.4

Application of developed models to evaluate Objectives partly fulfilled specific ship designs

Developed software for analysis of collision events could only be applied to the existing reference ship. No investigation into alternative structural designs was accomplished. Risk analysis for grounding considered only an artificial scenario due to a lack of seabed data, see subtask 1.1.

DEXTREMEL (BE97-4375) Final technical report page 65 / 75

6.2

Task 2: bow door loads

Two sets of model test data were collected and one computer program with accompanying manual was developed. Data, software and reports constitute the main results of this task. Overall performance in terms of initially planned and actually accomplished results Not acceptable
Objectives not fulfilled

Fair

Good

Very good

excellent
All objectives fulfilled

Some objectives fulfilled / partly fulfilled / not fulfilled

Subtask 2.0

Definition of relevant design scenarios

Objectives fulfilled

Subtask 2.1

Prediction of extreme wave contours

Objectives fulfilled

Subtask 2.2

Prediction of loads on bow doors

Objectives fulfilled

Subtask 2.3

Structural response of bow door, looks and bow

Objectives partly fulfilled

Only the original structural arrangement of the bow door was used in the analysis because i) an alternative structural design was not available and ii) the effort required to generate a finite element model for bow and bow door which was underestimated in the project plan.

Subtask 2.4

Application of developed models to evaluate Objectives partly fulfilled specific ship designs

The risk analysis of the bow door could not be fully exploited due to the missing criteria on ultimate failure of the bow door. Therefore, results were only considered as first step to demonstrate the applicability of the approach.

DEXTREMEL (BE97-4375) Final technical report page 66 / 75

6.3

Task 3: green Water loads

Two sets of model test data were collected. Data and reports constitute the main results of this task. Overall performance in terms of initially planned and actually accomplished results Not acceptable
Objectives not fulfilled

Fair

Good

Very good

excellent
All objectives fulfilled

Some objectives fulfilled / partly fulfilled / not fulfilled

Subtask 3.0

Definition of relevant design scenarios

Objectives fulfilled

Results of this subtask were included in the report of subtask 2.0.

Subtask 3.1

Prediction of green water wave heights

Objectives partly fulfilled

Comparison of model test results and predictions showed unsatisfactory agreement but no clear trend to refine the numerical method. Therefore, no modifications were made. Instead, results of model tests were later used to tune the numerical method for use in subtask 3.4

Subtask 3.2

Prediction of loads on deck structures

Objectives partly fulfilled

Investigations did not include the effects of smaller deck structures, e.g., winches and breakwaters, on the flow over the deck. Effects of ship motion on the flow were also not studied. The reason was a lack of resources due to the complexity encountered in the proper set up of the problem which was not foreseen in the project plan.

Subtask 3.3

Structural response of deck structures

Objectives fulfilled

Subtask 3.4

Application of developed models to evaluate Objectives partly fulfilled specific ship designs

Risk analysis was not fully completed because the prediction of consequences failed. To determine consequences of green water shipping, the flow over the deck has to be computed for predefined initial conditions that correspond to probability levels determined before. As the probability of occurrence was computed using a long term probabilistic method, information on specific wave situations was lost. No solution was found to solve this problem.

DEXTREMEL (BE97-4375) Final technical report page 67 / 75

6.4

Task 4: residual structural strength of damaged ship structures

Model test data were collected and one computer program with accompanying manual was developed. Data, software and reports constitute the main results of this task. Overall performance in terms of initially planned and actually accomplished results Not acceptable
Objectives not fulfilled

Fair

Good

Very good

excellent
All objectives fulfilled

Some objectives fulfilled / partly fulfilled / not fulfilled

Subtask 4.0

Review of extreme loads predicted from models Objectives fulfilled in tasks 1,2 and 3

Results of this subtask were included in the report of subtask 4.3.

Subtask 4.1

Prediction of large amplitude motions and Objectives fulfilled resulting structural response of a damaged RoRo ship

Subtask 4.2

Validation of developed method with small scale Objectives fulfilled model tests

Subtask 4.3

Prediction of residual structural strength and Objectives fulfilled quantification of risk of (total) loss of structural integrity by parametric study

Subtask 4.4

Risk acceptance criteria for evaluation improvement of safety measures

of Objectives not fulfilled

Partners decided to cancel this subtask due to a lack of available data. Efforts are ongoing in several European funded research projects and thematic networks to define socially acceptable risk acceptance criteria for maritime transport.

Subtask 4.5

Formulation of guidelines and recommendations

Objectives fulfilled

Note: Identified research needs will be included in the final technical report.

DEXTREMEL (BE97-4375) Final technical report page 68 / 75

7 Management and co-ordination aspects


7.1 Performance of consortium
All partners of the DEXTREMEL consortium were highly motivated during the project's duration. However, some partners having a smaller share of research work in the end did not contribute as efficiently as the other partners that still had work in the final phase of the project. All partners were dedicated to the success of the project and contributed as expected. Discussions at project meetings were fruitful and added value to the results of the reports. All deliverables were distributed in time either according to the original work program or following announcements of delay. However, two deliverables were not available at the official end date of the project.

7.2

Organization and communication

Organization and communication in the project was carefully planned and a project manual was prepared by the project coordinator following the kick-off meeting. It was agreed on by the partners after the first month of the project's duration. Representatives of all contractors formed the project steering committee. All decisions were taken unanimously. All draft versions of periodic progress reports were issued in advance of the steering committee meetings. The project coordinator as representative of the coordinating partner was responsible for the management, administration and control of the project. This included acting as first contact to and issuing reports to the European Commission,

preparing periodic progress reports and consolidated cost statements, coordinating the exchange of documents and information between the partners, announcing and preparing the meetings and distributing the minutes, monitoring the progress and schedule of the project by means of 3-monthly informal progress reports required from all partners A project management task with an allocation of 5 man months (3% of the man months total) for the project coordinator was added to the technical tasks. However, in particular preparations of progress reports consumed more time than initially planned.
Effective communication was ensured using email from the beginning of the project. All documents were exchanged in electronic format. An Internet homepage was established in the first 6-monthly period of the project. The homepage served project partners as electronic document library and presented the project and the partnership to the public.

DEXTREMEL (BE97-4375) Final technical report page 69 / 75

7.3

List of contact persons for follow up


Dr. P.C. Sames Tel.: +49 40 36149 113 Fax.: +49 40 36149 7320 Email: pcs@germanlloyd.org Prof. P.T. Pedersen Tel.: +45 45 25 1386 Fax.: +45 45 88 4325 Email: ptp@ish.dtu.dk Dr. J.O. de Kat Tel.: +31 317 493911 Fax.: +31 317 493245 Email: j.o.dekat@marin.nl Prof. M. Samuelides Tel.: +30 1 772 1421 Fax.: +30 1 772 1412 Email: msamuel@deslab.ntua.gr

Germanischer Lloyd Vorsetzen 35 D-20459 Hamburg Germany Technical University of Denmark Department of Ocean Engineering Building 101 E DK-2800 Lyngby Denmark MARIN 2, Haagsteg P.O. Box 28 NL-6700 AA Wageningen The Netherlands National Technical Universtiy of Athens Dept. Naval Architecture and Marine Engineering 9 Heroon Polytechniou GR 15773 Zografou - Athens Greece SIREHNA 1 rue de la No BP 42105 F-44321 Nantes Cedex 3 France University of Newcastle upon Tyne Department of Marine Technology Armstrong Building Newcastle NE1 7RU United Kingdom Astilleros Espaoles Ochandiano 16 ES-28023 Madrid Spain

Dr. P. Corrignan Tel.: +33 2 51 86 0284 Fax.: +33 2 40 74 1736 Email: Philippe.Corrignan@sirehna.ec-nantes.fr Prof. A. Incecik Tel.: +44 191 222 6724 Fax.: +44 191 222 6724 Email: atilla.incecik@ncl.ac.uk Dr. C. Arias Tel.: +34 1 387 81 38 Fax.: +34 1 387 81 16 Email: carias@astilleros.net

DEXTREMEL (BE97-4375) Final technical report page 70 / 75

Results and Conclusions

Project work closely followed the work program. Therefore, all main objectives were fulfilled. Although significant delays occurred in one task, the overall success of the project was not endangered. Smaller delays were accommodated through favorable planning that concentrated work in year 2 of the project. This buffer proved were valuable and, thus, the project could be finished as planned. In the following all results are summarized starting with the reference vessel. Then results of collision and grounding analyses are presented followed by results of bow door load and green water load investigations. Finally, structural integrity of the vessel is reviewed. Identified research needs are summarized in the end. A reference ship was defined in the beginning of the project. A so called "fast conventional" RoRo passenger ferry was selected. The ship had a length of 173 m and a service speed of 28 knots. Two reference routes were defined but subsequent analysis focused on the route from Cadiz to the Canary Islands. Wave climate and traffic data were generated only for this route. It turned out that accurate traffic data was difficult to collect. Analysis of collision events succeeded as planned. A new method was developed and coded to predict collision probability. It takes into account actual traffic data and effects of crew behavior and weather conditions can be incorporated. The program was successfully applied to predict the collision probability of the reference ship. Damage following the collision was predicted in terms of probabilistic distributions for damage size and location. The integration of these results into a risk analysis framework showed that one crucial component is still missing to fully exploit the potential of risk analysis. This component was an accurate and probabilistic method to predict damage stability of the damaged vessel. Analysis of grounding events was restricted to predict damage probability distributions. Prediction of grounding probability was not feasible due to a lack of seabed topology data along the route that is needed to determine when available water depth is less than the ship's actual draft. A new method was developed and coded to predict damage probability distributions for hard grounding taking into account structural arrangement of the double bottom and obstacle geometry. Integration into a risk analysis was performed assuming an artificial island along the route as obstacle. Therefore, results of the risk analysis can only be considered as example. In addition, the lack of a tool to predict damage stability limited the value of the consequence predictions. Loads and structural response of bow doors were analyzed using numerical methods and model tests. A new method was developed to predict loads on the bow door in regular waves taking into account impact pressures. This method was then validated using the custom made model tests that included measurements of the force on the bow door by means of a bow segment attached to the rest of the model. In addition, local pressures were recorded. The most innovative feature of the test campaign consisted in laser augmented measurements of the wave contour at the bow. A laser system was installed onboard the model and light was reflected upwards through the bulbous bow to form a sheet of laser light. Digital processing of the recorded images yielded the wave contours as a function of time. The new computational method was successfully validated and systematically employed to generate input data for the risk analysis of the bow door. Consequences of bow door loads, i.e., damages or total failure, could not exactly be predicted using nonlinear finite element analysis due to the complexity of the locking system. However, applying simplified consequence criteria the effect of bow flare was analyzed in the risk analysis. Investigations on green water shipping centered on prediction of loads and structural response of the front bulkhead of the superstructure. Again, model tests were performed to yield data for validation of computational methods. Relative wave height along the edge of

DEXTREMEL (BE97-4375) Final technical report page 71 / 75 the fore deck and wave elevation on the fore deck was measured. Loads on the front bulkhead were measured using panel sized force transducers varying the stiffness. The flow over the fore deck was visualized using the same sheet of laser light as before but now directed upwards through a slit in the deck. However, digital processing of the images was difficult due to overturning waves and associated bubbles of air that scattered the laser light. The numerical method to predict relative wave height was not capable of predicting all features of the flow around the bow. However, a second numerical method that predicted the flow over the deck could successfully predict the loads on the front bulkhead. Analysis of the structural response showed that vertical front bulkheads may suffer from large pressure loads. The risk analysis of green water induced structural failures could not be completed. Problems in the specification of initial conditions for the flow over the deck were not be solved in this project. Analysis of structural integrity of damaged ships in waves started with the development and programming of a new method to predict large amplitude motions of a damaged ship in waves. A dedicated series of model test runs were conducted to yield measurements for validation of the new numerical method. A two-compartment damage including the engine room was selected as worst case. Using a second numerical method, ultimate strength of the main hull girder was analyzed. Results were cross-checked with elaborate finite element analysis. It was shown that a very large damage (from keel to upper cabin deck, resulting from, e.g., a collision) and extreme wave conditions resulted in a slightly larger vertical bending moment that the hull girder was capable to withstand. However, the assumed scenario was considered unlikely to occur. Prenormative guidelines were established that summarize the experience gathered during the project. For each hazard that was investigated in the project, guidance is given on data requirements, application of numerical methods and the processing of results. Research needs emerged during the work performed in the project. They are summarized in the following. To fully exploit risk analysis for collision and grounding events, an accurate method to predict damage stability of the damaged vessel in a probabilistic framework is needed. In addition, systematic collection of traffic data is a prerequisite for collision analysis. Information on seabed topology is required input for grounding analysis. In the future, both data sets may be collected by means of satellite observation techniques. To make risk analysis work for bow doors and superstructures, a reliable method to predict ultimate strength is needed in a probabilistic framework. However, it is assumed that for analysis of bow doors no tool short of a full blown nonlinear finite element analysis is capable of predicting ultimate failure. Prediction of green water loads depends on correct initial conditions for the flow over the deck. Measurements of the so called "green mass" - that is ready to flow over the deck - through an array of wave gauges at the bow is required together with a refinement of numerical tools.

Acknowledgements

Project partners thank the European Commission for funding DEXTREMEL. We also thank the project officers Dr. E. Campogrande and Dr. T. Fairley for the support during the project's duration. The success of the project would not have been possible without the efforts of our colleagues at the respective organizations: GL: Dr. C. stergaard, Mr. S. Otto, Mr. H. Rathje, Mr. C. Schiff, Dr. L. Zhang. DTU: Dr. A.M. Hansen, Ass. Prof. P. Friis Hansen, Mrs. M. Ltzen, Ass. Prof. B. Cerup Simonsen, Mr. R. Trnquist, and Dr. S. Zhang. NTUA: Mr. D. Servis, Ms. T. Louka, Mr. G. Voudouris, Mr. G. Katsaounis UNEW: Drs. M. Atlar, H.S. Chan, E. Korkut and Mr. D. Lamb.

DEXTREMEL (BE97-4375) Final technical report page 72 / 75

10

References

Adegeest, L. J. M. (1994): Nonlinear hull girder loads in ships , PhD Thesis, Faculty of Mechanical Engineering and Marine Technology, Delft University of Technology. Borresen, R. & Tellsgard, F.(1980): Time history simulation of vertical motions and loads on ships in regular head waves of large amplitude. Norwegian Maritime Research, 3. Buchner, B. (1998), A New Method for the Predictions of Nonlinear Wave Motions, in Proceedings of the 17th international Conference on Offshore mechanics and Arctic Engineering (OMAE), Lisbon, Portugal. Buchner, B. (1995), On the Impact of Green Water Loading on Ship and Offshore Unit Designs, in Proceedings of the 6th International Symposium on Practical Design of Ships and Mobile Units (PRADS), Seoul, South Korea. Chan, H. S. and Incecik, A. (2000): Structural Integrity of a Damaged Ro-Ro Ship , Proceedings of the International Conference on Marine Design and Operations for Environmental Sustainability, Newcastle upon Tyne. Chen, Y. K., Kutt, L. M., Piaszczyk, C. M. and Binek, M. P. (1983): Ultimate strength of ship structures, Transaction of SNAME, Vol. 91, pp. 149-168, 1983. Chiu, F. C. & Fujino, M., (1989): Nonlinear prediction of vertical motions and wave loads of high-speed crafts in head sea International Shipbuilding Progress, 36 , 193-232. Clauss, G., Lehman, E., stergaard, C. (1994), Offshore Structures, Volume 2, Springer. Dow, R. S., Hugill, R. C., Clark, J. D. and Smith, C. S. (1981): Evaluation of ultimate ship hull strength, Proceedings of SSC-SNAME Extreme Loads Response Symposium, Arlington, Virginia, pp. 133-148, 1981. Fang, M. C. & Her, S. S. (1995): The non-linear SWATH ship motion in large longitudinal waves. International Shipbuilding Progress, 42 , 197-220. Fang, C. C., Chan, H. S. & Incecik, A.(1997): Investigation of motions of catamarans in regular waves II. Ocean Engineering, 24 , 949-66. Faulkner, D. (1975): Compression strength of welded grillages, chapter 21 in Ship Structural Design Concepts, edited by J.H. Evans, Cornell Maritime Press. Fekken, G., Veldmann, A.E.P. and Buchner, B. (1999): Simulation of Green Water Loading Using the Navier-Stokes Equations, presented at 7th Int. Conference on Numerical Ship Hydrodynamics, Nantes, July 1999. Fujino, M. and Yoo, B. S. (1985): A study on wave loads acting on ship in large amplitude (3rd Report). Journal of the Society of Naval Architects of Japan, 158. Fujii, Y. and Mizuki., "Design of VTS Systems for water with bridges", Proceedings international Symposium on Advances in Ship Collision Analysis, Copenhagen, 1998 Ghose, D. J., Nappi, N. S. and Wiernicki, C. J. (1994): Residual strength of damaged marine structures, Ship Structure Committee Report SSC-381. Guedes Soares, C., Schellin, T. (1995), Long-Term Distribution of Nonlinear Wave Induced Vertical Bending Moments on a Containership, Marine Structures, Vol. 9, pp. 333-352. Hansen, P. Friis and Pedersen, P. Terndrup (1998): Risk Analysis of Conventional and Solo Watch Keeping Presented at Int. Maritime Organisation (IMO) Maritime Safety Committee by Denmark at the 69th Session. Kinoshita, T. Kagemoto, H. and Fujino, M. (1999): A CFD Application to Wave-induced Floating-body Dynamics, presented at 7th Int. Conference on Numerical Ship Hydrodynamics, Nantes, July 1999.

DEXTREMEL (BE97-4375) Final technical report page 73 / 75 Kriebel, D. L. and Dawson (1993), T. H., Nonlinearity in wave crest statistics, 2nd intl. symposium on ocean wave measurement and analysis, pp.6175. Kvlsvold, J., Svensen, T. and Hovem, L. (1996), Bow Impact Loads on RoRo Vessels, in Transactions of the Royal Institution of Naval Architects (RINA), London, United Kingdom. Mansour, A. E. and Thayamballi, A. (1980): Ultimate strength of a ships hull girder in plastic and bending modes, Ship Structure Committee Report SSC-299. Minorsky, L. (1959): An analysis of ship collisions with reference to protection of nuclear power plants, Journal of Ship Research, Journal of Ship Research, Vol. 3, No. 2, pp. 1-4. Ortloff, C.R. and Krafft, M.J., Numerical Test Tank: Simulation of Ocean engineering Problems by Computational Fluid Dynamics, in Proceedings of the Offshore Technology Conference (OTC), May 1997, Houston, Texas, USA. stergaard, C. and Schellin, T.E. (1995), Development of an Hydrodynamic Panel Method for Practical Analysis of Ships in a Seaway. Trans. STG, Vol. 89, pp. 147-156. stergaard, C., Rathje, H. and Sames, P.C. (1996), RoRo Ship Bow Door Design: First Principle Analysis of Wave Loads and Stresses, in Proceedings of the 13th International Conference & Exhibition on Marine Transport Using Roll-on/Roll-off and Horizontal Handling Methods (RORO96), Lbeck, Germany. Paik, J.K. (1994): Cutting of a longitudinally stiffened plate by a wedge, Journal of Ship Research, Vol.38, No.4, pp. 340-348. Paik, J.K., and Wierzbicki, T. (1997): A benchmark study on crushing and cutting of plated structures, Journal of Ship Research, Vol. 41, No. 2, pp. 147-160. Pedersen, P. Terndrup (1995): "Collision and Grounding Mechanics" Proc. WEMT'95, pp. 125157, Copenhagen. Pedersen, P. Terndrup and Zhang, S. (1998): "Impact Mechanics of Ship Collisions" Marine Structures Vol. 11 1998, pp. 429-449 Pedersen, P. Terndrup and Zhang, S. (1999):"Collision Analysis for MS Dextra" SAFER EURORO Spring Meeting, Nantes, France, April, Paper No. 2, pp. 1-33. Ratje, H., Schellin, T., Otto,S., stergaard, C. (2000), Predicting Nonlinear Wave Induced Design Loads for Ships, Proc. OMAE 2000 Joint Conference, OMAE 00-6122, New Orleans. Rutherford, S. E. and Caldwell, J. B. (1990): Ultimate longitudinal strength of ships: a case study, Transaction of SNAME, Vol. 98. Sames, P., Schellin, T., Muzaferija S., Peric, M. (1998), Application of a Two-Fluid Finite Volume Method to Ship Slamming, in Proceedings of the 17th International Conference on Offshore Mechanics and Arctic Engineering (OMAE), Lisbon, Portugal. Sames, P.C., Muzaferija, S. and Rathje, H. (1998): Computation of impact loads for ships in waves (in German), presented at STG Hydrodynamic Committee Meeting, Hamburg, Sept. Samuelides, M. (1984): Structural Dynamic and Rigid Body Response coupling in Ship Collisions, PhD Thesis, Glasgow University. Samuelides, M. (1999), Prediction of oil outflow based on energy considerations, Journal of Ship Research, Vol. 43, No. 3, pp. 194-199. Servis, D. P. (1997), Development of a model for the simulation of ship ship collisions, National Technical University of Athens, Final year thesis (in Greek). Servis, D.P. and Samuelidis, M. (1999): Ship Collision Analysis using Finite Elements SAFER EURORO Spring Meeting, Nantes, France, April, Paper No. 3

DEXTREMEL (BE97-4375) Final technical report page 74 / 75 Simonsen, B.C (1997a): Mechanics of ship grounding, PhD thesis, Technical University of Denmark, Department of Naval Architecture and Offshore Engineering. Simonsen, B.C (1997d): Ship grounding on rock: I Theory, , Marine Structures, Vol. 10, pp. 519-562, 1997. Simonsen, B.C (1997c):, Ship grounding on rock:I I Validation and application, Marine Structures, Vol. 10, pp. 563-584, 1997.. Wang, G. and Ohtsubo, H. (1995): Energy absorbing mechanisms involved in grounding, MARIENV95, pp. 81-85. Skjong, R. and Ronold, K.O. (1998): Societal Indicators and Risk Acceptance, presented at 17th Int. Conf. On Offshore Mechanics and Arctic Engineering, paper OMAE98-1488, Lisbon. Smith, C. S. (1977): Influence of local compressive failure on ultimate longitudinal strength of a ships hull, Proceedings of International Symposium on Practical Design in Shipbuilding, Tokyo, pp. 73-79. Smith, C. S. and Dow, R. S. (1981): Residual strength of damaged steel ships and offshore structures, Journal of Constructional Steel Research, Vol. 1, pp. 1-15. Tao, Z. & Incecik, A.(1996): Large amplitude ship motions and bow flare slamming pressures in regular head seas. In proceedings of 15th International Conference on Offshore Mechanics and Arctic Engineering, Vol. I, pp. 16-20. Turan, O. & Vassalos, D. (1994): Dynamic stability assessment of damaged passenger ships. Transactions RINA, 136, 79-104. Wang, G. Ohtsubo, H. and Liu, D. (1997): A simple method for predicting the grounding strength of ships, Journal of Ship Research, Vol.41, No.3, pp. 241-247. Yamamoto, Y., Fujino, M. & Fukasawa, T.(1978): Motion and longitudinal strength of a ship in head sea and effects of nonlinearities, (1st & 2nd Reports). Journal of the Society of Naval Architects of Japan, 143 & 144 . Zhou, Z.Q., De Kat, J.O. and Buchner, B. (1999): A nonlinear 3D approach to Simulate Green Water Dynamics on Deck, presented at 7th Int. Conference on Numerical Ship Hydrodynamics, Nantes, July 1999. Final report on the capsizing on 28 September 1994 in the Baltic Sea of the Ro-Ro passenger vessel MV ESTONIA. The Joint Accident Investigation Commission of Estonia, Finland and Sweden. Edita Ltd., Helsinki 1997, ISBN 951-53-1611-1. Report on the sinking of the MS Derbyshire, by UK/EC assessors, ISBN 1-85112-072-6. FSEA (1996): Formal Approaches to Risk Assessment for Seaborne Transport in European Waters, Concerted Action WA-96-CA1155. GL Rules (1999), GL Rules for Classification and Construction, Part 1 Seagoing Ships Global Wave Statistics (1990), PC GWS V2.10, BMT Fluid Mechanics Ltd IACS (1999), IACS Blue Book, IACS Unified Requirements S8, Rev 2 Joint North West European Project (1996):Safety Assessment Methodology of Passenger/RoRo Vessels, DNV Technical Report. Maritime Industries Forum (1999): The Maritime Industry R&D Masterplan 1999, R&D Coordination Group. SAFER EURORO (1997): Design for Safety: An Integrated Approach to Safe European RoRo Ferry Design, thematic network, ERB BRRT-CT97-5015. SOLAS, "International Convention for the Safety of Life at Sea" (SOLAS) 1974, THEMES (2000): Safety Assessment in Waterborne Transport, thematic network, 2000TN.11080. .

DEXTREMEL (BE97-4375) Final technical report page 75 / 75

Appendix
Wave scatter diagram for MS-Dextra sailing between Cadiz and Tenerife
P [Hs,Tz] 14.5 13.5 12.5 11.5 10.5 9.5 8.5 7.5 6.5 5.5 4.5 3.5 2.5 1.5 0.5 3.5 0.000000 0.000000 0.000000 0.000000 0.000000 0.000000 0.000000 0.000000 0.000000 0.000000 0.000000 0.000002 0.000020 0.000185 0.001184 4.5 0.000000 0.000000 0.000000 0.000000 0.000000 0.000000 0.000000 0.000000 0.000000 0.000004 0.000030 0.000210 0.001300 0.006117 0.010522 5.5 0.000000 0.000000 0.000000 0.000000 0.000000 0.000001 0.000002 0.000006 0.000030 0.000147 0.000753 0.003655 0.014748 0.039042 0.026885 6.5 0.000000 0.000000 0.000001 0.000002 0.000004 0.000011 0.000033 0.000108 0.000387 0.001487 0.005734 0.020124 0.055131 0.087057 0.028082 Zero upcrossing period Tz 7.5 8.5 0.000002 0.000012 0.000003 0.000011 0.000006 0.000022 0.000013 0.000045 0.000031 0.000098 0.000076 0.000222 0.000202 0.000533 0.000580 0.001359 0.001771 0.003630 0.005605 0.009777 0.017177 0.024715 0.045808 0.052092 0.089127 0.075027 0.088231 0.048983 0.014419 0.004259 9.5 0.000028 0.000023 0.000042 0.000081 0.000163 0.000339 0.000745 0.001715 0.004067 0.009525 0.020368 0.034920 0.038516 0.017368 0.000838 10.5 0.000038 0.000027 0.000048 0.000088 0.000165 0.000321 0.000648 0.001360 0.002901 0.006000 0.011047 0.015744 0.013677 0.004436 0.000124 11.5 0.000035 0.000022 0.000038 0.000065 0.000115 0.000210 0.000394 0.000761 0.001473 0.002720 0.004375 0.005279 0.003700 0.000893 0.000015 12.5 0.000024 0.000014 0.000022 0.000037 0.000061 0.000104 0.000182 0.000326 0.000578 0.000962 0.001369 0.001420 0.000819 0.000151 0.000002 13.5 0.000013 0.000007 0.000010 0.000016 0.000026 0.000042 0.000069 0.000114 0.000186 0.000282 0.000358 0.000324 0.000156 0.000022 0.000000

Sign. Wave Height Hs m]

Note: all directions, all seasons

You might also like