You are on page 1of 29

MarineStructures8 (1995) 451-479

1995 ElsevierScienceLimited
Printed in Great Britain.All rights reserved ELSEVIER

095i -8339/95/$9.50

The Vessel in Port: Mooring Problems

T. E. Schellin & C.

Ostergaard

Germanischer Lloyd, 20149 Hamburg, Germany


(Received 19 July 1993)

ABSTRACT Discussed are problems concerning safe mooring of ships in port, comprising environmental forces applied to the ship, general principles that determine how the applied forces are distributed to the mooring lines, and application of these principles to establish a good mooring arrangement. Z~sues addressed include an overview of good mooring principles, guidance for berth layout and shipboard mooring equipment, and prediction of wind and current loads on the ship. Emphasis is placed on computer prediction techniques to assess effects of moored ship dynamics caused by the natural seaway. As an example, calculated maximum surge excursions of a large tanker moored to an exposed jetty and subject to the action of long-crested natural seaways are presented for seastates having significant wave heights ~,f up to 5 m and zero-upcrossing periods of up to 11.6 s. Restoring force characteristics in surge were assumed linear. Representative restoring force constants ranging from 500 to 4000 kN/m corresponded to natural surge periods ranging from 52 to 148s. Frequency domain analyses were performed using spectral expressions of wave drift force. Expected maximum surge motions of the moored tanker are plotted as functions of significant wave height. Key words." m o o r i n g , safety, ship in p o r t , surge, seaway, drift.

1 INTRODUCTION The safe mooring of a ship in port depends on factors such as size and type of the moored ship, variety and characteristics of available mooring
451

452

T. E. Schellin, C. Ostergaard

ropes and equipment, terminal layout requiring flexible ship mooring plans, and physical conditions at the port facility ranging from a protected pier to an offshore sea island exposed to wind, current and wave forces. Numerous standards, guidelines and recommendations exist in the worldwide marine industry concerning mooring practices, mooring fittings and mooring equipment. The most widely used standards are the International Standards (ISO), the British Standards (BS) and the Japanese Industrial Standards (JIS). These requirements, however, sometimes lack essential information such as strength of fittings, and their non-uniformity and variations in interpretations of practices has resulted in the misunderstanding of the complete vessel mooring system by both designers and ship and terminal operators. To establish an operational standard that ensures the ship is properly equipped for safe mooring, it is necessary, depending on the probable trading route, to determine environmental forces that would normally be experienced by the moored ship. Only then can the ship be properly equipped with a mooring system of sufficient restraint capability to resist these forces and to allow some degree of flexibility within normal safety tolerances. Therefore, strict rules for mooring systems are not specified by international regulatory bodies such as the United Requirements of the International Association of Classification Societies (IACS). However, Classification Societies specify minimum recommendations for mooring ropes in their rules, but compliance is not a condition of class. ~ Recommendations for hawsers are based on empirical tables derived from an equipment numerical obtained using displacement and dimensional criteria. Moreover, classification rules contain only little guidance on the design and installation of bits, bollards and fairleads and, generally, they give no information on safe working loads, strength of materials, structural reinforcements, and foundations. This lack of guidance on the mooring outfit resulted in a wide variation in the number of mooring lines, type of mooring equipment, capacity of winches, etc. found across the range of similar vessels. Furthermore, many loading berths (e.g. tanker terminals) are situated at exposed locations, and existing mooring equipment on some vessels may not permit their maintaining a safe moor at these berths during all prevailing environmental conditions unless the berth operator can provide the additional restraint required. Safe mooring of a vessel requires close cooperation between ship and shore operators and between ship and terminal designers. To achieve this objective, all personnel involved must be familiar with the principles of good mooring practice, the magnitude of forces acting on the ship, the magnitude of loads in the mooring lines, and the limitations of the mooring

The vessel in port: mooring problems

453

equipment aboard ship and at the terminals. During 1978, in an attempt to achieve a uniform approach to the safe mooring of tankers, the Oil Companies International Marine Forum (OCIMF) compiled and published Guidelines and Recommendations for the Safe Mooring of Large Ships at Piers and Sea Islands.2 More recently, OCIMF set up a task group to investigate the concept of a total mooring system for tankers and to develop guidance on the installation of mooring equipment consistent with safe mooring practices. 3 The International Association of Independent Tanker Owners (INTERTANKO), the International Chamber of Shipping (ICS), and the IACS participated in the task group activity. This broadbased industry group developed mooring guidelines initially for tankers. In the ,development of the guidelines it was realized, however, that many of the principles adopted and the equipment strength and installation criteria proposed could also be applied to other kinds of vessel moorings. Certainly the general principles apply to safe mooring of any size vessel. An efficient mooring system is essential for the safety of the ship, the terminal and the environment. In dealing with the problem of optimizing the moorings with sufficient restraint capability, it is necessary to be concerned with environmental forces applied on the ship, general principles; that determine how the forces are distributed to the mooring lines, and application of these principles to establish a good mooring arrangement. The purpose of this paper is to discuss such items in view of problems concerning safe mooring of ships in port. Issues addressed include an overview of good mooring principles, guidance for berth layout and shipboard mooring equipment, and prediction of wind and current loads on the ship, but the concentration is on computer prediction techniques to assess effects of moored ship dynamics caused, for example, by the natural seaway. Computer predictions, resulting in time records of ship motions and mooring line forces or statistical measures thereof, are recommended when wave conditions cause significant dynamic loadings in mooring components.

2 REVIEW OF GOOD MOORING PRINCIPLES A mooring system must resist forces due to some, or possibly all, of the follIowing factors: wind and current, surges from passing vessels, and effects of waves, swells, and seiches as well as ice and tides. Normally, if the mooring arrangement is designed to accommodate maximum wind and current forces, reserve strength is sufficient to resist other forces which may arise. Considerable loads can be developed in the ship's moorings, however, if appreciable surge, wave conditions, or changes in ship eleva-

454

T. E. Schellin, C. Ostergaard

tion exist at the terminal. Surge is usually resisted by utilising additional spring lines, sometimes provided by the terminal. Wave-induced forces are minimized by providing a degree of elasticity in the moorings and by a judicious selection of the fender system. Forces in the moorings due to changes in ship elevation from either tidal fluctuations or vessel unloading operations must be compensated for by proper line tending.

2.1 Wind and current forces


The longitudinal wind force is relatively small since the head wind only strikes a small portion of the total exposed area of the ship. A beam wind, on the other hand, strikes the entire exposed side area of the ship, and a large transverse force is exerted on the vessel. For a given wind velocity the maximum transverse wind force on a very large crude carrier (VLCC), for example, is about five times as large as the maximum longitudinal wind force. With the exception of ahead or astern and broadside wind, the resultant wind force does not have the same angular direction as the wind. The point of application of the resultant force is either forward or aft of the transverse centre-line, thus producing a yawing moment on the vessel. Procedures for calculating wind forces and moments are generally based on wind force coefficients obtained from wind tunnel measurements. Wind force coefficients for VLCCs are given in an O C I M F publication; 4 for other kinds of ships, see publications by Wagner 5 and Blendermann, 6 for example. Current forces on the moored ship must be added to wind forces when evaluating a mooring arrangement. In general, the variation of current forces with current velocity and direction is similar to the variation of wind forces. Assessment of current forces, however, needs to consider the significant effect of water clearance beneath the keel. The majority of terminals are orientated essentially parallel to the current, thereby minimising current forces. Model tests indicate that the current force due to a 1 knot beam current on a loaded VLCC with a 2 m keel clearance is about 40 times as large as the current force on that ship with the same underkeel clearance due to a 1 knot head current. Procedures for calculating current forces and moments are generally based on current force coefficients obtained from model tank tests. Current force coefficients for VLCCs are given in Ref. 3; for other kinds of ships, see the literature dealing with the resistance of ships in calm water, for example, Ref. 7.

2.2 Mooring pattern


The effectiveness of a mooring line is influenced by two angles, namely, the vertical angle between line direction and pier deck and the horizontal

The vessel in port: mooring problems

455

angle between line direction and parallel side of the vessel. The steeper the orientation of a line, the less effective it is in resisting horizontal loads. For instance, a line orientated at a vertical angle of 45 is only 75% as effective in restraining the vessel against wind forces as a line orientated at a 20 vertical angle.

2.3 Elasticity of mooring lines


The elasticity of a mooring line depends on the material of manufacture, the line diameter, and the line length. It is a measure of its ability to stretch under load and to attain its original length after unloading. In general, if two lines of different elasticity are tied to the ship at the same point, the stiffer one will always sustain a greater portion of the load even if the orientations are identical. Short hawsers will generally carry more of the load than longer ones, and loading in large diameter hawsers will be grearer than loading in small diameter hawsers. Thus, mixed moorings consisting of lines of different materials used in the same direction of restraint should generally be avoided. The effect of line length on load distribution is often overlooked when evaluating the mooring arrangement. This is one reason why long bow and stern lines are inefficient when used in combination with shorter breast and spring lines.

2.4 IVlooring pattern arrangement


Assuming the moored vessel is exposed to strong wind or current from any direction, the following mooring pattern guidance should be observed to optimise load distribution on the moorings. (1) Mooring lines should be arranged as symmetrical as possible about the midship point of the vessel. (2) Breast lines should be orientated as perpendicular as possible to the longitudinal centre-line of the vessel and as far forward and aft as possible. (3) Spring lines should be orientated as parallel as possible to the longitudinal centre-line of the vessel. (4) Head and stern lines are generally not required for safe mooring provided shore mooring points are suitably designed and arranged. This means that a vessel can be moored most efficiently within its own length. (5) The vertical angle of mooring lines should be kept to a minimum. The flatter a mooring line is orientated, the more efficient it is in resisting horizontal loads acting on the vessel.

456

T. E. Schellin, C. Ostergaard

(6) Mooring lines of the same size and type (material) should generally be used for all leads. Mooring lines of the same size and type should always be used for all leads in the same service such as breast lines, springs, head lines, etc. (7) If tails are used on wire mooring lines, the same size and type of tail should be used on all lines run out in the same service. Synthetic tails are often used on the ends of wire mooring lines to permit easier handling and to increase line elasticity. The addition of a 10 m nylon tail, for instance, increases line elasticity of a 45 m long wire line fiveto six-fold. (8) Mooring lines should be arranged such that all lines in the same service are about the same length between the vessel's winch and the shore bollard. In practice, final selection of the mooring pattern for a given terminal must also consider local operational and weather conditions and pier geometry. For example, pilots may insist on head and stern lines to assist a vessel moving into, along, or out of a berth although many ships use their springs for this purpose. Head and stern lines, however, can be an advantage at berths where the mooring points are too close to the vessel and good breast lines cannot be provided or where the bollards are so located that the springs have an excessive vertical angle in the light ship condition.

3 G U I D A N C E F O R BERTH L A Y O U T The layout of a pier or sea island to permit the safe berthing and mooring of the whole range of vessel sizes is specified, in principle, by the spacing of breasting dolphins, the location of mooring dolphins, and the placement of spring line bollards. At a conventional pier for berthing other kinds of ships, breasting and mooring dolphins may be absent, with bollards or quick-release hooks, placed at corresponding locations, providing the shore mooring points.

3.1 Spacing of breasting dolphins


At conventional piers the main fenders are often part of the pier face. In new VLCC berth constructions, however, breasting dolphins are generally provided. Breasting dolphins should be set apart a distance approximately onethird the overall length of the vessel, symmetrically about the central

The vessel in port: mooring problems

457

loading station. Wider spacing of the breasting dolphins is desirable, but the range of vessel sizes must be considered. Spacing limits are generally set between a minimum of 0-25 and a maximum of 0.4 times the overall length to ensure contact with the vessel's parallel sides. Breasting dolphins may also serve as spring line mooring points if they are located such that spring lines have sufficient length.

3.2 Location of mooring dolphins


Mooring dolphins are located to handle vessel head, stern and breast lines. For an ideal berth, four mooring dolphins are normally provided. However, where a wide range of vessels must be handled or where operations require head and stern line dolphins for first-line-ashore, six dolphins may be required. As discussed above, a vessel can be safely moored within its own length using only breast and spring lines. Head and stern lines are inefficient in providing restraint capacity. However, a mooring point that provides a good breast line lead for a large ship becomes a head or stern line mooring point for a smaller Ship. The two outermost mooring dolphins should be spaced slightly further apart than the longest vessel to be handled to avoid interference between mooring lines and hull. The immediate mooring dolphins are then located to optimise breast line orientation and to accommodate the range of vessel sizes to be handled.

3.3 Placement of spring line mooring points


Normally, four mooring points for vessel spring lines should be provided on the loading platform or on the breasting dolphins. They should be placed as close to the berthing face as possible to keep the spring lines as parallel as possible to the longitudinal axis of the vessel. As with breast lines, the vertical angle of spring lines should preferably not exceed 25 during loading or unloading, and line lengths should be comparable. Where a wide range of vessel sizes must be accommodated at berth, spring line interference may occur at the breasting dolphins from small vessels in loaded conditions. To diminish this problem, breasting dolphin elevations must be kept low and spring line mooring point positions carefully selected. Generally, two spring line mooring points forward and two aft are required, and they may be constructed as double hooks. However, special consideration must be given to smaller ships using the berth.

458

T. E. Schellin, C. Ostergaard

4 G U I D A N C E FOR SHIPBOARD M O O R I N G E Q U I P M E N T LAYOUT

The layout of shipboard mooring equipment concerns placement onboard ship of items such as winches, fairleads, chocks, pedestal rollers, etc. To accomplish the goals of good mooring principles, the following guidance should be considered. (1) The mooring arrangements should be symmetrical, and all mooring lines should be capable of being run to either side of the vessel. (2) Mooring operations should be located as far forward and aft as possible. Bow and stern fairleads should be located as far forward and aft and as low as possible on the ship. Spring line fairleads should be placed as far forward and aft on the main deck as possible to provide adequate spring line lengths. Breast lines should be as far forward and aft as possible to be most effective in restraining the yawing tendency of the moored ship. (3) Winches should be located so that the length of all breast lines from the winch drums to the shore mooring points are equal. Since clear deck space is limited, this may not always be realised. Nevertheless, winches and chocks should be located to make the total length as nearly equal as possible. (4) Mooring areas should be kept as clear as possible, and winch control positions should be located to provide a clear view of the mooring operations for the officer-in-charge of mooring. The correct alignment between fairlead and winch drum should be observed, and mooring lines in the same service should have about the same length between winch and chocks and between chocks and shore mooring points. (5) Consideration should be given to providing additional bits and fairleads onboard to permit use of shore moorings or shore augmentation of ship moorings.

5 MOORING EQUIPMENT In recent years several mooring incidents involving crude-oil carriers occurred resulting in damage to the terminal's loading arms and pollution of harbour waters. The usual causes of these incidents were broken mooring lines or tails, winch slippage, and excessive vessel movements permitted by the high elasticity of synthetic lines. Selection of and operating criteria for ship as well as shore mooring equipment comprising

The vessel in port: mooring problems

459

mooring lines, winches (including winch brakes), chocks, and pedestal rollers should be based on the characteristics of these components.

5.1 'Wire mooring lines


Wire mooring lines are generally recommended for VLCCs because they prevent excessive ship movement which is the usual cause of damage to shore-based loading arms. Another reason is that when mixed moorings are used, the wire lines take most of the load, and early line failure can occur. The excessive elongation of synthetic fibre lines even at relatively low load levels is illustrated in Fig. l, showing curves of percent breaking load versus percent extension of length for wire lines and used synthetic fibre', ropes. To keep line-handling ability within manageable proportions for the larger sizes, mooring lines should be selected that have the highest minim u m breaking load (MBL) compatible with ease of handling. The general consensus of ship operators is that a 42 m m line meets these criteria although there are some who operate VLCCs with 44 m m wire mooring lines'.. Steel wires with an independent wire rope core (IWRC) are recommended over fibre core steel wire lines because of their higher resistance to crushing, higher MBL for a given diameter, and greater strength retention when bent. The harmful effects of bending on both fibre core and IWRC wire lines are shown in Fig. 2.
xoo / / , i / / l /

-//
60 /

/ /
/

USEDDOUBLE ONBRA,D J
POLYESTER ROPE /

/ / /

,"

/"

r /j

6x361WRC WIRE LINE /

/ / / / ~/"

o,

i 8 ST..NO,',..r,EO R O , ' E > . - " / . . "


H I .6x36FIBERCORE

I USEDPOLYPROPYLENE

/ ~USEDNYLON / - - - - " - ' - 7 8 STRAND ,/ ,,,' PLAITED ROPE / " / / USED DOUBLE BRAID

/ - - NYLO.ROPE

II/~ ~RE',NE
20

~.~.~~~...
o

III Ill III

/5I.

" /

/..-" /
/ 5
I

//

//r

//

/.~

/'"

i
I

CURVESWILLVARYFROM IMANUFACTURER TO MANUFACTURER 15 20

N O T E :

10 % EXTENSIONOF LENGTH

Fig.

I.

Typical curves for % breaking load versus % extension of length for various materials.2

460

T. E. Schellin, C. Ostergaard

'

,0

-CORED W,BE RO E

751~ 8

I 10

I 12

I 14

I 16

I 18

I 20

I 22

24

BENDING RATIO D/d WHERE: D = DIAMETER OF BITT, BOLLARD, FAIRLEAD, ETC. d = DIAMETER OF WIRE ROPE

Fig. 2. Effects of bending on wire rope strength.2


To meet the requirements of increased strength for wire mooring lines, manufacturers developed preformed, drawn-galvanised wire with high tensile strengths. For mooring VLCCs, it is recommended that, as a minimum, a 42 m m diameter 6 37 class IWRC preformed, heavily drawn galvanised wire line (minimum tensile strength of 1.77 k N / m m 2) with a typical MBL of 1130 kN be specified. A wire line in service for a short time experiences a permanent lengthening, known as constructional stretch, due to a slight rearrangement of the wires. Thereafter, a wire line undergoes a recoverable, linear elastic stretch for the load range from approximately 20 to 65% MBL of the wire line when new. For six-strand IWRC construction, the extension varies between 0.25 and 0-5%; for a fibre core wire, between about 0.5 and 0.75%. For loads above 65% MBL, the elastic limit is exceeded and the stretch increases nonlinearly until the line breaks (Fig. 1). For design purposes wire mooring lines should not be loaded in excess of 55% MBL for normal maximum use. This level of loading allows for reduced strength due to eye splices and minor degradation of the wire and ensures loads remain below the yield point of the wire. Mooring lines on large ships should have a minimum length of 275 m to allow for berthing at multi-buoy moorings as well as piers and sea islands. Splices other than eye splices are not recommended.

5.2 Synthetic tails


Wire mooring lines of some large ships are fitted with a length of synthetic rope on the shore end, a so-called synthetic tail. At some terminals, tails

The vessel in port: mooring problems

461

are provided when ships fail to have them. Whether use of these tails is beneficial has not been established. Model tests, field measurements, and experience indicate the general desirability of the additional elasticity provided by tails. This additional elasticity reduces the loads induced in wire lines under dynamic loadings by allowing the ship to respond more favourably to the action of wind, current and waves as well as to ships passing in close proximity. However, improperly designed tails may introduce too much elasticity, allowing the vessel to move more than can be tolerated by the terminal's cargo transfer syste,m. In mooring lines in the same service, tails tend to distribute the loadings more evenly. The additional elongation of the mooring line system permitted by tails reduces the risk associated with poor line tending and the frequency and precision of line tending, particularly at berths with large tidal variations and high loading/unloading rates. Tails are also valuable at berths having short breast or spring leads as they provide the same effect as longer, all wire lines. Tails facilitate the handling rate of wire lines by boatmen and mooring gangs. However, tails may introduce a weak link into the moorings, and this may not be readily apparent to ship operators. Failures of tails have occurred in many ship moorings, indicating that tails have a lower minimum breaking strength than the wire lines to which they are attached, that they have a poor connection to the wire line, or that they are not replaced at regular intervals. The synthetic rope of tails should have a breaking strength at least 25% greater than that of the wire line to which it is attached. This ensures that the loading as a percentage of the breaking strength of the tail is lower than that in the wire mooring line. Reducing the maximum percent loading increases the useful life of the tail. Experience indicates that synthetic lines; degrade more quickly than wire lines under similar load conditions. The size of rope of the tail should be capable of easy handling and be made of a material with high breaking strength, such as braided or plaited nylon. The melting point of the material grade should also be considered. A higher melting point is generally more desirable, thereby reducing the possibility of damage due to internal heat generated at high cyclic loadings. Tails should be connected to the wire mooring lines in a manner that prevents abrasion. A wire line/synthetic tail shackle should be used, and the eyes of the tail should be served with leather or plastic shoaling to protect them from chafe at their ends. Tails should be inspected before use and those showing signs of damage removed from service. They should be replaced at least every 18 months unless experience and/or inspection indicates otherwise. General informa-

462

T. E. Schellin, C. Ostergaard

tion on inspection and replacement of a line is found in an Appendix to Ref. 2. Specific information is available from line vendors.

5.3 Mooring winches


The heart of the mooring system is the deck-mounted reel-stowage mooring winch, including winch drums and brakes. For the mooring system to operate properly, an adequate number of winches must be installed, a sufficient number of drums must be provided to cater to the mooring wire requirements, and the winch brakes must be designed, operated and maintained to furnish the designed holding load capacity. Mooring winches aboard ships are either manually operated or they are automatic tension winches. Automatic tension winches are designed for a specified line tension that can be preset so that the winch pays-out mooring wire whenever this preset value is exceeded by a set amount. The winch hauls in whenever the line tension falls to a certain amount below the preset value. If automatic tension winches are installed, they should be put on manual brakes while the vessel is moored and its automatic features are not used. The reason is that forward/aft springs and head/stern lines can work against one another when the ship is exposed to wind and current forces. This decreases their effectiveness in counteracting these forces and may cause the ship to move along the pier, possibly exceeding the operating tolerances of the cargo transfer devices, resulting in damage and possibly pollution. Manually operated winches require that mooring wires be hauled-in or slacked and the brake set either mechanically or hydraulically to a predetermined torque or pressure setting, respectively, to ensure design holding capacity. The hauling load capacity of a winch with one layer on the drum should not exceed 33% of the mooring wire's MBL. The preferred no-load speed of the first layer of the drum is 1.5 m/s; however, this speed should never be less than 0.5 m/s. At maximum hauling load, design speed should be 0-5m/s, and this design speed should never be less than 0-12 m/s. It is also recommended that winch engines have a slippage to hauling-in ratio sufficiently low to prevent yielding of wire lines when an overload condition occurs. Winch brakes should be set to hold a minimum load of 60% MLB of the wire on the appropriate wire layer on the drum. If set for a higher level, slippage of the wire should be permitted before breaking. The 60% minimum is selected because evidence available indicates that IWRC wire lines yield at 65-75% MBL when new. Considering friction in fairleads, it is believed that the 65% level is readily obtained in the outboard portion of the mooring lines. Further evidence indicates that a line in service for a

The vessel in port: mooring problems

463

few '.years experiences a reduction in braking load due to cyclic loadings, abrasion, etc. Thus, when considering further reductions due to splices and bending, it seems that a brake holding load of 60--65% MBL when new is consistent with the objective of protecting the line from breakage. Tile physical condition of the winch gearing and the brakeshoe linings have a significant effect on brake holding load capacity. It is recommended that all winch brakes be tested after installation and periodically tested and inspected thereafter. For mechanically actuated brakes, test results should be used to determine torque levels to obtain design brake holding capacity. Oil, moisture, or heavy rust on the brake linings or brake drum can reduce brake holding capacity up to 75%. Many operators run the winch with the brake set slightly to burn off or wear off the oil or moisture. Care must be taken to prevent excess wear due to the buildup of heat when using composite brake linings.

5.4 Fairleads
Since the braking strength of wire is affected by bending, all roller fairleads should have a minimum radius which is 10 times the radius of the wire.. With Panama-type fairleads, wire is subjected to greater friction forces, and thus a minimum bending ratio of 12:1 is suggested. Fairleads should be chafe-free from any direction either inboard or outboard.

5.5 Pedestal rollers


To minimise friction, pedestal rollers are at times installed because it is not always possible to have direct leads from the winch drum to the fairlead. As for a roller fairlead, a minimum bending ratio of 10:l is prescribed for the pedestal rollers. The directional change of a wire lead should be as small as possible and not exceed 90 .

5.5 Shore-based mooring equipment


Shore-based mooring equipment must be compatible with ship moorings, located properly to utilise ship moorings, and of sufficient capacity to ensure safe mooring and efficient handling of lines. This equipment should be designed utilising appropriate safety factors and fabricated and installed to withstand the minimum design loads. Fixed bollards should be provided at appropriate locations to serve as shore mooring points. Bollards are not recommended, however, at berths for large ships such as VLCCs. At these berths quick-release hooks should be installed to allow remote release of mooring lines for expediting emer-

464

T. E. Schellin, C. Ostergaard

gency sailing with minimum available manpower. Each quick-release hook, whether a single hook or part of a multiple hook unit, should have a safe working load not less than the MBL of the largest line anticipated. Only one line should be placed on each quick-release hook. A sufficient number of hooks must be provided. All hooks should be capable of being released separately and safely from the mooring point area under full to no load conditions. At berths for large ships such as VLCCs, capstans having vertical spindles should be located at each mooring point for pulling mooring lines ashore. They should be built into the quick-release mooring device or be located adjacent to them to enable the eye of the mooring line to be slipped over the mooring hook. Capstans should be capable of holding the load with the motor stopped. Instead of capstans, winches with horizontal drums are frequently preferred by some operators. They have the same pulling and load characteristics as capstans and are generally preferred for use with gallows that permit the mooring line to be placed over the quick-release hook more readily.

6 M O O R I N G ANALYSIS At most berths the mooring system designed to accommodate maximum wind and current forces is generally adequate to also resist forces arising from other sources such as waves. For such situations, calculations to determine the number of mooring lines to install on a given vessel can be accomplished with static analysis methods based on specified design mooring conditions selected to meet conditions that may be encountered at the terminal. For conditions that may be encountered at most world wide VLCC terminals, for instance, design mooring conditions were compiled by OCIMF. 2

6.1 Low-frequency ship motions


Large tankers, bulk carriers and LNG carriers tend to be moored to cargo-handling jetties built close to deep water to reduce the costs of dredging navigation channels. As a result, mooring locations for these ships are more exposed to wave action, and under certain conditions a moored ship experiences large horizontal motions, causing breakage of mooring lines. 9-1~ The problem is generally due to low-frequency excitations, which may be caused by slowly-varying wave drift forces associated with the development of wave groups as experienced, for example, during

The vessel in port: mooring problems

465

swell conditions; but they may also be caused by harbour basin longo period standing waves (seiches) and by wind speed variations. 12' 13 A significant feature of a moored ship system is that damping at natural periods is relatively small. Therefore, although low-frequency excitations are also small as they are generally of second-order, large amplitude resonant horizontal oscillations may occur because the periods of lowfrequency excitations may fall within the range of the natural periods of the raoored ship. 6.2 Prediction of moored ship behaviour In the past, the only way to obtain accurate estimates of ship motions and mooring loads was to perform expensive scale model tests. During the last decades, numerical methods became feasible, brought about by powerful computers and progress in the development of mathematical models. 14~2'46 The predictions made using these mathematical models generally rely on the following simplifying assumptions. (1) First-order ship responses at wave frequencies are modelled as linear responses to harmonic waves using response functions due to waves of unit amplitude (transfer functions) for the six degree of freedom motions of the ship in waves. Effects of the (linearised) stiffness of the mooring system can be accounted for. (2) Low-frequency ship motions in surge, sway and yaw are not affected by first-order ship motions. They are modelled as responses to wind, current and wave drift forces acting on the moored ship. (3) To obtain total response, first-order (high-frequency) response is combined with low-frequency response using an appropriate method. (4) Dynamic behaviour of mooring lines and fenders does not significantly affect motions of the moored ship. Consequently, mooring forces are determined from the instantaneous position of the fairleads and from load~teflection characteristics of mooting lines and fenders.
6.2.1 Forces on the ship

First-order hydrodynamic forces (and the resulting high-frequency motJions) of the moored ship in regular waves can be calculated by means of a singularity method based on three-dimensional potential flow. The incident waves are assumed to undergo scattering (diffraction) at the ship, leading to a diffraction wave potential that is superimposed on the incident wave potential and the radiated wave potentials caused by the

466

T. E. Schellin, C. Ostergaard

motions of the oscillating ship. The wetted surface of the ship is subdivided into surface elements (Fig. 3), and the singularities are distributed on these elements. Hydrodyamic forces are obtained by integrating pressures due to the incident, radiated, and diffracted wave potentials. 22 25 Computations using unit amplitude waves of different frequencies yield transfer functions. Figure 4 shows a sample transfer function for surge motion in head waves of the tanker idealized in Fig. 3. The wave drift forces have small magnitudes proportional to the square of the wave height. In general, there are two methods to solve for wave drift forces in regular waves, namely, the far-field and near-field approaches. The far-field approach obtains drift forces from consideration of conservation of total energy or total momentum of the fluid. 26'27 This method requires knowing the total velocity potential and its derivatives far away from the ship. The near-field method, 28 in contrast, calculates drift forces from direct integration of the pressure on the ship's wetted surface. Both methods have advantages, but the near-field approach is often preferred because it allows for a physical interpretation of the components of drift forces. The near-field method employs regular perturbation expansions to formulate a set of boundary value problems to solve for the velocity potentials. With this method only first-order potentials and their derivatives are required to determine drift forces. Neglecting second-order motion responses in time except for the steady drift components, mean drift forces are obtained by integrating over one wave period and taking the time average. Based on the near-field approach, 29' 30 the non-dimensional drift forces in regular head waves for the tanker idealised in Fig. 3 were calculated (Fig. 5). The magnitude of low-frequency damping to a large extent determines the (resonant) low-frequency motions of the moored ship. Since potential damping calculated using linear theory is small at low frequencies in the

Fig. 3. Surface element idealisation o f a tanker. 2t

The vessel in port." mooring problems


2~
\ ~ ESl ~S1 T~
= PHASE ANGLE

467

\
o \

to_ I

_ .._

1.2
----- s1 = COMPUTED MEASURED SURGE AMPL. WAVE AMPL.

1.0

Is~L
~

0.8

0.6

/ / / /

0./,

/.

0.2

/
i

L - - 4 4 " O ~1 ~ * " \ V /

10

15
WAVE PERIOD [sl

20

Fig, 4, Surge m o t i o n transfer function in head seas for a tanker. 21

neighbourhood of the natural frequencies of the moored ship, other sources of damping such as wave drift damping and viscous damping become important. The low-frequency damping can be subdivided into two components, namely, still water damping and wave drift damping. Stillwater damping can be estimated from extinction tests in calm water; wave drift damping, from extinction tests in waves by registering the increase in damping as a function of incident wave frequency. 3~ Using potential theory based analysis techniques, wave drift damping has also been obtained by the so-called added resistance gradient technique 32 or the approximate drift force gradient approach. 33 Damping due to viscous pressure drag or eddy making is important for sway and yaw motions and may be estimated from cross-flow drag coefficients of the ship. Forces in mooring lines are a function of their respective terminal point coordinates onshore and their fairlead coordinates aboard ship. Restoring forces for each line are determined from force elongation characteristics and catenary effects.

468
-2.0

T. E. Schellin, C. Ostergaard
I I I

~_(1) = -1.5 T=(1) ~'Pg


1 V1/3 [2

COMPUTED MEASURED MEAN DRIFTFORCE DENSITY OF WATER ACC. OF GRAVITY VESSEL DISPLACEMENT WAVE AMPLITUDE

p g V 4a

= = = =

-a -1,0

-0.5

0 0 5 10 15 WAVE PERIOD [s] 20

Fig. 5. Mean longitudinal drift force in head seas for a tanker.

Fender reactions comprise friction forces alongside the ship as well as compression and relief forces perpendicular to the pier wall for a simplified modelling of fender hysteresis. Wind and current forces from the largest part of the steady forces that have to be restrained by the moorings. Wind and current forces are caused by viscous effects and may be estimated using empirically derived force coefficients (see Section 2.1). To account for wind gusts during simulations, the correct level of turbulence is superimposed on mean wind speed. Wind and current may also contribute to damping forces of low-frequency motions. This effect can be accounted for by considering relative velocities between ship and air or water in the aerodynamic and hydrodynamic force formulations, respectively. Recent experience 34 demonstrated the effective use of a four-quadrant manoeuvring model 35'36 to determine the quasisteady hydrodynamic response and control forces on an anchored tanker subject to current, wind and waves. Passing ships may induce large peak loads in the mooring lines of the moored ship. Many instances are known where lines broke as a result of a ship passing close by at high speed. Loads on the moored ship caused by a passing ship are most reliably obtained from scale model tests. 37'38 They were found to be proportional to the square of the speed of the passing ship. Figure 6 shows sample results of such measurements, here for the lateral force on a captive 100000 D W T tanker due to the passage of a 110 000 D W T tanker.

The vessel in port." mooring problems


20
I
PASSING SHIP MOORED SHIP

469

I
110 MDWT 100 MDWT PASSING DISTANCE

60 20

t1

I
--

Fy v
x

= =
=

LATERAL FORCE ON CAPTIVE SHIP SPEED OF PASSING SHIP LONGITUDINAL DISTANCE BETWEEN THE MIDSHIP POINTS OF THE VESSELS LENGTH BETWEEN PERPENDICULARS OF PASSING SHIP LENGTH BETWEEN PERPENDICULARS OF CAPTIVE SHIP
=

m "10 L1 L:Z -i x -= =

(L, + L 2) _x . (L1
L~J

-1 CORRESPONDS TO THE BOW-TO-BOW POSITION +1 CORRESPONDS TO THE STERN-TO-STERN POSITION

B20

I
-2 0

I
2

LONGITUDINAL POSITION OF PASSING SHIP

(L~+ L~

Fig. 6. Lateral force on a captive 100000 DWT tanker due to the passage of a 110000 DWT tanker,as

6.2.2 Surge motion predictions of a moored tanker


It appears that the mooring pattern of large ships at exposed jetties is always rather similar, using an average o f about 16 mooring lines of vari~ous materials and thicknesses. The natural period o f vessel surge with this kind of mooring system usually ranges between 50 and 150 s. During storm conditions, wave groups having periods within this range frequently occur at exposed jetty locations. From a designer's standpoint, to assure the safety o f ships and jetty structures and to reduce downtime when loading the ships, it may be useful to know the expected maximum surge motion o f a moored ship. We calculated maximum surge excursions of the 240 000t tanker, idealised in Fig. 3, moored to an exposed jetty and subject to the action of long-crested natural seaways with waves propagating in the longitudinal

470

T. E. Schellin, C. Ostergaard

ship direction. We approximated wave conditions according to the Pierson-Moscowitz spectrum ~ and specified the seastates by the significant wave height Hs, defined as the average of the one-third highest wave heights, and the average zero-upcrossing period T~, defined in terms of the zeroth and second spectral moments m0 and m2 as Tz ----2It(too~m2) 1/2. We considered seastates having significant wave heights of up to 5 m and selected the matching zero-upcrossing periods, ranging up to 11.6 s, that corresponded to swell wave conditions as recommended by the American Petroleum Institute. 45 We assumed the mooring system's overall restoring force characteristic in surge to be linear and chose five representative restoring force constants of 500, 700, 1000, 2000 and 4000 kN/m that correspond to natural surge periods of the moored ship of 148, 125, 105, 74 and 52 s, respectively. By considering only surge response and assuming that low-frequency surge can be decoupled from all other ship motions, the low-frequency surge equation can be written in linear form with constant coefficients. This enabled us to perform a frequency domain analysis using spectral expressions of wave drift force. We assumed that low-frequency surge response, XLF, is the solution of the motion equation
(m + a)JCLF + b-~LF + CfXLF = fLF

(1)

where m and a represent, respectively, the tanker's mass and its hydrodynamic added mass in surge at low-frequency, b is the low-frequency damping coefficient in surge, C/ is the mooring system restoring force constant, and fLF is the low-frequency wave (drift) force for the tanker in head waves. We assumed a constant added mass in surge equal to 15% of ship mass and a damping coefficient equal to 1.2% of critical damping. The critical damping, be, being a function of the mooring restoring force coefficient, was obtained from bc = 2[(m + a)Cr] t/z (2)

The natural surge period, Tn, was of course also a function of the mooring restoring force coefficients as given by T, = 2rc[(m + a)/Cf] j/2
from

(3)

The response spectrum of low-frequency surge, SxLF (P), was determined S~LF (P) = SrLF (#) X DRO 2(/2) (4)

where SxLF (#) is the spectral density of the low-frequency wave drift force, DRO (p) is the amplitude of the low-frequency surge motion response function, and p is the circular wave frequency at low frequencies.

The vessel in port: mooring problems

471

Using the previously calculated mean drift forces in regular waves as given in Fig. 5, the spectral density of the drift force was calculated from

SfLF (~) : 2p2g2V2/3 0 SC(co)S(o~+ U) L1/2p g V1/3~J dw

J~

[F"(w + #/2)] 2

(5)

where .~l(w) is the mean drift force in regular waves of circular frequency co; p is the density of water; g is the acceleration of gravity; V is the ship displacement; G is the wave amplitude; Sc(o9) is the spectral density of the seaway; and

DRO(#) = 1/~//[Cf- (m + a)kt2] 2 + b2/fl

(6)

The standard deviation of the low-frequency surge amplitude, axLF, was calculated by integrating its spectrum and taking the square root
O'xLF = ~ / l O SxLF(J~) d/-t

(7)

and its maximum value, XLFmax, based on a 2 h storm duration, was obtained from
XLFmax = 3.72 O'xLF (8)

The mean surge excursion, X, was calculated by dividing the mean lowfrequency wave (drift) force, fLF, by the mooring system's restoring force coefficient, Cr, that is

=fLF/C/
wherefL F was obtained from

(9)

L~ = p ~ V "

,o~

S~(~,) 1/2p---~-~/~j~

pl(~)

].

(lO)

The tanker's high-frequency surge response was determined from standard spectral techniques for a linear system. The response spectrum of highfrequency surge, Sx.F(og), was obtained from the surge motion transfer function as given in Fig. 4, that is S~HF(W) = [Sl(W)/~a]2 S(w) (11)

The standard deviation of high-frequency surge amplitude, trxHF, was calculated by integrating its spectrum and taking the square root, and the

472

T. E. Schellin, C. Ostergaard

maximum high-frequency surge amplitude, XHFmax, based on a 2 h storm duration, was determined from
XHFmax z

3.72 trxHF

(12)

The expected m a x i m u m surge of the moored tanker, Xmax, comprising the low-frequency surge amplitude, the mean surge excursion, and the highfrequency surge amplitude, was obtained by adding to the mean surge a combined value of the low-frequency and high-frequency surge. This combined value was the sum of the maximum value of the dominant component and the significant value of the other component, where the dominant component is that with the highest value, i.e. ifXLFmax > XHFmax, then Xmax = ~ + XLFmax+ XUFsig ifXLFmax < XnFmax, then Xma x 2 -~- XLFsig -~- X n F m a x (13)

Here subscript sig stands for the significant value, which is equal to twice the standard deviation. This method follows AP145 recommended practice for computing maximum mooring line tensions. The resulting expected maximum surge motions of the moored tanker are plotted in Fig. 7 as functions of the significant wave height. They depend on the mooring system restoring force coefficient. It is seen that
8 , , ,

6
Xmax

Cf = 500 700 1000 2000 4000


Hs [m] 1 2 3 Tz [s]

kN/m kN/m kN/m kN/m kN/m

[m]
4

5.0 7.3 9.0

H s [m]

Fig. 7. Maximum expected surge excursions of a 240000 tanker moored to an exposed jetty.

The vesselin port: mooringproblems

473

surge motions can be relatively large when significant wave heights exceed 2-3 in, which is in agreement with reported experience. 9 For low restoring force,' coefficients (Cf = 500-1000 kN/m), the major part of tanker surge (over 70%) is due to the low-frequency motions; for high restoring force coefficients (Cf= 2000 and 4000 kN/m), the low-frequency motions contribute about 50% of the total surge motion.

6.2.3' Simulation method


Sometimes it is necessary to account for system non-linearities to obtain reliable predictions. Non-linearities may be due to, for example, increasing force-elongation characteristics of the moorings, frictional effects of the fenders, or viscous effects of hydrodynamic forces. Non-linear simulations are then suitable to numerically predict the moored ship's response in the time domain including the forces in the mooring lines. Such a simulation method generally comprises the following steps. (1) No-load static equilibrium. The position of the moored ship is calculated with no environmental forces acting. Mooring line pretensions are specified as inputs. (2) Mean-load static equilibrium. The position of the moored ship is calculated under influence of mean environmental loads. Mooring line tensions and fender forces are calculated from force-elongation and force<leflection characteristics, respectively. (3) High-frequency motions. By means of computed transfer functions, the six degrees of freedom high-frequency ship motions in waves are calculated in the time domain by superimposing responses to individual wave components as specified by the wave spectrum of the natural seaway. Effects of mooring system stiffness are not directly accounted for, only indirectly when computing the motion transfer functions. (4) Low-frequency motions. A set of three coupled non-linear secondorder differential equations describing surge, sway and yaw motions of the ship as affected by low-frequency wind and wave drift forces, current forces, and mooring system reactions are integrated at small successive time steps to obtain time series of low-frequency ship motions. The linear hydrodynamic frequency response to the ship's own motion can be approximated in the time domain by means of a finite state space model, 39"40 where recursive state vectors effectively store motion history during time domain simulation (memory effect). (5) Total ship motions. The sum of high- and low-frequency ship

4~

2.00 t " ..... MEASURED COMPUTED

WAVE [m]

L i , l h . . I b , . _ . k .,. _J, lllm. . . . . . . . l l l , . . , = . , . ,

SURGE [m]
_ , -,.,.=~ , A . J,n=, =,,,=_

SWAY [m]
--"+" ~ ""'~'-

1 . o ;n _ r . -

.'--I

-1.ooL 0.50

YAW _ ~
. " Y ' - " ' ~ , . ---

I ~ , +

WIND

0.50

M LINE 5 [t] :
I
I

100.00 ~ , , 0L -100.00 I
I

FENDER 1 [t]
I I

lr 0 200 400
TIME Is] 600

-,oo.oo L

WAVES

Fig. 8.

Comparison between computed and measured time series of system response, bow quartering seas and beam wind. 43

The vessel in port." mooringproblems

475

motions yields total ship motions, and corresponding mooring tensions and fender forces are computed on the basis of the ship's position relative to the terminal. 6.3 Comparison with measurements Data on full-scale measurements are scarce. Some field observations were recorded from eight ships at various berths in three Icelandic harbours. 4 Unfortunately, comparative computer simulations of ship motions and mooring forces were not performed. The Maritime Research Institute Netherlands (MARIN), however, carried out an extensive validation study 42 for a large tanker moored to an offshore jetty that also included model test measurements in their wave and current basin (60 m long, 40 m wide, 1 m deep). The random wave signal generated in the model basin was used as input for their computer simulations, allowing a direct comparison of measured and computed records of ship motions and mooring loads. An example of their results is reproduced in Fig. 9. 44 The results generally confirm the validity of their simulation computer model. Computed time histories of ship motions and mooring forces can generally be Fourier-analysed to obtain energy spectra of system response and statistical measures such as mean and RMS values, significant amplitudes, and maxima or minima. Validation studies at M A R I N 44 also presented response spectra. Computed and measured spectra of surge
SPECTRA OF SURGE MOTION - - - ..... MEASURED COMPUTED SPECTRA OF MOORING LINE FORCE - MEASURED COMPUTED
. . . . .

SPECTRA OF FENDER LOAD MEASURED ..... COMPUTED

m=

1__

m i,

I
it f

0.5 FREQUENCY [rad

1.0

0.5

1.0

0.5

1.0

S "1]

FREQUENCY [rad s "1]


BOW QUARTERING WAVES

FREQUENCY [rod s -~]

Fig. 9. Comparison between computed and measured response spectra. 44

476

T. E. Schellin, C. Ostergaard

motion o f the moored ship, tension in one of the mooring lines, and load of one of the fenders are reproduced in Fig. 9. Their close agreement validates the simulation model. From a practical standpoint, response spectra are generally more useful than response time functions. The spectra in Fig. 9, for example, show that horizontal motions and mooring loads occur mainly at low frequencies in the neighbourhood of the natural frequencies of the moored ship, whereas responses at wave frequencies are comparatively small.

7 CONCLUDING REMARKS Adhering to the guidance of good mooring practice should generally endure safe mooring of ships. There are situations, however, where such guidance may be inadequate. This is generally the case when it is difficult to stipulate with sufficient accuracy external loads acting on the moored vessel, for example, when mooring large ships at exposed terminals subject to the occurrence of relatively severe seastates or when, in a harbour, a large ship passes a berthed vessel at close range. In such situations it may be helpful to resort to computer predictions, possibly in combination with scale model tests for validation. Although the understanding of the behaviour of moored ships, the ability to analyse the ship's motions, and the prediction of mooring loads has improved considerably in recent years, there are still uncertainties that have to be dealt with because even the most advanced analysis methods are simplified models of reality. These models are valuable tools to make certain that most relevant aspects are accounted for in the analytical computation, but continued comparison with full-scale observations and measurements should be endorsed.

REFERENCES 1. Germanischer Lloyd, Rules for the Classification and Construction of Seagoing Steel Ships. Hamburg, Germany, 1981. 2. OCIMF, Guidelines and Recommendations for the Safe Mooring of Large Ships at Piers and Sea Islands. Witherby & Co. Ltd, London, UK, 1978. 3. OCIMF, Mooring Equipment Guidelines (Task force Report, Final Draft). June, 1990. 4. OCIMF, Prediction of Wind and Current Loads on VLCCs, Witherby & Co. Ltd., London, UK, 1977. 5. Wagner, B., Wind forces on surface ships. Trans. Schiffbautechnische Gesellschaft, 61 (1967) (in German). 6. Blendermann, W., The Wind Forces on Ships (Report Nr. 467). Institut f~ir Schiffbau, University of Hamburg, Germany, 1986 (in German).

The vessel in port: mooring problems

477

7. Lewis, E. V., Principles of Naval Architecture (Vol. II, 2nd edn). SNAME, Jersey City, 1988. 8. I-latplapa Maschinenfabrik GmbH, Brochure on Mooring Winches, Uetersen, Germany. 9. Moes, J. & Holroyd, S. G., Measurement techniques for moored ship dynamics. In 18th Int. Conf. on Coastal Engineering, Cape Town, 1982. 10. Bowers, E. C., Long Period Oscillations of Moored Ships Subjected to Short Wave Seas (Transactions, Paper W4). The Royal Institution of Naval Architects, 1975. I 1. Nagai, S., Oda, K. & Lee, T. T., Impact forces, mooring forces and motions~ of super-tankers at offshore terminals subjected to wave actions. In Proc. 24th Congress of PIANC, Leningrad, 1977. 12. Wilson, B. W., The mechanism of seiches in Table Bay Harbor, Cape Town. In Proc. 4th Conference on Coastal Engineering, Chicago, 1954. 13. Keith, J. M. & Murphy, E. J., Harbor study for San Nicolas Bay, Peru. J. Waterways and Harbors Division, ASCE, 96 (WW2) (1970). 14. Van Oortmerssen, G., The motions of a moored ship in waves. Dissertation, MARIN Publication 510, Delft, The Netherlands, 1976. 15. Seidi, L. H. & Ishi, B. T., A system for the analysis of the dynamics of vessels and platforms moored offshore. In Proc. 3rd Int. Syrup. on Offshore Eng., COPPE, Rio de Janeiro, Brazil, 1981. 16. Roberts, J. B., Nonlinear analysis of slow drift oscillations of moored vessels in random seas. J. Ship Research, 25(2) (1981). 17. Schellin, T. E., Scharrer, M. & Matthies, H. G., Analysis of vessels moored in shallow, unprotected waters. In Proc. Offshore Technology Conference (OTC 4.243). Houston, 1982. 18. Fylling, I. J. & Andersson, F., Simulation of motions of berthed vessels-simplified simulation model. In Proc. Advances in Berthing and Mooring of Ships and Offshore Structures, (NATO ASI Series, Series E: Applied Sciences Vol. 146). Kluwer Academic Publ., Dordrecht, The Netherlands, 1988, pp. 298-303. 19. Germanischer Lloyd, Annual Report 1979. Hamburg, Germany, 1979. 20. Schellin, T. E., Calculation of line loads of moored objects. Handbuch ffir Hafenbau and Umschlagtechnik (Vol. XXXI). Schiffahrts-Verlag, Hansa, 1986, pp. 163-71 (in German). 21. ()stergaard, C. & Schellin, T. E., Comparison of experimental and theoretical wave actions on floating and compliant offshore structures. Appl. Ocean Res., 9(4) (1987) 192-213. 22. Clauss, G., Lehmann, E. & Ostergaard, C., Offshore Structures. Springer~erlag, Berlin, Germany, 1988 (in German). 23. Ostergaard, C., Schellin, T. E. & Stikan, M., Safety of marine structures: hydrodynamic calculations for compact structures. Schiff und Hafen, 31(1) (1979) 71-6 (in German). 24. Sfikan, M., Calculation of wave effects on compact structures of arbitrary shape. Schiffstechnik, 29(4) (1982) (in German). 25. Papanikolaou, A., On the evaluation of motions and loads of arbitrary bodies in waves. In Proc. Int. Syrup. on Ocean Space Util. '85, Tokyo, 1985, pp. 75-86. 26. IVlaruo, H., The drift force of a body floating on waves. J. Ship Res., 4 (1960) ]1-10.

478

T. E. Schellin, C. Ostergaard

27. Newman, J. N., The drift force and moment on ships in waves. J. Ship Res., 11 (1967) 51-60. 28. Pinkster, J. A. & van Oortmerssen, G., Computation of the first and second order wave forces on oscillating bodies in regular waves. In Proc. 2nd Int. Conf. on Num. Ship Hydrod., Univ. of California, Berkeley, 1977, PP. 136159. 29. Clauss, G., Siikan, M. & Schellin, T. E., Drift forces on compact offshore structures in regular and irregular waves. Appl. Ocean Res., 4(4) (1982) 20818. 30. Papanikolaou, A. & Zaraphonitis, G., On an improved method for the evaluation of second-order motions and loads on 3D floating bodies in waves. Schiffstechnik, 34(4) (1987) 170-211. 31. Schellin, T. E. & Kirsch, A., Low-frequency damping of a moored semisubmersible obtained from simulated extinction tests and mean wave drift forces. Appl. Ocean Res., 11(4) (1989) 202-1 I. 32. Hearn, G. E. & Tong, K. C., Added resistance gradient versus drift force gradient-based predictions of wave drift damping. Int. Shipbuilding Progress, 35(402) (1988) 155-81. 33. Standing, R. G., Brendling, W. J. & Wilson, D., Recent developments in the analysis of wave drift forces, low-frequency damping and response. In Proc. Offshore Technology Conf. (Paper No. OTC 5456). Houston, 1982. 34. Schellin, T. E., Jiang, T. & Sharma, S. D., Motion simulation and dynamic stability of an anchored tanker subject to current, wind and waves. Ship Technol. Res./Schiffstechnik, 37(2) (1990) 64-84. 35. Sharma, S. D., Drift angle and yaw rate touring tests in four quadrants-Part 2. Schiff und Hafen, 34 (1982) 219-22 (in German). 36. Oltmann, P. & Sharma, S. D., Simulation of combined engine and rudder maneuvers using an improved model of hull-propeller-rudder interactions. In Proc. 15th Syrup. on Naval Hydrodynamics. National Academy Press, Washington, DC, USA, 1984, pp. 83~-108. 37. Brix, J. & Kleinw~ichter, J., MTI-Stellungnahme zum Thema Aus-demRuder-Laufen von Schiffen. Hansa, 18(20) (1979). 38. Remery, G. F. M., Mooring forces induced by passing ships. In Proc. Offshore Technology Conf. (Paper No. OTC2066) Houston, 1974. 39. Jiang, T., Schellin, T. E. & Sharma, S. D., Maneuvering simulation of a tanker moored in a steady current including hydrodynamic memory effects and stability analysis. In Proc. Int. Conf. on Ship Manoeuvrability. (Paper No. 25) RINA, London, UK, 1987. 40. Jiang, T. & Schellin, T. E., Motion prediction of a single-point moored tanker subjected to current, wind and waves. J. Offshore Mechanics and Arctic Eng, A S M E Trans., 112 (1990) 83-90. 41. Viggosson, G., Field observations of ship behavior at berth. In Proc. Advances in Berthing and Mooring of Ships and Offshore Structures (NATO ASI, Series E: Applied Sciences Vol. 146). Kluwer Academic Publ., Dordrecht, The Netherlands, 1988, pp. 198-264. 42. Van Oortmerssen, G., Pinkster, J.A., & van den Boom, H. J. J., Mathematical simulation of the behavior of ships moored to offshore jetties. In Proc. 19th ICCE, Houston, 1984. 43. Van Oortmarssen, G., Forces related to motions of moored ships/analytical

The vessel in port: mooringproblems

479

methods of moored ship motions. In Proc. Advances in Berthing and Mooring of Ships and Offshore Structures INATO ASI, Series E: Applied Sciences Vol. 146). Kluwer Academic Publ., Dordrecht, The Netherlands, 1988, pp. 26581. 44. Van Oortmerssen, G., The mooring of ships: from skill to science. In Proc. Symp. on Ship Transport, Rotterdam, 1982, pp. 407-27. 45. American Petroleum Institute, Recommended Practice for the Analysis of Spread Mooring Systems for Floating Drilling Units, API RP-2P, 1987. 46. Schellin, T. E., Jiang, T. & Sharma, S. D., Numerical prediction of lowfi'equency surge of two moored floating production platforms. In Proc. OMAE (Vol. l-A). Offshore Technology, ASME (1991), pp. 165-74.

You might also like