You are on page 1of 87
3 Design Loads 3.1_GENERAL ‘The purpose of this chapter is to provide the port designer with guidelines on loads and forces to be used in the design of port related marine structures. All applied design loads and forces which should be considered in marine structures design may be defined in the following basic categories: = Loads and forees from the water side of the structure ‘= Loads from the land side of the structure Loads and forces related to dock operation + Load due to the temperature effect Waterside related loads and forces are basically environmental forces such as wind, waves, current, and ice. Land-side related loads are typically the weight of the struc ture, the vertical and horizontal compo- nents of soil pressure, and the hydrostatic pressure from the groundwater behind the structure. ‘Loads and forces related to the operation of the dock are basically those associated with the following: ‘Ship docking and mooring operations, in- cluding berthing impact, and mooring forces due to winds, waves and current acting on the vessel + Vertical and longitudinal friction forces between the ship and fenders due to tidal action, change in ship dra due to Joading/unioading, and the ship rubbing along the structure during berthing ma- neuvering ‘+ Loads and forces due to the operation of cargo handling and hauling equipment such as fixed, rai-mounted, and mobile cranes, dry bulk loading /unloading equip- ment, conveyers, miscellaneous tire and crawler track-mounted equipment, trucks and other vehicles, liquid-eargo handling arms and piping, and others ‘+ Uniform distributed loads of miscella- neous nature Loads due to temperature effects may have a substantial impact on the spacing between expansion joints. In general, these loads may generate considerable horizontal loadings on marine structures and be espe- 24a 244 Design Loads cially detrimental when applied to some specific structural components. ‘Additionally, loads acting on a marine structure can be classified as permanent, temporary, and special. Permanent loads include the weight of the structure along with the weights of ware- houses, sheds, and administrative offices all permanently installed on the marine struc- ‘ure, and also the weight of permanently in- stalled fixed cargo handling equipment. In further discussions, this kind of load will be referred to as dead load. (Note that lateral soil pressure may also be defined as a perma- nent load.) ‘Temporary loads and forces include those due to the environment and dock operation. The latter includes all kinds of live loads (surcharge, moving equipment), ship berthing forees, and mooring forces. Special loads include accidental loads, earth- quake forces and other unusual loads, ‘As with the design of conventional struc- tures, the loads, and their combinations, used in marine structure design are distin- guished as normal and extreme. Normal load combinations are those that ‘could exist under normal dock operations. The ‘most severe, realistically expected combina- tions of dead load and soil pressure, ship impact and mooring forees, loads generated by cargo handling and hauling equipment, and appropriate surcharge loads are usually selected as the normal design load combina- tion for a given structure. To proportion the structural elements exposed to normal load combinations, normal allowable stresses are used. ‘The normal loading, in combination with unusual forces such as an earthquake, an accidental load of miscellaneous nature or operation under a damaged condition, is considered to be an extreme load condition. In such cases, allowable stresses used in design would be appropriately increased (see “Load Combinations,” Section 3.8) Safe operation of any structure depends on the loadings subsequent to completion of construction, Because the structure is usu- ally loaded after it is built and not always as designed, the selection of design loads is a problem of statistics and assessment of probability. In some cases, neither the ‘owner nor the designer is sure what maxi- ‘mum loading the structure may be required to sustain in the course of its useful life. In addition, the selected loading is often sub- jective and largely depends on the designer's experience. To avoid divergent designs and to formulate reasonable design loads and load combinations, the designer should fol- low relevant guidelines, codes, and stand- ards, but also exercise common sense. Rea- sonable design loads are those that do not result in either prohibitive cost of structure construction nor in structure obsolescence shortly after commissioning. The designer also should exercise great care in formulat- ing design loads, particularly those gener- ated by mobile cargo handling and hauling equipment. 3.2_ ENVIRONMENTAL LOADS Wind, waves, current, and ice provide the principal environmental loads on marine structures, Typically, environmental design loads are based on storm conditions having an average expected recurrence interval of 50 years; some important marine structures may be designed on the basis of an expected event recurrence of up to 100 years, Because most of the port related marine facilities are located in areas sheltered from the action of sizable waves and strong cur- rents, the direct action of these environ- mental agents against marine structures are, in general, negligible. Under typical design conditions, berthing and mooring loads and lateral soil thrust are, as a rule, the dominant horizontal forces’ acting on a berthing structure. However, even in protected harbors the traveling vessels may generate substantial waves, the effect of which could be considerable on moored ships. However, at exposed locations (ie., breakwaters, offshore terminals, or some terminals located in inland waterways) the effects of wave and current forces must be studied closely. ‘In countries with a cold climate, ice rep- resents a major environmental load on dock structures. The ice may impose lateral thrusts against a structure, produce uplift and pile jacking loads, increase the weight of the structure, and abrade the structural material. 3.21 Wind ‘The magnitude and strength of a wind load acting on a dock structure depends on the velocity of the prevailing wind, dock orien- tation, the amount of structure exposed to wind, and the size of a ship moored along- side the dock structure. In general, the wind force that acts on a ship is proportional to the square of the wind velocity. The wind effect, which by its nature is dynamic, is usually decomposed into a static action (e.g., no variation in wind speed) and a dynamic action (e.g., gust) is covered by a special coefficient(s). Two components of wind forces are usually considered: one that acts per~ pendicularly to the dock structure and the other that acts parallel to it. The maximum wind foree is usually obtained when the wind is blowing at 90° to the ship's center- line, ‘Total wind force, P,, acting on a ship alongside a structure may be expressed as follows: bp, 0,0,( EA, sin* + YA, c05%) on Design Loads 245, ‘Table 3-1. Value of factor C, Length of the Ship 2m 60m 100m = 200m 08 065 05, where P,,= total wind force on the dock structure in Kilonewtons. k= 12, shape factor, which considers suction increase on the leeward side. C, = coofficient that considers the length of the ship; the recommended values of C, are given in Table 3-1 Cz = gust factor, average gust-factor value range 1.25-1.45, Note: when selecting a design value of C, the size of the ship is to be considered. Because the gust duration may be insufficient to de- ‘velop a full mooring foree or maximum fender strain on a large ship (the “inerta effect”), smaller values of Cy are usually used when dealing with large ships. Fang (1979) has re- ported on test results to verify wind and cur- rent loads on tankers 170,000-280,000 DWT. It has been found that 20 s of sustained gust. can fully excite the moored very large crude carrier (VLCC), whereas approximately a 1- min gust duration is required to induce the maximum ship response. EA, andEA, exposed to wind areas of the ship, cargo handling and hauling equipment located on the structure, and the dock structure area (in m?) above water line exposed to wind respectively in x and y directions. If the dock structure is occupied by ships on both sides, the sheltering effect of the windward ship is taken into account. The wind force on the sheltered vessel is usually assumed to be 50% of the value of an unsheltered vessel had it not heen sheltered against the wind. Note: The maximum wind load on a dock structure will be when the ship is in a light condition (maximum exposed to wind area). If accurate wind area information for a specific 246 Design Loads design ship is not available, then approxi- rate values obtained from Figures 3-1 and 5-2 may be used forthe preliminary design. 46 angle of wind direction to ship centerline Py ~ specific wind pressure (in KPa), which varies with the square of the sustained wind velocity. The sustained wind velocity is the average speed over 1 min, The value of 2, may be obtained from Pe =474x10-V2 (8-2) where V,, = velocity of the wind (in km/h). The value of p,, varies from place to place and usually is specified by local building codes. Wind speed is usually assumed to be the sustained wind speed 10 m above the water surface (V;o), obtained from available wind records. If the wind is recorded at an elevation other than 10 m, the wind speed at the 10-m level is determined using Ven Vio= 10/2)", (3-8) where Vip = wind speed at an elevation 10 m over surface of water Z = elevation of recorded wind (in m) V, = recorded wind speed at elevation Z. For fetch length, that is, the horizontal ex- tension of the wave generating area, less than or equal to 15 km, the wind speed recorded over land (V,) should be adjusted to obtain the over-water wind speed (V,.) nee 0") expec to wind aes (ns 103) Figure 3-1. Approximate projected (vessel in ballast condition) to wind area as function of tanker DWT. =1y, (a4) If the fetch is longer than 15 km, the ratio R=V,/V;, can be obtained from Figure 3-3. Design Loads 247 For preliminary design Udovichenko and ‘Yakovlev (1976) recommended the following equation to determine the ship's exposed to wind area, A, when she is in a light condi- tion: More information on wind speed and Aq=alt -5) wind forces calculations can be found in NAVFAC DM-26.2 (1982), Gaythwaite Where (2981, 1990), Headland (1995), Braun A = exposed to wind area of @ ship in light (1989), Séhngen (1995) and PIANC (1995). condition (m). so. 40. 7° z & ~ i Fa 10. op Lj tj | ° 7 2 > 7 3 ° ‘Ship exposed to wind area (m? x10) Figure 3-2. Approximate projected to wind area as a function of vessel capacity: 1. general cargo ‘ight condition) 2. container ship. ‘Table 8-2. Coefficient a in o9. (2-2) ‘Type of Vessel 00m 150m ‘Mixed general cargo/passenger O12, O12 General cargo on 010 ‘Tankers oa 010 Length (L,) 300m 200m 250m and More 012 oat oan 009 8, 0090.08 oor 248 Design Loads 1, = maximum length of a ship (in m). a= empirical coeficient which takes into ac- count the ship's shape and length. The ree- ommended values of « are given in Table 3-2 Additionally, it should be recognized that the area of a ship exposed to the wind is not sym- ‘metrical about the midship line; the latter may help to develop « moment of rotation about the ship's center of gravity. In designing the dock structure for wind forces, it is usually assumed that cargo han- dling equipment such as cranes and loading towers would not operate with a wind stronger than 25 km/h. It is also customary to consider that a ship would not remain alongside the dock in light condition in a severe storm; she would either take on bal- last or leave for sea. (p, = 0.25 kPa is usu- ally considered adequate for the dock struc- ture) In designing the dock, use of unit wind pressure greater than 1kPa should be Justified by special requirements. For the preliminary design, values of de- sign wind load (KN/per linear meter of ship) can be obtained from Figure 3-4. The val- OvERLAND ‘ues of wind load presented in Figure 3~4 are obtained from miscellaneous sources and should be considered as rather conser- vative. The wind load is transmitted to the quay structure via the ship's mooring lines, or fender units effectively in contact with a ship. In some cases, the berth could be partly shielded against the seaward wind by exist- ing structures, such as warehouses, office buildings, or by specially constructed wind- protective structures. In the past, such structures have been effectively used in ports at Le Havre, Marseilles, Dover, and other locations (Minikin, 1963). 3.2.2 Currents In the design of marine structures there are three major kinds of currents to be considered: 1. River currents 2, Tidal currents 3. Wind-induced eurrent in harbors 6 mWOTS. 3S os ees Hou INOSPEED, Figure 3-3. Ratio (R) of over-water (V,) to over-land (V;) wind speed as a function of overland wind speed. (After Resio and Vincent (1976)] Design Loads 249 | ind toed (KM per ue. m of hie) Figure 3-4. Approximate wind load per linear meter of ship, River and tidal currents are usually se- lected from available statistical data, whereas the wind-induced currents in still water are usually taken as equal to 1% of the velocity of sustained wind at 10 m above the surface of the water. Typically, current velocity decreases slowly with depth, and current load (drag) on a dock structure de- pends on the square of the current velocity and the shape and magnitude of the exposed-to-current areas of both the fully laden ship and the submerged part of the structure. The force of the current exerted on the ship/structure system in both longitudinal and transversal directions (P.) (in kN) can be obtained from Pwe Dave G6) where smpirical drag coefficient; ¢ = 0.5 to 1.0 (aN s?/m*). Smaller values are used for ships and larger values for the dock structure, These values of ¢ are conserva: tive and are recommended for the prelim- inary design stages only. The true value of ¢ may be obtained from model tests of the design ship and dock structure. V, = velocity of the current (in m/s) A = exposed-to-current areas ofthe fully laden ship and the structure system (in m). In a strong current, a dock structure is usually placed parallel to the stream. The design current force is typically based on a current having an expected average recur- rence interval of 5 years. Current force acting on the vessel is usu- ally transmitted to the dock structure through the mooring lines and very seldom by the ship bearing against the fender sys- tem, For more information on the direct, quasi-static current forces imposed against ship/dock structure system, the reader is referred to the British Standard Code of 250 Design Loads Practice for Maritime Structures, Part 1 (1984), NAVFAC (1968, 1982), Palo (1983), Gaythwaite (1981, 1990), Bruun (1989), Headland (1995), and PIANC (1995). ‘Newman (1977) provides an in-depth treatment of hydrodynamic current forces on ships, where, under certain site condi- tions, strong current may produce substan- tial standoff forces exerted on the vessel. In 1968, serious problems associated with a strong tidal current (velocity of about 5.5 knots) was experienced during construction and operation of the Immingham Oil Termi- nal for 100,000-DWT tankers on the Hum- ber River in the United Kingdom (Jackson, 1973). One problem that appeared was vio- lent oscillation of unbraced piles across the flow. Even when restrained by bracings or a superstructure, the piles were not immune to some vibration and extensive additional bracing was subsequently added in order to ensure safe terminal operation. Another problem observed, during a high current the moored ship was pushed off the berth by considerable standoff force that was built up between the ship and the pier structure. The cause of this force attributed to the effect of slowing down water passing through the pier piling, resulting in a pres- sure head by raising the water level. The increased head of water in the area of the piles deflected the current outward, with the result that there was a flow across the berth line. As a result, a ship moored to a piled berth structure was pushed away from the berth by a combination of the differen- tial head and the angular current (Fig. 3-5). The differential head across the moored ship (Ah) was found to be equal to Ah = C,V2/2g, where the coefficient C, ~ 0.42, V is the undisturbed current (flow) velocity, and g is gravity acceleration. Khanna and Sorensen (1980) and Wood et al. (1980) reported on model test results carried out on a 1:100 scale model to study the standoff effect phenomenon. They found the value of coefficient C, to range from 0.1 to 0.15 and 0.07 to 0.12 for scaled current velocities of 5 and 7 knots, respectively, and dependent on the berth location. The ap- proximate one-third difference in C, value measured in the above experiments and those measured at Immingham Oil Termi- nal were mainly attributed to the different area densities of piling. Specifically, at Im- mingham it was about 3% according to Khanna and Sorensen, and in the Wood et al. experiments it was equal to about 1% of the structure's effective area. However, in the latter case, a 16% increase in pile den- sity did not alter standoff forces signifi- cantly. More details on the effects of pile density and configuration on standoff forces is found in the work of Ball and Wilcock (1981), and PIANC (1995). In general, the reduction in the depth/ Graft ratio from 1.5 to 1.32 resulted in about a 35% increase in standoff force. The mo- tion of the moored ship usually started when the flow of water exceeded a threshold ve~ locity of approximately 2 knots. Under such conditions, substantial dynamic mooring line loads (two to three times that due to the standoff force) are encountered are attributed to strong surge-yaw and sway motions. A phenomenon similar to that at Im- mingham Oil Terminal can be experienced by a ship moored alongside a berth struc- ture located in close proximity to the river- bank or in a narrow tidal estuary with strong currents. In this case, the acceler- ated flow of water around one side of the ship may result in differential head (a hy- draulic gradient across the ship) pushing the ship away from the bank. Therefore, at terminal sites with strong current and where a potential for standoff forces exist, the designer should carefully investigate the problem, as no typical de- sign solution or theory exists in practice at this time. In conclusion it should be noted that an essential parameter to be considered in cal- culating the force of current, besides the current velocity, is the underkeel clearance, Design Loads 251 LAN [ons cevt/20 —— Cte ALA Figure 3-5. Standoff force on moored ship. ice, the vertical distance from the bottom of the ship to the bed of the basin. According to PIANC (1995), if the underkeel clearance is small, the current force can be up to six times the value of that in deep water. 3.23 Waves ‘As was stated earlier in this chapter, most marine terminals are built at protected lo- cations and therefore they are not exposed to direct, wave forces of significant magni- tude, Goda (1985) suggested that any harbor where a significant wave, H,, is less than approximately 1.0 m, even under design storm conditions, can be considered a calm harbor; he offers the following criteria for harbor tranquility: 1. The harbor should have a broad interior. 2, Any portion of the harbor that can be viewed through the entrance should have 252 Design Loads a natural beach or wave-absorbing struc- ture. 3. Small craft should not be viewable from ‘the open ocean at any angle. 4. Apportion of the waterfront should be wave dissipating in character. 5. Wave reflection from the back side and the use of energy-absorbing quay walls such as perforated or slit caissons should be considered, In protected harbors, berthing and moor- ing forces typically are much more severe than those which can be generated by wave action, and therefore the latter are usually neglected. However, even in a protected harbor moored ship movements can occur, as depicted in Figure 3-6, due to the re- sponse of the water surface in the harbor to excitation by open ocean long-period waves known as seiches. Seiches may cause sub- stantial damage to both ship and mooring structures. Long-period waves, or long waves, in the open sea are rather small, but they are amplified in coastal regions due to a variety of nonlinear effects. When entering en- closed bays or harbors, long waves may be further amplified and natural water oscilla- tion may develop. It is very difficult to dissi- pate energy of these waves. Their periods are close to the natural periods of surge, sway, and yaw for medium and large ships, which, combined with usually low damping of the ship mooring system, may result in resonant phenomena (Jensen and Warren, 1986; Stammers and Wennink, 197; Viggosson, 1988). ‘Harbor protection from long waves will require construction of a very long break- water, that in most practical cases is uneco- nomic. Instead, careful design of the berth mooring system and correct positioning of the berth with the harbor is needed to mini- mize response of ships to long waves. This can be best achieved by physical modeling of port (harbor) layout. Furthermore, waves created by a passing vessel under specific conditions may pro- duce standoff forces on a moored ship simi- lar to those discussed in the previous sec- tion. Naturally, structures located offshore should be designed for sea waves of sub- stantial magnitude. The purpose of this sec- tion is to introduce some basic state-of-the- art concepts for wave loads under the above-noted conditions, In general, the interaction between wave forces and a moored ship is a very complex phenomenon, and analytical treatment of the problem typically involves the combined effects of wind, wave and current loads, coupled with ship motion and response of the mooring system. Progress in develop- ment of an analytical solution can be fol- lowed in the works of Wilson (1951, 1958, 1959), Van Oortmerssen (1976), Van Oort- merssen et al. (1986), Bomse (1980), Vasco Costa (1983), NATO (1965, 1973, 1987), and PIANC (1995). In the above works, sophisticated mathe- ‘matical modeling techniques have been pro- posed. The proposed methods formulate a mathematical approach for obtaining water surface response, ship motion, and mooring forces due to a given offshore wave spec- trum for a vessel moored at a particular location in the harbor. It should be noted that at their present stage of development, mathematical models of moored ships are best used in preliminary stages of port de- sign, where several alternative solutions are usually investigated to achieve the most efficient design. Physical scale modeling may be employed to calibrate (validate) the mathematical model. Because physical scale models are expensive and time-consuming they are usually employed to model the final harbor layout obtained on the basis of a calibrated mathematical model. Principles of hydraulic modeling are deseribed in most fiuid mechanics texts. ‘As a rule of thumb, standoff forces on a moored ship, generated by passing vessels, are proportional to the square of the speed of the passing ships and the separation distance. Additionally, the underkeel clearance, the vessels’ block coefficients, the softness of mooring lines, and the ratio of length of moored ships (Lp,) to length of passing ships (L,,) affect the maximum standoff forces and subsequently, the mooring line loads. Forces on the moored ship are greater when L,,/Ly * 1 and are slightly notice- able when L,,/Ly,* 1, even for a very close separation of both vessels. By its very nature, the effect produced by the passing ship on a moored vessel is very similar to that of a vessel transiting a channel in close Design Loads 253. proximity to the bank or another vessel. For details, consult Chapter 2. Additional infor- mation on dynamic effects of ship passage ‘on moored vessels is found in Wang (1975), Jensen et al. (1990), and PIANC (1995). As was stated earlier, the evaluation of hydrodynamic loads, associated with long- period waves, and standoff forces on a moored vessel generated by passing ships on marine structures is extremely complex. The problem's complexity is aggravated by the fact that the added mass and damping coefficient of an oscillating vessel are dif- ferent for each of the six degrees of freedom (Fig. 3-6) and are dependent on the fre- quency of the ship's oscillation and, to a great extend, her underkeel clearance. De- tailed treatment of this subject is beyond the scope of this text. For a comprehensive treatment of the subject matter, the reader is directed to the sources cited earlier in this section. Offshore marine terminals, such as transhipment islands or individual cargo loading/unloading berths, are exposed to substantial wave loads. Perimeter walls of the offshore islands are usually of verti- cal configuration and typically exposed to wave forces similar to those acting on breakwaters of the same configuration. Forees of this type are discussed in detail in Chapter 10. 3.2.3.1 Wave Force on Circular Piles and Large-Diameter Cylinders ‘The majority of individual offshore berth structures, such as piers, breasting and mooring dolphins, cargo handling equip- ment supporting platforms, and approach trestles are typically of the open-type con- struction, which, in general, comprises of a concrete deck structure supported on piles of miscellaneous construction. Circular piles are used more often than piles of other cross sections because they offer less resis tance to wave or passing current action. 254 Design Loads ‘The approach to calculating wave forces on piles depends on the ratio of the pile diameter (D) to the incident wavelength (L). When the ratio is small (D/L < 0.2) the wave motion is relatively unaffected by the presence of pile, and the analytical ex- pression of the wave force acting on a pile (Fy) is normally calculated using the well- known Morison equation (Morison et al., 1950). This equation is based on an as- sumption that the pile (cylinder) does not disturb the waves and consists of the sum of the inertial force (F,) and the viscosity drag force (Fp): Fy=Fo+ on where Fy =o. sc( 7) Dulul (8) @ and yw) 2D*) (du _ neal (TF) 6% where Cy = empirical dimensionless drag_coeffi- cient taking into account the resistance of the pile against the flow pressure Cy, empirical dimensionless inertia coeffi- cient taking into account the resistance of the pile against the acceleration of water particles = gravity acceleration (m/s*) ‘yy = density of water (KN/m*). For freshwa- ter, yw = 10 KN/m® and for seawater ‘yq = 10.25 kN/m? = horizontal component of the instanta- neous orbital velocity of the water par- ticles at the studied pile location (m/s) du/dt = horizontal component. of the instanta- ‘neous orbital acceleration of the water particles at the studied pile location (m/s) D=pile diameter or (at noncircular piles) characteristic width of the structural ‘member (m). ‘The velocity and acceleration of the wa- ter particles used in the Morison equation are calculated from the conventional wave equations [see for example, the U.S. Army (1984), Shore Protection Manual]. ‘The em- pirical dimensionless coefficients Cp and Cy for cylindrical piles are given in Table 3-3. For detailed information on the coeffi- cients Cp and Cy, the reader is referred to Isaacson (1988) and Chakrabarti (199). Some preliminary data on Cp and Cy, ob- tained from field measurements in the surf zone, are reported by ‘Tsrum (1989). Values of Cp and Cy for square piles, H-piles, and piles of other shapes are found in the British Standard Code of Practice for Maritime Structures (1984), EAU (1990), and Det Norske Veritas (1977). The assessment of drag and inertia co- efficients is made not on the basis of theo- retical considerations, but principally by means of measurements on models and pro- totypes. They correlate the computed force to the existing force and are greatly depen- dent on the Reynolds number. They are very sensitive to the pile roughness caused by such effects as fouling, which may sub- stantially increase the pile's projected area, ‘The Reynolds number (Re) is given by Re=U,D/v (s-10) where kinematic viscosity (m?/s) approximate maximum horizontal water- particle velocity (m/s) at 2 = m (see def- inition sketch, Fig. 3-7) ‘Table 3-3. Drag Cp) and inertia (Cy) coefficients {or eylindrieal piles Cy ____ Reynolds Number (Re) = 20 > 25x 10° 1207 15-20 2x 105 x 10° o7 - > 5x 108 ~ 15 5x10" 12 - << 25 x 10° Note: Cp ~ 0.7 and Cy 15 are typically used for most design tppleations. In deep water, = TH goeepns In = ae Lke ean ‘where 6 is the wave phase angle and U,, will occur at 6 = 0. mas) Additionally, the coefficient of inertia Cy takes into account the fact that the added ‘mass, subject to the force of inertia, has a ‘greater magnitude than the one correspond- ing to the volume of fluid displaced by the solid In order to obtain the total instantaneous ‘wave force the drag and inertia components must be integrated over the water depth Design Loads 255 and added vectorially. For this, the equa- tions for Fp and F, can be given by Fy= ff fpaZ 3 Sie (3-12) and ny [hae oa» where fo 086,| 7) Dui GH z iol BIER) o and z is the depth in terms of vertical Figure 3-7. Wave force on vertical ple. 256 Design Loads Figure 8-8. Wave force on battered ple distance along a coordinate axis with its origin at the bottom (Fig. 3~7). ‘The problem also could be solved by us- ing nondimensional nomographs and tables provided in the U.S. Army Corps of Engi- neers’ Shore Protection Manual, Vol. 1 (1984) and in NAVFAC DM-26.2 (1982), or by commercially available computer pro- grams, [eg., Pile developed by Bridgeman (1987). ‘The maximum drag force occurs at the ‘wave crest position and the maximum iner- tial force occurs at about one-quarter wave- length ahead of the crest. The practical ex- ample is given in Isaacson et al. (1995). For nonvertical (sometimes called batter or raking) piles, the wave phase angle 0, for the local coordinates X,, Yo, and Zo, is different for each individual section, as de- pieted in (Fig. 3-8). Because wave forces are perpendicular to the pile axis, it is pos- sible to calculate them by Morison’s equa- tion using components of velocities and ac- celeration perpendicular to the pile U.S. Army Corps of Engineers, 1984). Exper ments by Brusnall and Loftin (1951) sup- port this assumption. They found that the drag force component depends on the resul- tant velocity rather than on the velocity ‘component perpendicular to the pile axis. With due consideration to these experi- mental observations, the following proce- dure has been recommended for calculating wave forces on batter piles. For a given location on the pile (Xp, Yo, %) the force per unit length of pile is taken as the hori- zontal force per unit length of a fictitious vertical pile at the same location. For a detailed procedure and numerical examples the reader is referred to U.S. Army Corps of Engineers (1984). It must be noted that besides drag and inertial forces, a trans- verse force due to alternate eddy formation and shedding generated downstream from the pile known as the “lift force” (F,), will also result from the wave-pile interaction. The lift force acts perpendicularly to both wave direction and pile axis and is analo- gous to a drag force. It reaches its maxi- mum value at the same moment as the drag force does, but could be four to fives times greater than the drag force (Rogan, 1978). For design of rigid structures, NAVFAC ‘DM-26.2 (1982) recommends adoption of a conservative value of Cp for determination of lift forces, regardless of the foree direc- tion. Due to the oscillatory nature of the lift force in the design of flexible structures, the dynamic response of the structure must be considered (Laird, 1962; Chakrabarti, 1991). The magnitude of this lift force can be de- termined by nxo(*2epe aw B where C,, is the lift force coefficient which can be taken as approximately equal to Cp/8 for circular cylinders and steady flow velocities. It generally does not depend on pile roughness. For more information on C, values under different conditions, the reader is referred to Chang (1964), Sarpkaya (1976), and Sarpkaya et al. (1977), Both linear and nonlinear wave theories are used for the calculation of wave forces. However, use of nonlinear wave theory [e.g., the fourth and fifth order of Stoke's wave theory can produce better results (Tsuchiya and Yamaguchi, 1974; Isaakson and Qi-Hua Zuo, 1989)]. If nonlinear wave theory is used, the Morison equation is usually solved by numerical integration. ‘The moment of wave force about the mudiine is given as follows: M, = [fot ede (3-17) where z is the depth in terms of vertical distance along a coordinate axis with its origin at the mudline level (Fig. 3~7). The single small-diameter pile structure is rare in marine applications. The usual open-type piled structures are piled piers, platforms, or dolphins. It is generally ac- cepted that unless piles are placed too close to each other (1.5D or less) the pile density may be ignored in calculating total wave force on a pile group (Quinn, 1972; Bushnell, 1977). Wave forces on a combina- tion of piles in group are calculated by plot- ting the forces on a single pile as a function of wave phase angle (0). Design Loads 257 ‘The combination of piles is then superim- posed over the plot of wave forces, and the maximum force for the system is found by summing the forces on the individual piles for small changes in phase angle (8) (Fig. 3-9). The phase difference between piles is, given as AO = (2n/L) Ax (3-18) where = wavelength ‘Ax = spacing between piles in the direction of wave advance (j., along the x-coordinate axis). Assuming that the two-dimensional wave is not changing while moving, the force ap- plied on individual pile and pile groups for different wave phases may be determined as follows (Rogan, 1978): Foca) ~ Fon 608 8, £98 8,1 Fy sin 8, (3-19) Megay = Moya 608, 08 8+ Mr 5 (20) where Fyg) and Mga) correspond to the total force and moment acting on the nth pile in a pile group, respectively. The total Fm - ory le Figure 2-9, Wave force on pile group. 258 Design Loads force (Fr) and moment (Mycn)) on a structure composed of n piles will then be expressed as Fay E Faw (-21) 7 Muay=E Muy @-22) ‘The maximum wave force on a group of piles generally occurs when the central pile is located near the wave crest. Wave forces on a pile bracing system are determined in a manner similar to that used for vertical piles; that is, the wave force per unit length, 4.2, of pile (Fig. 3-7) is also the drag foree per unit length of horizontal bracing when Az =D , where Dz is the diameter of the horizontal brac- ing. Naturally, the wave force on the hori- zontal bracing is a function of the bracing position above the seafloor. Because brac- ings are typically beams of small diameter, only the drag force is usually considered. ‘The wave load on an angular bracing may be obtained by the integration of loads per unit length at different elevations. Although notable progress has been made in recent years regarding understanding of the breaking wave-pile interaction, well- founded recommendations on a part of the breaking wave force on pile are sparse, Apelt and Piorewiez (1987) state that no theory is available for predicting wave forces in the breaking zone, although a considerable amount of experimental work has been re- ported (Swift, 1989; Terum, 1989). For practical engineering applications in shallow water, the inertia component of breaking wave forces on a pile is normally ignored, and the drag coefficient is usually taken as equal to 1.75 to 2.5 times greater than the one used in the deep water (US. Army Corps of Engineers, 1984; NAVFAC DM-26.2, 1982; Rogan, 1978) Because the wave crest velocity (u,) in shallow water approaches the wave celerity at breaking, it is usually given as equal to (gd,)°%, where d, is the water depth at breaking, which is approximately equal to the breaker height (H,). Current velocities of any kind (steady, tidal, or generated by the sustained wind) should be added veetorially. ‘As was stated earlier, the Morison equa- tion is used to obtain wave forces on a small-diameter pile (D/L < 0.2). This equation simplifies the general problem by assuming that the pile (cylinder) does not disturb the incident wave in any way. How- ever, as the cylinder’s diameter becomes large relative to incident wavelength, the large-diameter cylinder causes reflection and diffraction of the incident waves, which makes Morison’s equation unapplicable, and a diffraction theory must be used instead. ‘The diffraction theory assumes that the fluid is homogeneous and inviscous, and that the fluid motion is irrotational. Addi- tionally, the wave heights are assumed small relative to the wavelength so that terms involving wave steepness to the second (or higher)-order power may be neglected. MacCamy and Fuchs (1954) developed a linear diffraction theory for calculating wave loads on a vertical circular cylinder extending from above the water surface to the mudline. This theory was supported by limited experiments con- ducted by Chakrabarti and Tam (1973) for D/L up to 0.54. ‘According to MacCamy and Fuch (1954), the total force on a large-diameter cylinder (is given by pail? Gat fa tanb(kd) cosCangle) (3-23) where P= 1/8 ~ fluid density 1 fe “loos apy” where i = first derivative of the Bessel function of the first kind Y{ = first derivative of the Bessel function of the second kind for the angle (wt)— a; here a= arctan(J,/Y,) and @= 24/T in wave circular frequency k= 2n/L 1 = wave period with 9,, L, H, and d as previously defined. ‘Mogridge and Jamieson (1976a, b) have presented a simplified graphical version of the linear diffraction theory of MacCamy and Fuchs to estimate wave forces and mo- ments on large cylinders. Garison et al (1974) and Hogben and Standling (1975) have extended the diffraction theory to structures of arbitrary geometry. Theoreti- cal treatment of the diffraction theory is found in Sarpkaya and Isaacson (1981). Ac- cording to Sarpkaya and Isaacson, the total wave force on a large-diameter cylinder can be obtained from y- (Zpunoten(222)oy cater) 2 where all of the terms are as previously defined for small-diameter cylinders. Although considerable efforts have been spent on research with wave forces on verti- cal cylinders subjected to regular and ran- dom waves, studies on the wave-induced dynamic pressure distribution around verti- cal cylinders, especially due to random waves, are rather limited. The most recent study to date on wave impact pressures on large-diameter vertical circular cylinders in deep water has been conducted by Sundar et al. (1989) and Khalfin (1990). It has been concluded that. the values of the dynamic pressure coefficients derived from the pres~ sure spectra are in satisfactory agreement with the linear diffraction theory of Mac- Camy and Fuchs. Bridgeman (1984) has Design Loads 259 introduced to the market a computer pro- gram based on MacCamy and Fuchs’ origi- nal formulation of wave load on large cylin- ders. For a discussion on wave loads on square caissons and conical-type structures, the reader is referred to works by Mogridge and Jamieson (1976), Jamieson et al. (1985), and Rahman and Chakravarty (1986) 3.23.2 Wave Load on Continuous Vertical Wall When the structure width is longer than the incident wavelength (D/L > 1.0), it can be treated as a continuous wall exposed to the wave force per unit length along the wall. This case of wave load is treated in virtually all general textbooks and hand- books on coastal and port engineering [e.g. US. Army Corps of Engineers (1984), Gaythwaite (1981, 1990), Braun (1989), Herbich (1991/1992), and many others. ‘The extensive discussion on wave loads and their impact on vertical walls is found in special issue of Coastal Engineering (Vol. 22 (1994)], in Proceedings of International Workshop on Wave Barriers in Deep Waters (1994), Osaka, Japan, published by Port and Harbor Institute, Yokosuka, Japan, and in Proceedings of the Annual Offshore Technology (OTC) Conferen- ces, Houston, Texas, and International Off- shore and Polar Engineering (ISOPE) con- ferences, In general, where the water depth at the face of the wall is deep enough (ie., equal to approximately 15-20 times the incident wave height), the wave will be reflected and form a standing-wave pattern, also known ‘as clapoties. In general, itis assumed that a standing wave is equal to twice the height of the incident wave. In shallow water (eg., H < 15d, where His the height of the incident wave and d is the depth of the water), the wave may break against the wall, thus generating greater forces on it. 260 Design Loads On the basis of his extensive laboratory tests of the breaking waves impacting (by plunging) directly on vertical, 10°, and 30° backward inclined walls with 1/10 fore- shore slope, Kirkgéz (1992) concluded that within a range of water depth conditions used in his investigation the most frequent location of the maximum impact pressure for all three walls tested remained almost at the still-water level Figure 3-10 illustrates typical wave load diagrams attributed to the action of stand- ing and breaking waves. Wave forces gener- ated by a nonbreaking wave (Fig. 3-10a) generally are calculated using the methods of Sainflou and Miche-Rundgren, which ac- count for partial reflection, as described in step-by-step instructions and graphical aids in the US. Army Corps of Engineering (1984) Shore Protection Manual. ‘When waves break against the wall, the Minikin (1963) and Goda (1985) methods are generally employed. The peak dynamic pressure which is assumed at the sea water FS ese » Figure 8-10. Wave pressure on a vertical wall (a) by a standing (reflected) wave and (b) by a breaking wave, H and I, are respectively the height of an incident wave and of a breaking wave; hy is the setup ofthe water level, level (SWL) in the Minikin method is given by the semiempirical formulation. The total dynamic force that is centered about the ‘SWL is added to the hydrostatic pressure to obtain the total profile of the pressure dia- gram (Fig. 3-10b). Again, complete step- by-step instructions and graphical aids for applying the Minikin method are found in the U.S. Army Corps of Engineering (1984) Shore Protection Manual. A complete dis- cussion on wave forces acting on vertical wall breakwaters is given in Chapter 10. 3.3_ MOORING LOADS ‘Mooring forces acting on a berth structure arise from environmental forces, passing vessels, and forces due to ship berthing ma- neuvers. They are transmitted to the struc- ture via ship mooring lines secured to the mooring hardware such as bollards, bitts, quick-release hooks, and other mooring ‘As was mentioned in the previous sec- tions, at most sea port locations wind is a dominant force to be considered in mooring system design; at inland ports, located along riverbands with strong currents, the force of current may control mooring forces; at exposed offshore locations, a combination of all environmental forces (e.g., wind, waves, and current) must be considered. In the latter case, berth operations are usually limited to certain acceptable weather conditions. Standoff forces due to current and/or wave effects, as well as ship surging mo- tions, due to effects of long-period waves must also be considered at certain locations (as described in previous sections of this chapter). Naturally, vessel size and type are key issues in the determination of mooring forces on a berth structure and the design of mooring systems, ‘Typically, the general arrangement for mooring lines for a ship lying alongside a marginal wharf is as shown in Figure 3-11, and a typical arrangement of mooring lines, at tanker berth, is shown in Fig, 3-12. 3.3.1 Mooring Lines Arrangement Ship mooring lines (hawsers) are usually of three types: bow and stern lines, springs, and breasting lines. Bow and stern lines, are used to resist current and wind forces imposed against the vessel and acting par- allel to the berth. These lines are attached to the respective ends of the ship and usu- ally make an angle of 30°45" with the face of the quay. Larger angles are usually not recommended. Springs are also used to ab- sorb wind generated and current generated forces acting parallel to the quay direction, but they are also used to prevent the vessel Figure 3-11, Marginal wharf. Typical layout of ‘mooring lines: 1—dock structure; 2—ship; 3—bollard; 4 bow line; 5—stern line; 6—breasting ine; spring line, Design Loads 261 from surge motions. They run fore and aft, typically at an angle of 5°-10° to the quay face. Breasting lines are used to provide lateral resistance to wind and current forces imposed on the vessel. These lines are nearly normal to the quay face. In general, a minimum of four lines are required to safely secure a ship at a berth; however, in practice, the actual number of lines in each category basically depends on vessel size. In the case of mooring a large ship, as many as 24 (and more) lines may bbe employed. As mooring lines are usually regarded as shock-absorbing elements, they should not be less than about 30 m in length to provide for sufficient elasticity. ‘To be efficient, mooring elements, such ‘a8 mooring and breasting dolphins, must be properly positioned to handle the expected range in ship sizes. This often means that additional mooring dolphins are spotted be- tween the ones on the extreme ends of the berth. Sometimes it is also necessary to have the breasting dolphins double to pro- vide for sufficient ship support and to han- le spring lines. Where practical (eg., at berths where berthing is only on one side), the mooring dolphins should be set back from the breasting dolphins by as much as 30-45 m to limit the upward component of Figure 3-12. Tanker berth. Typical layout of mooring lines: hip; 2—mooring line; loading /anloading platform; 4—breasting dolphin; 5—mooring dolphin; 6-—access ‘trestle; 7—gangway. 262 Design Loads the mooring pull (vessel in light condition) and also to reduce navigational hazards. More information on this subject is given in Chapter 8). Bulk cargo terminals frequently employ the same concept for mooring arrangements as is used for tanker terminals (Fig. 3-12). At these terminals, however, a greater number of breasting dolphins may be re- quired, especially in the cases where the ship needs to be shifted into position for access to all hatches by loading /unloading equipment. 3.3.2. Mooring Line Materials ‘Mooring is provided by mooring lines ex- tended from the ship and secured to moor- ing hardware, fastened to the berth struc- ture. All mooring lines are typically made from the same material. They are usually made from natural (manila) or synthetic (nylon, polyester, terylene, polypropylene, etc.) fibre materials, or galvanized steel (wire ropes). For better elasticity, individu- als fibers are woven together into strands, which, in turn, are woven into ropes accord- ing to specific patterns. Fibre ropes are usu- ally of three- to nine-strand construction, and steel wire ropes typically consist of six strands, each containing 19, 24, or 37 indi- vidual wire filaments built around a fibre or steel core. A detailed discussion on the elas- tic properties of mooring lines is found in Wilson et al. (1958) and Wilson (1967). For more information on steel mooring cables the reader is referred to NAVFAC (1986), ‘Tsinker (1986), and Gaywaite (1990). ‘Natural fibre lines are relatively cheap, easy to handle (when dry), but have a very moderate resistance against abrasion and therefore have a relatively short lifetime. ‘These lines are sensitive to high tempera- tures (due to sunlight or friction) and chem- icals. They absorb water easily, making the handling process cumbersome in wet condi- tions. Synthetic fibre lines also are rela- tively cheap; however, have much longer lifetime than lines made from natural fibres. Compared to natural fibres, the synthetic fibre lines have a higher load/ diameter ratio, are relatively light, and are easier to handle. The elongation of syn- thetic lines is rather large and their strength characteristics can be affected in case of high temperature (eg., caused by friction). Steel wire ropes are relatively cheap and have a long lifetime if well maintained. They have a low elongation being proportional to the mooring line load and a high load/ diameter ratio. The disadvantage of these lines is that they are stiff and heavy and therefore are difficult to handle. Also owing to their low elongation rates, they are vul- nerable to a potential breakage under dy- namic load conditions. Sometimes steel wire rope lines are used in combination with synthetic tails which enhance the shock ab- sorbing characteristics of steel wire rope lines. In most cases tails are not longer than 10 m and are connected to steel wire ropes by means of shackles, protecting the tail eyes with leather or plastic sheathing, ‘These lines most often are used for mooring large tankers. ‘Typical values of breaking strength for different types of hawsers is provided in Table 3-4. The elastic stretch at break ranges from approximately 10 to 15% for manila ropes, to about 50% elongation for nylon lines. Consequently, a nominal work- ing load is approximately in order of 10 to 15% of tensile strength for manila ropes, and 10 to 20% for lines made from syn- thetic materials where the line stretch can be tolerated. The elongation of steel wire mooring ropes at break typically range from 2 to 5%. 3.3.3 Mooring Forces Mooring forces are calculated from the envi- ronmental design loads, as previously dis- cussed, and are assume to act on the largest Design Loads 263, Table 3-4. Breaking strength of hawsors (tonnes) bi 6x 24 Wire 6x37 Wire ameter Polyester Poly- Rope, Galvanized Rope, Galvanized Ga)_Gum) Manila Nylon (Dareon) _propylene Plow Steel Plow Steel 1 2 4 13 10 6 26 2 z 85 1 2 uv u - - i = - — 6 65 2 5 48 0 Py 100 3 3 1% 29 at 9 52 = = 4100 48180 136 86 = - 6 me = - - = vessel for which the facility is designed. It is necessary, however, to relate the calcu- lated forces to the breaking strength of the mooring lines. It is therefore recommended that the designer consult with the dock op- erator to obtain information on the design vessel and on number and size of mooring lines to be considered. ‘The critical wind load is transmitted to the structure when the wind is blowing nor- mal to the ship's centerline, pushing her out from the dock. When the wind force is symmetrical, the normal component of mooring force per bollard (Nj) is given by Ns= = (3-28) where P,,= the total wind force onto the ship (in kN) expressed by Eq. (3-1) ng = the number of simultaneous working bol- ards (mooring lines) ‘The actual number of simultaneous work- ing bollards (n,) depends on the length of the ship; unless special mechanisms are used to equalize mooring forces in hawsers, ng can be obtained from Table 3-5. ‘The British Standard Code of Practice for Maritime Structures, Part 4 (1985) rec- commends that when the ship is moored at, six points, one-third of the total force should be assumed at any one point, and when the ship is moored at four points, one-half of the total force should be assumed to act at any single point. ‘Table 3-5. Number of simultaneously working bollards (na) ‘Minimum Distance Length of Ship Between Bollards (om) =) np 250 20 2 150 25 4 250 30 6 = 300 30 8 ‘The magnitude of mooring force per bol- lard, and its vertical and longitudinal com- ponents, can be obtained by the following equations (see Fig. 3-13): Ne a 3-26 Qa Fina cos B oe) Va = @nsin p (2) T= QpeosacosB (3-28) Figure 3-18, Mooring forces 264 Design Loads In general, the normal component of the mooring force (Np) used in Eqs. (8-26)- (3-28) should be increased by at least 20% to allow for nonuniformity of distribution of mooring forces between working bollards. If the actual values of a and B are not known, then for the preliminary design they may be obtained from Table 3-6. As the ship mooring lines are generally not fully stressed simultaneously and the mooring line forces partially cancel each other, in the absence of actual mooring line forces supplied by the port operator, the values in Table 3-7 and 3-8 should be used for the preliminary design. ‘The values given in Table 3-7 are suffi- ciently accurate for seagoing vessels, and the values given in Table 3-8 can be used for the design of inland port mooring structures. It must be noted that bollards at the ends of each berth are usually designed for double the value of mooring force used for @ regular bollard. At docks where strong cur- rents exist the values of Qp, obtained from Tables 3-7 and 3-8, should be increased by 10-25%. A smaller value applies to smaller vessels and a larger value to larger vessels. Normally, bollards are spaced uniformly along a quay face and symmetrical about ‘Table $-6. Values of « and 6 in Eqs. (3-36)-(3-28) 8 Vessel Bollard Taden Vessel in Light ‘Type Location @ Vessel Condition Seagoing Atthequay edge 30" 20° 10 ‘Second line of 0" rd 20° bollards Riverboat Passenger Avthequay edge O45" = 30° 30" ‘vessel Cargo vessel ___Atthe quay edge 0-90 = 80" 30" ‘Table 8-7. Mooring forces Qp generated by seagoing vessels Vessel Vessel Displacement Line Pull Displacement Line Pull onnes) dan onnes) en) 2,000, 100 100,000 14000 10,000 300 200,000 1500 20,000 600 > 200,000 2,000 50,000 800 — — Table 3-8, Typical values of Qp transmitted by inland waterway ship mooring lines to bollards (kN) ‘Maximum Vessel Type F100 6001000 F600 3000 5,000 tonnes) o}000 20,000 60,000 100,000 000,000 > 200,000 200 30 100 50 180 200 250 150 enger ship cargo ship 600 800 250-300 1,900 1,500 2,00 the center of a quay section (between ex- pansion joints). Other arrangements are also used, depending on the actual length of the section and bollard spacing require- ments, Typically, bollards should be de- signed to hold several hawsers simultane- ously, and, therefore, the load per bollard must be adjusted accordingly. In most seaports, bollards are installed ‘at ground surface level. In inland ports, where the water level may fluctuate signif- icantly, bollards are usually installed on at least two levels, but quite often they are installed at various elevations to allow for ship convenient and safe operation. Some practical examples are found elsewhere in this book. ‘At exposed locations, a soft mooring sys- ‘tem comprising of mooring lines with nylon tails in combination with soft fenders are used (Khanna and Sorensen, 1980; Wood et al., 1980; and PIANC, 1995). Nylon tails provide an additional elasticity to the moor- ing system, thus reducing the load in the ‘mooring lines due to motions of the moored ship. Furthermore, in a multicomponent ‘mooring arrangement, the length of nylon tails can be varied in order to maintain approximately equal loads among lines un- der anticipated ship movements. Nylon tails are routinely used by oil com- pany fleets, particularly at exposed loca- tions. Soft fenders, although absorbing ‘equal amounts of the ship’s energy, produce smaller forces on the mooring structure than stiff fenders. They are provided to be compatible with the softer mooring lines. Pulling a ship against a fender system by pretensioning of mooring lines may reduce hor motion and thereby reduce the forces in mooring lines (Brunn, 1989). ‘The magnitude of pretension required in the mooring lines is not precisely defined Normally, a pretension to about 10% of a line's breaking strength will remove the ini- tial sag from a synthetic line. In practical terms, a relatively small amount of preten- sion on the order of about 10 tonnes is Design Loads 265 expected to be satisfactory at berths accom- modating large vessels. A potential “side effect” of the pretension of mooring lines is the creation of vertical up and down friction, forces between the ship and fender system due to tidal variation, or changing draft of the ship due to loading or unloading. These forces are then transmitted to the dock and have to be applied to the front of the dock structure for calculation of general stability and local strength. Naturally, fenders must be strong enough structurally to sustain this load. ‘The mooring system must be designed to limit ship movements at berth to prevent possible damage to the ship, mooring struc- ture, and loading/unloading equipment. Information on ship movement allowances obtained from miscellaneous sources is pre- sented in Table 3-9. An extensive research program was carried out by the Nordic countries to establish acceptable criteria of motion of moored ships (Jensen et al., 1990). Motion of moored ships were measured at a number of ports for various ships being loaded /unloaded by different gear, and ship acceptable motions were recommended based on interviews with ship crews and port operators. Based on these recommen- dations PIANC (1995) suggested somewhat more liberal motion criteria than those given in Table 3-9. More information on acceptable ship motions while under load- ing/unloading operations is found in Braun (1989), Bratteland (1974), Slinn (1979), ‘Thoresen (1988), Jensen et al. (1990), and PIANC (1995). Safe berth operation to a great extent depends on mooring systems reliability. Therefore, redundancy should be provided as much as is practical; factors of safety on the mooring system components typically range from 1.5 to 2.0, depending on the severity of environmental conditions and the nature and consequence of a failure. It must be noted that standard practice in the wire rope industry is to use a factor of safety of 266 Design Loads ‘Table 3-9. Movement allowance for ship at berth ‘Surge Sway> eave Yaw Piteh Roll ‘Type of Ship @ =) (x) G@ogree) (degree) _(dogree) ‘Tanker as 20 10 10 10 10 Ore carrier (crane operation)* 15-30 05-2005 10 10 10 Grain carrier (at elevator) 10 10 10 10 10 10 Container 08. os os: 10 10 10 o/Ro (side ramp)* 03-05 08-05 © 08-05 10 10 10 Ro/Ro (bow orstern ramp) OL oa 02 10 = 10 General cargo! 15 10 10 20 20 10 LNG 15 15 — 20 20 20 Fishing Vessel® 015-20 015-10 04 30 30 30 "Can be larger ata single mooring buoy or single-point mooring systems, "Movement away from the berth, “Allowance movements dependent on ramp design. "Depends on cargo handling equipment. 3-5 on maximum breaking strength when specifying the working load. ‘Monitoring of the mooring lines forces to keep them constant is the latest technologi- cal tool used for improving operations at a berth. This technology is basically used for handling of large ships, particularly at loca- tions exposed to waves, wind, and current. ‘The system triggers alarms when loads ex- ceed a setup limit, which permits initiation of necessary action, such as early line tend- ing or use of a tug to hold the ship. Finally, if the berth structure comprises of more than two breasting dolphins, then possible misalignment of dolphins (A) and/or other factors (e.., eccentric load ap- plication) could result in an unequal load distribution between individual dolphins. (Fig. 3-14). To account for this, it is recom- mended that the design load generated by a ship against individual dolphin be in- creased by a factor of 1.3. Additionally, if these dolphins are of different construction, then the total component of mooring force normal to berthing line, P,,), must be re- distributed among the dolphins according to Figure 3-14. Berthing forces due to misalignment of dolphins: 1—ship; 2—breasting dolphin; 8—loading /unloading platform; 4—berthing line, their respective stiffnesses. Therefore, con- sidering the possibility of dolphin misalign- ment, the normal component of mooring load exerted on an individual dolphin, Pc, can be given by Pay) = 1.32 (BD) “4 08 Feng ED! (3-29) where X(ED, ~ total stiffness of the system of breast ing dolphins in direction normal to the berthing line (ED, = stiffness of the particular breasting dolphin within the system in direction normal to the berthing line Essentially, if all dolphins are of the same construction, then (3-30) where n is the number of breasting dolphins. For the case in question, the horizontal component of a total mooring force P,y), is transmitted to dolphins via mooring lines and through friction forces between the ship and the fender system. 3.4 LOADS FROM CARGO HANDLING AND HAULING EQUIPMENT AND UNIFORM DISTRIBUTED LOADS 3.4.1 General Considerations Historically, during the era of sailing ships, vessels were loaded and unloaded manu- ally. With the introduction of steamships, the average weight of a general cargo unit Design Loads 267 handled was increased from 20 kg to a few hundred kilograms; the cargo was trans- ferred with the help of powered winches, which enabled lifting of larger and heavier units, In the 1950s, palletization and pres- linging came into common use; the aim was to standardize units of up to 2000 kg. A pallet is essentially a wooden or metal base on which the cargo is placed. The cargo is typically secured with different kinds of banding. The pallet is transported by fork- lift vehicles and may have provision for being lifted by sling. Pallets are generally loaded to a maximum height of about 2.5-8.0 m and may be stacked up to about 5.0m. In the 1960s, the standardization process took a step further through containeriza- tion and refinement of other systems for products requiring protective containers (eg, liquid and bulk cargoes). Productivity, safety, and the working en- vironment have been improved with the de- velopment of the innovative cargo handling methods. The present trend in industrial shipping is the use of the roll-on/roll-off (Ro/Ro) systems, which under certain con- ditions may be several times as efficient as conventional systems. The chief merits of a Ro/Ro system is greater flexibility toward various cargoes, no demand for expensive port installations, and high utilization of all port and stevedoring resources. However, conventional lift-on /lift-off operation is very often found less costly for high-density cargo. Naturally, developments which occurred in ship construction and cargo handling and hauling technology inevitably affected de- sign and utilization of terminal marine structures. In modern practice, port related marine structures are typically designed to support relatively heavy loads imposed on them by cargo handling and hauling equip- ‘ment such as fixed cranes, rail-mounted and mobile cranes, vehicular loads, railroad track loads, miscellaneous material hauling 268 Design Loads equipment, and miscellaneous uniform dis- tributed (surcharge) loads. ‘In some cases, particularly at dedicated terminals, the structures are designed for very specific loads. Depending on structure usage, the formulation of possible design loads ‘is practically limitless. During the useful life of a terminal the kind of vessels calling at the terminal may suddenly change, with a subsequent change in the required quay edge cargo handling equip- ment. Therefore, in the formulation of the dock design loads, unless the dock is de- signed to serve a highly specialized kind of vessel, the designer must exercise a certain flexibility to strike a balance between a cost-prohibitive structure and a structure that becomes repeatedly unusable or even obsolete because of an inability to support unanticipated loads. At present, various trends can be observed with regard to the selection of quay edge handling equipment. For example, if traffic is sufficiently homo- geneous, dense, substantial, and stable over time, then highly specialized cargo han- dling equipment is used, whereas in other cases the port authorities prefer multipur- pose, more versatile cargo handling equip- ment. ‘The recent changes in quay edge equip- ment are a function of a dramatic evolution of ships. The ship specialization and their increase in sizes conditioned the need of ‘quay edge cargo handling equipment. Also, the loads to be handled in ports are getting heavier, as the ships’ beam increase. Hence, the lifting capacity of cranes and reaches of their booms are also modified. ‘The general trend concerning multipur- pose rail-mounted and mobile cranes ob- served in modern ports is as follows (PIANC, 1987): © Minimum reach of a crane boom, mea- sured from the berthing line of 35 m. + 850 KN of lifting capacity under hook, or 280 kN under spreader + Lifting velocity of 1-15 m/s ‘The container cranes most commonly used feature the following characteristics: ‘+ Minimum height between the lower part of the spreader and quay level of 30 m. * Minimum outreach, measured from the face of the quay, 35 m, and back reach of bm. ‘+ Rail gauge between 16 and 35 m; a rail gauge of 30.5 m is most common. ‘+ A clearance of at least 16 m is provided between the legs in order to leave room for containers and cargo hatches. ‘+ At Teast 400 KN of lifting capacity under the heaviest spreader. In general, container throughput averages 20 containers per hour per crane. According to PIANC (1987) there is a tendency to anticipate heavier containers and larger ships, which may result in required con- tainer crane lifting capacity under the spreader of 500 KN and boom outreach, measured from the face of the quay, of 40 m. These cranes will also need a larger clearance between the legs to allow for pas- sage of the larger containers. The required space under the crane is determined by the horizontal cargo handling operation, result- ing in a certain number of traffic lanes. Probably the latest development in ship- ping technology is the use of specific types of Ro/Ro ships with stern, side, and/or quarter ramps, the weight of which can be quite substantial for some. Because one end of these ramps is supported on the quay structure, the resulting load must be con- sidered in the design of the structure. Ships provided with a ramp can be ‘moored at any location on the quay; there- fore, the whole structure must be designed to support the ramp loads. If operation of Ro/Ro vessels with ramps is foreseen, then all edge bollards, and any kind of surface rails must be located below the deck surface level (Fig. 3-15). The following is a guideline for typical loads which are to be considered in the Figure 2-15, Bdge bollad installed below grade at Ro/Ro terminal design of a modern marine facility. Tt must be understood, however, that each port is different and, therefore, recommended gen- eralized design loads provided herein must be treated with caution. Design Loads 269 3.4.2 Design Load Assumptions ‘Two concepts are presently employed in the formulation of design loads acting on ma- rine structures; one is a “real-life” load as- sumption which is based on the use of mis- cellaneous loads which are falling in a line or concentrated load category (eg., crane, outrigger pads, gantry, and road traffic loads) (Fig. 3-16a); the second concept is based on an equivalent uniform load as- sumption (Fig. 3-16b). ‘The equivalent uni- form load in many cases is a convenient simple assumption; it is easily recognized and understood and is used in lieu of “real” concentrated and line loads. Pallets and containers are typically considered as uni- form distributed loads. ms Figure 3. ‘Terminal design load assumptions: a) real loads; (b) equivalent uniform distributed load; (€) compromise between aseumptiona (a) and (); (@) example of loading case (o)—general view. 270 Design Loads In general, at cargo berths the equiva- lent uniform load is usually assumed as being placed within the apron area (area between berthing line and transit sheds, or the open storage area). It is appropriate to note here that in current practice 20 m is considered as the minimum acceptable apron width for the general cargo berth; however, a wider apron (up to 40 m) is usually more desirable because it provides for more flexibility in cargo handling and hauling operations. In some practical cases, however, use of the equivalent uniform load concept may be misleading. As is pointed out by Junius (1983), in some instances the gross load intensity of stacked pallets may only amount to 20-25 KN/m?, but due to the specific configuration of the pallet, the real local concentrated load may be several times as much. Therefore, depending on the rela- tive configuration of the berth structure, serious overstressing may occur locally even though the design equivalent uniform load rating of the dock is not exceeded. In prac- tice, a compromise is recommended in the combination of both concepts (Fig. 3-16¢). Because the quay edge area is usually the most sensitive to closely situated con- centrated (point) loads, all kinds of related Toads, such as rail-mounted and mobile crane outrigger loads, and the like, must be considered, provided that the equivalent uniform load case is less critical than the concentrated load(s). Tt must be noted that in the design of backfilled structures, the design concen- trated load must be considered as dis- tributed in all directions in relation to the berthing line. This may significantly reduce the effects of heavy concentrated loads on backfilled structures On the other hand, this must not be expected in the case of open-type berth structures with conventional concrete (or other) deck structure supported on piles. This type of berth structure, as well as approach trestles of similar construction, should be designed for both a maximum design uniform live load and a maximum concentrated live load. The latter usually has a very noticeable effect on the deck slab, somewhat less pronounced effects on the supporting beam system, and usually insignificant effects on the foundation. This is because each successive structural ele- ment is affected more by the uniform load than by the concentrated loads, the effect of which is successively diminished. In general, regular concentrated loads from permanently installed and mobile cargo handling and hauling equipment, such as cranes of different kinds, fork lifts, strad- dle carriers, tracks, and others, must. be accounted for in the dock structure design Occasional usage of special heavy lift equip- ment such as large mobile cranes and the like, must be carefully evaluated. Because the load from such equipment can substan- tially increase the cost of the structure, it can be treated as an extreme load, there- fore, in such cases, the structural members designed for normal loads should be checked for larger allowable stresses and smaller factors of safety. In practice, loads from seldom used heavy lifts are spread out by different means (eg., steel beams, mats and others). Gurevich (1969) reviewed available re- search data on loads produced by 45 portal revolving cranes, used in different combina- tions (total 350), with railway track loads placed under the portals of these cranes. It has been concluded that, in most eases, por- tal revolving cranes, in combination with a railway load, produced an equivalent load of about 35 KN/m?; some 150-KN lift capac- ity cranes occasionally produced an equiva- lent load of about 40 kN/m?. In general, Gurevich recommends the use of an equiva. Jent uniform apron load equal to 40 KN/m? for ports operating conventional portal cranes in combination with other loads (e.., railway track and vehicular loads); he also recommends @ uniform equivalent load of 20 KN/m® to be used in general cargo berth design if conventional mobile and /or crawler cranes are employed for cargo handling, Again, it must be stressed that, in all cases, the effect of any kind of heavy con- centrated load on the local strength of a quay wall, or pile-supported pier structure, must be carefully evaluated. 3.4.3 Uniform Distributed Cargo Loads and Miscellaneous Live Loads A discussion on uniform distributed equiva- lent loads (q.) and conditions of load appli- cation have been provided in the previous section. Additionally, it must be said that this kind of load is typically used within the width of dock apron () only. Theoretically, an unlimited surcharge load (g,) may be applied beyond the apron because it does not affect strength and stability of the berth structure. In practice, however, this load is, limited by the very nature of the stored cargo (e¢., liquid or bulk material, general cargo piled on pallets, or in containers, etc.) and by the allowable bearing pressure on the storage area foundation material. For example, bulk materials such as sand, stone, coal, and ore can be stockpiled to produce surcharge loads up to 100-300 KN /m?. The use of retained walls can further increase ‘Table 3-10. Characteristics of standard container Design Loads 271 the surcharge load of stockpiled bulk mate- rial. In the case of bulk material, stockpil- ing the load diagram as depicted in Figure 8-16c should be modified to that as shown in Figure 3-17. ‘Typically, palletized general cargos nor- mally would produce a uniform load of around 20-30 kN/m?, and general cargo shipped in containers may produce a load up to 45 KN/m?, General data on standard containers are found in Table 3-10. In this, table, the basic characteristics including the weight of internationally used containers are given according to ISO (International Standards Organization) Standard 104. However, the actual weight of containers is, usually 50-75% less than the maximum values given in Table 3-10 and, in practice, 2high stacked and 4-high stacked contain- ers produce respectively 20 and 30 kN /m? of equivalent uniform load. Figure 3-17, Typical load diagram at bulk material hhandling terminal. ‘Maximum Theoretical Value of Uniformly Weight ot Distributed Load (kN/m*) Dimensions (mm) Max Container (KN) ‘Number of Layers ‘Types Tength Width ‘Height Weight’ Aluminum Steel 1 2 8 1A 12,190,485 ‘300 230 420 101-202 503, 1B 8125 2.495 250 180 325° «112224836 1c 60552486 200 130 230 136 «272408 1D 2990 -2.495 100 6 107A ALL 1 1965 2485 0 - = 46 292 438 F460 2.495 50 = = i 82a 272 Design Loads ‘Thoresen (1988) recommends that aprons and ramps for container traffic should be designed for a uniform live load of at least 40 KN/m? Multipurpose Ro/Ro cargo in modern port design is usually accounted for 30-50 KN/m? uniform load. Some repre- sentative ranges of uniform distributed loads, obtained from miscellaneous sources and recommended for design of marine structures, are given in Table 3-11. Loads in transit sheds depend to a great extent on the height to which palletized cargo are stacked. Typically, this load may vary between 20 and 50 KN/m? depending ‘on the type of cargo. If there is any possibility that the cargo may occasionally be stored in nondedicated storage areas, then it is reommended the same uniform distributed load be applied to the whole terminal area, including the apron. Such an approach will ensure safety of terminal operation and provide for m mum flexibility in cargo handling and haul- ing techniques. If appreciable quantities of cargo are likely to accumulate on the deck of rela- tively narrow piers (20-50 m wide), then the actual uniform load (but not more than 20 KN /m®), in combination with heavy con- centrated loads (if any), should be used in the pier’s design. Passenger piers, such as commuter boat piers and those, accessible ‘Table 3-11. Uniform distributed loads used ‘in modern port design Load kN /m? Passenger cars and ight trucks 5 ‘Care and trucks ofall kinds 10 General cargo 20 Paper 25-55 Forest products 0 ‘Steel products 40-100 Coal, sand 300-200 Ore 100-300 ‘Cement in bags 15-20 Fertilizers 5 Perishable produce and rain in bags 15-80 to pedestrian use only, are normally de- signed for a minimum uniform live load of 5 KN/m*. If piers are open to automobile traffic, then in addition to the uniform loads given in Table 3-11, the loads recom- mended by the latest issues of the National Bridge Design Codes must be taken into account, ‘A uniform live load of 10-15 kN/m? on the deck area is used for design of oil docks or similar docks, which handle bulk materi- als by conveyor or pipeline, and for which general cargo is of secondary importance. Liquid bulk cargos are handled by pipelines which exert local concentrated loads at pipe supports; bulk cargos trans- ported by conveyers usually exert live loads transmitted to the supporting structure through the conveyer's frame. Saturated weights should be considered for uncovered bulk materials; NAVFAC (1980) covers dock loads for navy piers and wharves; loads for ship repair and maintenance facilities are discussed in Mazurkiewicz (1995). 3.4.4 Rubber Tire and Crawler Track Mounted Equipment Contemporary vehicular hauling equipment such as mobile cranes, forklift trucks, strad- dle carriers, heavy trucks, and so on impose heavy concentrated loads on a marine structure, In fact, these loads may govern the design of the pile-supported dock sys- tem or apron and/or storage area pave- ment. Forklift trucks (Fig. 3-18) are probably the most common cargo handling equip- ment used in ports. Normally they have a lifting capacity in the range 30-100 kN and up to 420 KN, with a typical wheel load of 118 KN for 200-kN-capacity lift trucks and 215 KN for 420-KN-capacity lift trucks. Forklift trucks typically have a turning ra- dius in a range of about 4.0-6.0 m. Usually, Design Loads 278 Figure 3-18. Examples of forklift tracks, 214 Design Loads general cargo terminals and specialized straddle carriers used in container terminal piers are designed for 200-KN-capacity fork- operations (Fig. 3-19) have a lifting capac- lift trucks, However, heavier trucks must ity of 450 kN with a maximum correspond- also be considered where applicable. ing load per wheel equal to 170 and 160 KN, ‘Typically, 3-high and 4-high stacking respectively. Because forklifts, straddle car- Figure 3-19. Examples of typical 3-and 4-high stacking straddle carriers, riers, and other rubber tire mounted equip- ment of different capacities are fitted with a number of wheels proportional to their lift- ing capacities, the usual maximum contact pressure remains below 1000 KN/m?. Normally, terminal facilities are de- signed to accommodate a standard highway load. In designing for a vehicular highway Toad, it is typically assumed that a 20-t motor truck (traffic load class H20-44 or HS20-44) could, within reason, be placed anywhere on the dock’s deck, and load dis- tribution assumptions should comply with bridge construction regulations. However, the impact coefficients normally used for bridge design may be decreased by half, assuming reduced speed of vehicular traffic on the deck of the dock and access trestle (bridge). ‘The vehicular load in pier design basi- cally consists of the individual wheel, axle, and truck loads, as given by the appropriate specification. In the design of a pier deck, the uniformly distributed live load rarely governs because of the relatively short spans of the deck structural members. ‘The critical wheel load on the deck elements depends on the wheel contact area. In the case of the deck wearing surface, the di- rectly loaded area of the deck structure is somewhat larger than the contact area be- tween the tires and the deck riding surface. ‘This is due to the effect of the load dis- tributing action of the wearing layer. For a concrete/asphalt wearing surface, a 45° distribution is usually assumed. For design Design Loads 278 purposes, the loaded area usually is repre- sented as the equivalent rectangular area (ab), which is assumed to be uniformly loaded (Fig. 3-20). ‘The standard truck load typically used in North American bridge design practice is shown in Figure 3-21. The average actual dimensions of the standard North American truck tires (corresponding to approximately 36.5-, 54.5-, and 73-kN wheel loads) and the assumed dimensions of the rectangular wheel-load areas at the deck surface (ap- proximately the actual tire pattern dimen- sions) are given in Table 3-12. In addition to vertical loads, all moving rubber tire mounted vehicular equipment produces lon- gitudinal forces due to traction and braking. ‘To account of it, 10% of the vertical wheel load is usually applied longitudinally. Ap- plicable design codes such as AASHTO (1982), CAN /SA S6-M87, or similar speci- fications should be used as a design guide for defining the general vehicular loads. ‘Mobile cranes are a common cargo han- dling equipment in ports. They may be rub- ber tire or crawler track mounted. Lifting capacity of this type of equipment ranges from 30 to 100 KN and up to 3000 KN; note that to handle loaded containers, the crane must have at least 400 kN lifting capacity. Rubber tire mounted mobile cranes trans- mit their corner loads via outriggers to the pier deck structure or to the apron pave- ment of backfilled dock structures. In gen- eral, the wheels of mobile cranes are not used to transmit loads during hoisting oper- ©) | Lil hee 2 Figure 3-20, Wheel load 276 Design Loads ©) ©) ©) © 2} elo —£ Figure 3-21. Standard truck load. V is the variable spac- ing, 43-922 m inclusive. Spacing to be used is one which produces maximum stresses. For det bridge design code. ations. The crane’s outriggers are typically dimensioned in a way that the bearing pressure against the terminal pavement amounts at a maximum of about 400 KN/m?. This load may be too heavy for some older structures. In this case, the load must be properly distributed to the deck structure by means of mats or stee! beams. ‘The bearing pressure may also be reduced by enlargement of outrigger bearing area. ‘see relevant, This, however, is done at the expense of the crane’s mobility and, therefore, should be avoided where possible. Unless otherwise required, new general berthing piers, or quays should be designed to accommodate a minimum 700-KN- capacity mobile crane (NAVFAC, 1980). Using this minimum crane size for design, the maximum outrigger corner load may be on the order of 700 KN. Additionally, Design Loads 277 ‘Table 3-12. Truck wheel dimensions and loaded area (mm) Dimensions Used in Design Nominal ‘Top of Wreet Wearing ‘Top of Toads Wheel Dimensions Surface Deck Structure any T T ny SW oa @ 3 365 17s 226~«40x107—«200~=CGT==SCORSCS SCC 545. zl 24 BEX 10? 825526. 508545905 m0 24 19 Tx 1094126 Bk 0] 05 |. Mobile crane maximum loads Wheel Load* 0) 160 120 10 150 40 Outrigger Load aa 180) 300 3300 2400 1680 Source: Prom PIANC (1987) cranes may cause horizontal forces due to wind, and braking due to movement or during rotation. As a rule of thumb, this force is usually taken as equal to about 10% of the wheel load and can be applied in any direction, Some representative values for mobile crane wheel and corner loads are found in Table 3-13. Typical wheel and outrigger arrangements for mobile cranes with 250 and 450 KN capacities are de- picted in Figure 3-22. Other types of rubber tire mounted equipment include different types of tractors, with or without trailers, front-end loaders, automobiles, Ro/Ro trailers, and others. Some of the above vehicles may produce substantial wheel loads and should be considered in the structure's design; others, such as automo- biles, produce relatively light loads. As it, was pointed out earlier, piers designed to handle Ro/Ro and heavy truck traffic must be able to withstand substantial localized loads. Regular loads, obtained from the equipment manufacturer's literature, must be used with consideration of the dynamic nature of these loads. PIANC (1987) recom- mends that static wheel loads for mobile cranes and for forklift trucks be increased by 20% and 40%, respectively. Crawler track mounted mobile equip- ment (mostly cranes and bulldozers) are also used in ports. The load distribution under the crane’s tracks is typically trape- zoidal (nearly triangular) and the bearing pressure under the tracks may peak up to 700 KN/m?. Tread widths are typically within a range of 85-140 em. Crawler track ‘mounted equipment is usually used on un- paved surfaces of port territory. Otherwise, timber mats must be used to protect the pavement or concrete deck surface. Mats are also recommended for better load distri- bution where required. In addition to the rated load, an impact factor of about 20% is usually applied to account for the dynamic effect of the crane’s performance. Generally, specific data on the mobile cargo handling and hauling equipment are obtained from the manufacturer’s litera- ture. However, the manufacturer's general brochures can sometime be dangerously misleading in regard to deck load, warns Gaythwaite (1990), because the listed aver- age tread pressure for crawler-type cranes may be inadequate for deck structure de sign, Therefore, if in doubt, the designer should carefully verify any specific equip- ment characteristics with the manufac- turer. In conclusion, it should be noted that to prevent mobile equipment from rolling over the berth edge and into the water, a safety 278 Design Loads » t Inco |! leet Lt noo | t 3.00 © Figure 3-22, Examples of typical wheel and outrigger arrangements for sobile cranes, 250 kN (a) and 450 KN (b) capacities. curb 20-30 em high is usually installed along the berthing line. This curb is typi- cally designed to take a horizontal impact force of 15-25 kN, depending on the type of mobile equipment used. 3.4.5 Rail-Mounted Cargo and Material Handling Equipment This type of equipment includes revolving portal and container cranes of miscella- neous capacities and construction, railroad trains, and a variety of specialized bulk handling equipment. Customary general cargo handling cranes used in port operation are basically revolv- ing portal level-lifting cranes spanning one, two, or three railway tracks (6-15 m be- tween legs). Typically, portal crane capacity varies in the range of 30-400 kN at a work- ing radius of 20-40 m. At seaports, the crane’s front rail is typically placed at, 1.75-2.25 m from the quay edge. This is usually done in order to maintain sufficient room for bollards, to accommodate the out- board of a crane leg, to provide working space for line handlers, and for access to the berth. In general, the distance between front rail and quay edge should be maintained as, WA @ Design Loads 279 small as possible; however, care must be taken to avoid any potential collision be- tween the crane and a heeling ship, or the ship's flared bow. In inland ports, portal cranes are usually placed closer to the face of the berth than in seaports (Fig, 3-23). This is usually done for the following reasons: riverboats and barges generally do not have flared bow and therefore cannot hit the crane during the high water level while approaching the dock; they also cannot hit the crane by the stern during departure maneuvers; a closer crane location is required to improve ship observation by the crane operator dur- ing loading/unloading operation under a low water level. Inland port cranes usually have the front legs placed about 1.1-1.3 m from the berth’s face; this provides a minimum 0.8-m walk- way along the structure's edge. In some cases however, and particularly when a crane’s front legs rest on sheeting, or at pier-type structures built on a sloped por- tion of the quay, the walkway for the crew and maintenance personnel is provided be- hind the front leg of the crane. This type of arrangement, in general, is not desirable because of its vulnerability to ship impacts. Tt may also cause the problems associated with crane accessibility for maintenance from the water side and difficulties with the arrangement of mooring hardware and the © ® Figure 8-23. Typical pocitioning of portal cranes in inland ports in relation to the berthing line. 280 Design Loads placement and handling of ship mooring lines. ‘The crane boom can slew through 360°. ‘Therefore, the maximum load per leg that is usually attributed to maximum design wind load imposed on the crane can occur at each corner boom position. Maximum portal crane wheel loads are typically within the range 200-400 KN and are usually obtained from the crane manu- facturer. Some typical values of portal crane wheel loads used in North America are de- picted in Figure 3-24, Again, a dynamic factor of 1.2 is typically used to account for vertical and horizontal inertia forces that ‘occur during lifting or setting down of cargo. A horizontal load of 10% of the wheel load in the rail direction, due to crane traction or centrifugal forces, is also considered. ‘The most unfavorable crane load combi- nation on a berth structure is created by ‘two portal cranes operating at the closest distance (typically 2.5 m) between adjacent wheels. Naturally, under such conditions Figure 8-24. Standard North American portal cranes whee! load. two neighboring cranes cannot produce maximum loads in adjacent legs simultane- ously, otherwise the crane’s jibs would in- terfere with each other. ‘The German code EAU (1990) provides guidelines on typical portal crane corner loads. Useful information on rail-mounted crane loading may also be obtained from NAVFAC (1980). ‘A great variety of rail-mounted container cranes of various configurations are now used in ports throughout the world. Most container cranes have lifting capacity in excess of 400 KN. They are typically built as fall portals cranes with cantilever beams and trollies, whose supports have, as a rule, six or eight wheels. The crane gauge is typically in the range 15-18 m. An example ofa typical container crane of 530 KN lifting capacity (inclusive of spreader) commonly in use is shown in Figure 3-25 (EAU, 1990). These cranes are designed to handle the typical 40-ft container as well as 2 x 20-ft containers (so-called twin-twenty method). ‘The maximum corner load on a portal sup- port of either 3000 KN or 8300 KN (see Fig. Design Loads 281 3-25) should be increased by 10-15% due to wind action or inertia effects. ‘Maximum wind forces parallel to the ‘eraneway direction and perpendicular to it ‘on the order of 200-300 kN are applied to typical working cranes, or a load of 600-750 KN applied on idle cranes must be consid- ered, Smaller values of wind force are used in areas with relatively moderate wind con- ditions, and larger values in areas with severe or special wind conditions. If cranes with spans of more than 18 m are used, then maximum corner load, as well as max- imum wind force exerted on the cranes, must be corrected accordingly. For example, some container cranes with an effective span of 30 m may produce a maximum corner load in the order of 3000-3600 kN (EAU, 1990). In addition, the rated capacity, impor- tant crane features are its outreach and backreach. These effects on crane loadings must also be considered. Contemporary container ships are usu- ally built with flared bows which may sub- stantially overhang the berth line. In order aah igure 3-25, Typical container crane of 530 kN lifting capacity. 282 Design Loads to prevent bow collision a crane is usually place 2 m or more from the berth face. A wide variety of specialized rail- mounted equipment is employed in bulk cargo handling terminals. As a rule, this equipment is custom built to serve a certain particular project. It includes traveling gantry cranes (bridge cranes) with miscella- neous grab devices, traveling bucket eleva- tors which utilize various conveyor systems, traveling ship loaders, and others. Their design parameters are provided by the fab- ricators and typically include vertical and horizontal wheel loads due to material han- dling, and dynamic impact and inertia. Most often, maximum design wheel loads are limited to the design loads on a rail section and, similar to the crane corner loads, can be within 350-450 KN. In most cases, this type of equipment has substantial windage area able to attract significant lateral wind force, so that the dynamic response of elevated structures un- der wind loading must also be carefully addressed (Phang, 1977). In general, when nonstandard equipment is employed, it is advisable that the designer work closely with the equipment manufacturer to prop- erly understand the loadings and relevant operating criteria. 3.4.6 Fixed-Base Equipment ‘This may include regular stiff-leg (derrick) cranes (Fig. 3-26), stiff-leg cranes sitting on a ring which can slew the crane through 360° (ringer crane), loaders, such as that depicted in fixed in-place revolving cranes, miscellaneous loading towers, liquid bulk handling arms (Fig. 3-27), and others. ‘The weights of the above equipment are typically in a range 150-500 kN, with some up to 15,000 KN, and in special cases, even Figure 3-26. Stff-log crane (dere, Figure 3-27. Liquid bulk loading arms, more. The foundation structure for this type of equipment is designed on a basis of load data and the operating criteria provided by ‘the manufacturer, along with associated forces due to effects of environmental loads and berthing and mooring forces as previ- ously discussed in this chapter. In addition to normal service loads, an appropriate impact factor is used. Descrip- tion of some unigue installations of this ‘kind can be found in special publications by ASCE, PIANC, The Dock and Harbour Authority, and others. 3.5 SHIP IMPACT (BY M. SHIONO AND G. TSINKER) Ship impact force on berthing structures can be determined by statistical, empirical, or theoretical methods, or by physical and/or mathematical modeling. ‘The statistical method is based on in- formation obtained from prototype projects with similar port characteristics, such as exposure to environmental forces, mode of operation, type of vessels, and others. How- ever, one should realize that the data obtained from actual measurements of berthing energies cannot be applied auto- matically to new port design. For example, data based on measurements on ships ar- riving fully loaded cannot be used for load- ing terminals where ships will arrive in ballast. In the first case, the current could be a dominant factor in ship berthing en- ergy, and in the latter, the wind load may be more dominant than current force. It should be noted that actual berthing- energy measurements usually show a significant scatter, which can only be de- scribed statistically, Furthermore, the statistical approach implies that, regardless of which design value of berthing energy is chosen there is always a risk that it can be exceeded. Also, ‘the more berthing operations the facility (fender) is exposed to, the larger this risk becomes 284 Design Loads Based on field observation data and risk analysis, Dent and Saurin (1969) suggest the following design values for tanker ter- minals located at exposed-to-wind and waves locations: 1. 2.6 meter-tonnes (18.8 feet-kips) per 1000 DWT of design ship at yield stress in the fonders 2 185 meter-tonnes (11.2 feet-kips) per 1000 DWT as a normal maximum related to approximately maximum working stress in breasting structures and fenders For berths located in protected harbors and for loading terminals where vessels nor- mally berth in ballast, five-eighths of these values are recommended. More details on statistical methods for the determination of a ship berthing energy is found in works by Balfour et al. (1980), Piaseckyj (197, undated in 1982), and in PIANC (1984). In the empirical method, the following parameters are usually taken into account: 1. Berthing frequency of the design vessel 2. Exposure conditions 3. Number and maneuverability of vessels 4, Borthing method: with or without tug assistance 5. Berth mode of operation; loading or un- loading. Specific values are assigned to the parame- ters based on judgment and experience and are related to the cost of repair against both downtime and the probability of exceeding the fendering energy capacity. Girgrah (1977) has concluded that berthing energy is empirically related to ship displacement (W,) and to the inverse square root of vessel displacement as fol- lows. wy ae(t+m) mae 3) According to Girgrah, Eq. (3-31) may be applied in the design of the end fenders and isolated berthing points where the point of contact is amidship. For interior fenders, the value should be reduce by a factor of 05. Tt should be mentioned that both the statistical and empirical methods are more often recommended for determination of im- pact forces generated by large ships. The theoretical (kinetic) method of calculating the ship berthing energy is more universal than both the statistical and em- pirical methods and is also applicable to a ship of any size. It tends to account for the following factors affecting dissipation of the ship's total berthing energy: Rotation of the ship Elastic deformation of the berth structure Compression of the fendering system Displacement of water between the ship and berth structure Aen ‘Thus, in most cases, the fender system must absorb only a part of the ship's total kinetic energy, which is usually referred to as the “effective berthing energy” of a ship (E). The effective berthing energy may be ob- tained from (+ Wav? 7 (3-32) x0 where W, = displaced weight of the vessel W, = hydrodynamic or added mass V, = component of the vessel's velocity normal to the berth structure 4 = gravitational acceleration Cy = eccentricity factor Ci = ship and/or structure softness factor ‘The sum (W, + W,), which is usually re- ferred to as a ship virtual mass (W), is typically expressed in tonnes. The approxi- mate relationship among the ship's dis- placement tonnage and her gross, net, and deadweight tonnages for a variety of vessels is given in Chapter 2. Hydrodynamic mass, sometimes called “added mass,” is the mass of a water en- trained by a moving vessel as the result of a vessel's motion (Valentine, 1967). While in motion, a ship displaces water and pushes ahead of her a positive pressure field, ap- parent as a raised water level. The dis- placed water passes back along the hull. For a ship moving sideways, much of the returned flow moves under the keel (Fig. 3-28a), unless the underkeel clearance is too small, in which case the flow is forced to return around the stern and bow. In principle, the hydrodynamic mass of a ship moving in open water is approximately 10% of the ship's mass. It increases consid- erably during sideways movement, such as during the berthing process, because the ship's mass is augmented by the large vol- ‘ume of water involved in the transverse motion of the vessel toward the berth. This, volume varies with the following factors: 1. Berthing facility characteristics, such as type of structure (open, solid wall, or Figure 3-28. Water movement end wave propaga- ‘tion during ship side-way movement. [After Ball, (1982) ‘and Middendorp, (1981)] Design Loads 285 semiopen) and spring properties of the fenders 2. Ship characteristics, such as its under- water shape and draft, and the way she approaches the berthing facilities (eg., sideways, parallel to the pier, or straight forward at a certain angle) 3. Berthing velocity and underkeel clear- 4. Oceanographic and environmental fac- tors, such as water depth and possible impact of current, wind, and waves. Some studies indicate that the added mass of many slender structures immersed in fluid is comparable to the mass of fluid displaced by these structures. On this basis, and for simplification, some authorities rec- ommend using the value of hydrodynamic mass W, as equal to the weight of water contained in a cylinder that has a diameter equal to the vessel draft and a length equal to the vessel length. Therefore, Wy = 0.785D"Lp 3-83) where D = vessel draft L = vessel length p= specific weight of water According to Eq. (3-38), the added mass of a fully loaded large tanker, or fully loaded large cargo ship, becomes equal to 0.5-0.6 times the displacement tonnage. Further recommendations to aid in the calculation of the hydrodynamic or added mass are available. The majority of these recommendations describe the virtual mass in terms of its relationship with ship diplacement tonnage. Equation (3-34) presents another means of determining the virtual mass: W=C,Wy (3-34) where C,, is the hydrodynamic mass factor. For the determination of C,,, Grim (1955) proposed Eq. (3~35), which is based on re- 286 Design Loads sults obtained from model tests: 18D “, (3-38) where D = draft (in meters) B= beam of the ship (in meters) Surin (1963) published a comparison on energy absorption by fender systems using theoretical data and data derived from full- scale observations. The observations were based on 70 berthings of ships varying from 16,000 to 50,000 DWT, with various under- keel clearances. Despite the scattered re- sults of these tests, Saurin concluded that value of C,, = 1.3 would represent a rea- sonable average. However, the factual data presented by Saurin suggests that to be on the conservative side, a value of C,, = 18 should be used. ‘Vasco Costa (1964), on the basis of his experiments, suggested some corrections to Grim's formula and recommended the fol- lowing expression for determination of mass factor Cy: 10 2p (3-36) 422 : B Ruppert (1976) suggested modification to formulas proposed by Grim and Vasco Costa, which lead to somewhat lower C,, values. According to Ruppert, C,, can be determined from the following formulation: cn = 09+ 482 z (a-37) On the basis of model tests in five different water depths (H/D = 1.07, 1.14, 1.90, 3.31, and 8.81), Girandet (1966) proposed the fol- lowing equation which links C,, value to depth of water (H): “, 2+0.9{ 6-38) #5) Prior to 1989, the Japanese standard MOT considered formula (3-39) to deter- ‘mine the hydrodynamic mass factor, appar- ently based on a publication by Stelson et al. (1958): 0.25xD*Lp 7 (3-39) Cy = 10+ where D = draught of ship L = overall length of ship p= specific weight of water at berth W, = displacement of ship According to formula (3-89) for fully loaded large tankers and for fully loaded cargo ships, C,, approaches 1.51.6. Blok and Dekker (1979) remarked that expression (3-39) is basically the added mass following from potential theory on sway oscillation for a beam-to-draught ratio of 2 in infinite water depth. They also point out that the value of C,, is greatly influ- enced by fender characteristics such as spring rate and the character of its load /deflection relationship. Fenders with different stiffness charac- teristics by as much as a 1:10 ratio were used by Blok and Dekker in their tests. For the total impulse versus change of momen- tum for all fenders they found a value of C,, = 3.8 (with little scatter). The explana- tion for the larger hydrodynamic mass fac- tor probably lies in the fact that when a light vessel hits a stiff fender, she tends to change her direction of movement very rapidly. The resulting side movement, how- ever, meets resistance from ambient water as well as the momentum of ship move- ment, which prevents the ship from chang- ing her forward movement rapidly. This in- evitably, results in a larger impact load compared to that resulting from a collision between a ship and a soft fender. On the basis of field observations and model studies, Ueda (1981) concluded that the second part of Eq. (3-89) should be increased by a factor of 2 and hence by equal to 0.57DLp/W,. The latter resulted in C,, values in the range 1.6-2.3. For cal- culation of C,, value, Ueda proposed Eq. (8-40) in which he introduced the ship's block coefficient (C;): =D ¢ 2C,B -1+ (3-40) This value of C,, is currently recommended by the latest issue of Japanese Standard MOT. It should be noted that for C, = 0.785, the outcome of Eq. (3.40) is similar to that obtained from Eq. (3-36) proposed by Vasco Costa. From field measurements conducted from 1966 to 1968 in the port of Le Havre (France) at oil tanker berths receiving tankers ranging from 70,000 to 130,000 DWT despite inherent inaccuracies in the observations and calculations and in the variation of D/H ratios, the value of C,, ranged between 1.4 and 1.7, with an aver- age close to 1.5 (PIANC, 1984) ‘The numerous studies conducted in Japan (Motora, 1959; Nagasawa, 195: Heyashi, 1962; Fujino, 1968; Mizoguchi and Nakanayama, 1973; Otani et al., 1974; and others) have produced various values of C,, in a range between 1.9 and 3.6. ‘The value of the hydrodynamic mass is largely dependent on the berth structure geometry. For example, an open piled struc- ture allows water from around the ship to move relatively unimpeded past the piles. On the other hand, a solid-wall structure considerably obstructs the flow, particularly if the underkeel clearance is small. When the ship suddenly stops upon impact with the berthing structure, such as an open piled structure, the flow displacement rate drops suddenly to almost zero, causing the ‘water level to fall on the ship's side nearest to the berth and rise on the other side (Fig. 3-28b). This phenomenon is a result of the momentum of the mass of water under the keel, which maintains an underkeel flow when the ship is stopped, thus creating an Design Loads 287 imbalance between the flow displacement by the ship and the underkeel flow (Mid- dendorp, 1981; Ball, 1982). The difference in water level across the ship produces a hydrodynamic force toward the berthing structure, which is appended to the ship's own mass to give the “virtual mass.” ‘As it was mentioned earlier, when the forward-moving ship collides with a stiff fender at a certain angle, she rapidly changes direction, which is resisted by the ship's mass as well as by the resistance of ambient water. A laboratory investigation of the above mode of ship-dock collision has demonstrated that the magnitude of hydro- dynamic mass depends greatly on the angle between the dock and approaching vessel; accordingly, a smaller hydrodynamic mass is associated with a small approach angle (Goranov, 1980). Contrary to an open berthing structure, when a ship approaches a solidwall struc- ture, a gradual cushioning effect occurs, in- creasing with the decreasing distance from the structure. In general, for a preliminary evaluation of the value of added mass, some authori- ties recommend considering the following factors: 1. 10 for an open structure of any dimen- 2. 08 for a semisolid-wall structure longer than the length of the ship 3. 06 for a solid-wall structure longer than the length of the ship Similar recommendations are found in PIANC (1984). It should be also noted that the C,, value may be affected by currents (Tyrell, 1966/1967). For large ships, unless the de- signer has good reason to adopt other val- ues, Permanent International Association of Navigation Congresses (PIANC, 1984) recommends that the value of hydrody- namic mass (W,) range between 15W, (for very large underkeel clearance, say 0.5 x 288 Design Loads draught) and 1.8W, (for very small under- keel clearance, say 0.1 x draught). Depending ‘on site conditions, fender stiffness, and flexibility of berth structure, C,, values for smaller ships can be much greater than those for large ships. It should be noted, however, that some earlier field measurements of ship berthing velocity show no correlation with vessel size (Dent and Saurin, 1969; Svendsen, 1970). ‘The dock designer must always be aware that in all cases of the ship's direct ap- proach to the berthing structure, the value of added mass should be determined by exercising good judgement and by evaluat- ing all berthing conditions previously dis- cussed. A degree of uncertainty in the mechanism of ship-dock interaction does present a problem in determining the value of the added mass, which, apparent in the above mathematical equations, probably cannot be solved with mathematical preci- sion without further research. The above speculations on a part of the hydrodynamic mass factor, C,,, indicate that it has been and still continues to be a matter of debate among practitioners. Despite various methods that have been suggested to determine the value of the added mass, most practitioners in the field still rely heavily on past practice, good Judgment, and experience. Approach velocity and angle are very im- portant components to be considered in the evaluation of a ship’s berthing energy. As seen from Eq. (3-32), the approach velocity is an important parameter in determining energy to be absorbed by the fender system. The designer must always realize that the ship's approach velocity is almost entirely dependent on the size of a ship, mode of approach (eg, assisted by tugs or self- propelled), human factors, environmental conditions, particulars of the site, and the skill of local pilots. A wrong maneuver of the ship due to human error or a sudden change in weather conditions can results in much larger approach velocity and, hence, much larger berthing energies to be ab- sorbed by the fender system. Therefore, the ship's approach velocity data should not be relied on exclusively but used as a starting point in fender system design. A discussion ‘with local pilots who are going to use the berthing facility is also recommended. ‘The latest technology, such as sonar sys- tems (Fig. 3-29), provides dock and ship ‘operators with a means to keep the berthing vessel under better control while docking. Docking sonar systems assist in safe and controlled berthing operations by providing ‘an accurate and highly stable measurement of a vessel's approach speed and distance from a berthing structure. Distance and speed are measured at the bow and stern independently and are instantly available to both the pilot on board and the dock personnel throughout the berthing opera- tion. The data are also printed out for sub- sequent analysis, Docking sonar systems have been used frequently at oil and LNG terminals which typically receive large tankers. However, because of the relatively high cost of this technology it is seldom used in dry bulk and general cargo ports. One example of sonar system use at a bulk loading facility is Ridley coal loading terminal, Price Ruppert, Canada (Port Development Inter- national, February 1985). ‘The proper selection of ship speed is very important in arriving at a realistic value of kinetic energy to be used in the design of the fender system and subsequently the berthing structure. As a general recommen- dation, the following berthing velocities can be assumed for the design of a berthing structure, 1. For tug-assisted large ships ‘+ very favorable conditions, 10 em/s + in most cases, 15 em/s + in very unfavorable conditions with eross-currents and/or much wind, 25 cm/s 2. For any ship approaching the dock with no tug assistance, 20-25 em/s Design Loads 289 Figure 3-29, Docking sonar system: (a) In the first case, the design of the terminal is usually based on an approach angle of 10°. In all other cases, an approach angle of 10°20" is used based on local approach and environmental conditions. The larger ap- proach angles are more characteristic for inland ports, and smaller angles are most, often in seaports receiving large ships. ‘The likelihood of exceeding these design berthing velocities is great; hence, caution should be exercised in selecting ship design approach velocities and subsequently fender system design. For more information on berthing velocities, the reader is referred to PIANC (1984). general vow; (b) and (@ display board. The energy to be absorbed by the dock fender system is usually taken as equal to one-half of the ship's kinetic energy, based on Cy = 0.5. The remaining half is assumed to be absorbed by the ship and water due to ship's rotation about her contact point with the fender. This point is commonly referred to as a “quarter point” (Fig. 3-30). For other berthing points, the value of Cy can be obtained from Eq. (3-41) or Fig. 3-31. 1 -—_, 3 1+ ow Cy 290 Design Loads where = distance from the gravity center of a ship to the point of contact, measured along the face of the pier (Fig. 3-30) r= radius of gyration of a gravity conter on the horizontal plane, commonly, r= 0.25L, where L is the full length of the ship. Figure 3-31. Derthing point. Beventriity factor versus vessel When the vessel is brought to rest along- side of a stiff marginal wharf structure, it is assumed that the impact energy is primar- ily absorbed by the fender system only. Fur- thermore, it is usually considered that the potential deflection of any type of continu- ous berth structure cannot contribute to ab- sorption of the energy, as the time it takes for the structure to react is much longer than the time over which the impact is applied. To calculate the impact force re- sisted by the continuous, or otherwise very stiff, structure it is necessary to consult load-deflection curves provided by the fender manufacturer. In contrast, flexible breasting structures (eg, flexible dolphins and similar struc- tures) greatly contribute to the absorption of ship berthing energy (Fig. 3-32). As men- tioned earlier in this section, rubber tires are used often as fenders in small-craft har- bors, With varying degrees of success, they are also used to absorb berthing energy from larger ships up to 3000-5000 DWT. ‘Load-deflection curves for used tires, or other fender systems which include used tires (eg,, tires filled with a closed-cell foam (Tsinker, 1986), big tires stuffed with smaller ones, and others), can be obtained experimentally. As reported by Hendrick (1980), used rubber tires and particularly those of large Tetlection Figure 3-32. Energy absorption by combination of flexible dolphin and fender: (a flexible dolphin fender system; (b) system reaction versus deflection. diameter, can absorb substantial amounts of energy, especially when assembled in groups of three and more tires. On the basis, of rather comprehensive tests that have been carried out at the Building Research Structures Laboratory of the Canadian National Research Council, Hendrick has arrived at the following basic conclusions: 1. The condition of the tire greatly influ: ‘ences the ultimate load on it; factors to be considered are wear, amount of tread re- ‘maining, whether or not the tire has been retreaded, cuts (size and location) and bruises. 2. ‘The tire construction and type or tread design may influence the energy absorb- tion capacity and the resulting reaction force, Performance of used tires as a fender can be enhanced greatly by filling them with a closed-cell resilient foam (Tsinker, 1986). In experimental work, the F = f(d,) curve is obtained first. Here F is the force applied to the fender and dy is the corre- sponding fender deflection (Fig. 3-33). The area between curve F = f(d,) and the axis d, represents the work (energy) done by the fender. Hence, the curve E = f(d;), which is an integral curve in the relationship to the curve F = f(d,), can be plotted by inte- grating the appropriate areas. ‘The design berthing energy of a ship is used for selection of a desired energy ab- sorbtion (fender) system and ultimately for T Il 1 | I Figure $-88. Typical energy versus deflection dia- gram obtained from fender test. Design Loads 291 determination of the ship impact load that will act on a marine structure. Practically, it is done by consulting the reaction force /energy absorbtion in fender deflection curves, given in the fender manufacturer's catalog. Normally, these curves are ob- tained from the fender load tests that are carried out in accordance with recommen- dations established by PIANC (1984). Ac- cording to these recommendations, in order to obtain the aforementioned curves, the fenders are tested at relatively low com- pression speeds. This, according to Kiedaisch (1995), results in the unneces- sary redundancy (“reserve of energy absorb- tion capacity”) and therefore in fender overdesign. By the time this book is pub- lished at least one fender manufacturer is using the so-called “velocity correction fac- tor” to account for a ship actual berthing velocity. This resulted in smaller more eco- nomical fender units recommended by this manufacturer for dock fendering. Padron (1994) disagrees with this practice. He ar- gues that the velocity correction factor, as currently applied, is based on the assump- tion that the fender unit is compressed at a constant velocity throughout the impact ey- cle, while in practice this velocity varies from the initial impact velocity to zero as the vessel comes to a stop. In the process the fender unit deffects at a continuously decreasing rate. Padron pointed out: that use of the velocity correction factor does have validity with respect to the maximum reaction force that the fender unit imposes on the structure because it is important to the designer to be fully aware of the peak reaction force for which the structure must be designed. Padron further suggests that normally the velocity correction factor should be applied when considering the magnitude of the reaction force that. the fender unit can be expected to exert on the structure, and not applied to fender energy absorbtion capacity because, as presently used, the velocity correction fac-

You might also like