Sei sulla pagina 1di 26

Coastal Engineering 37 1999. 149174 www.elsevier.

comrlocatercoastaleng

Wave loads on rubble mound breakwater crown walls


Francisco L. Martin
a

a, )

, Miguel A. Losada b, Raul Medina

Ocean and Coastal Research Group, Uniersidad de Cantabria, Ada. de los Castros s r n, 39005 Santander, Spain b Uniersidad de Granada, ETSI de Caminos, C.y.P, Campus de la Cartuja s r n, 18071 Granada, Spain Received 10 March 1997; received in revised form 11 February 1999; accepted 18 February 1999

Abstract Crown walls are primarily built to reduce wave overtopping of mound breakwaters. Several methods have been proposed to calculate wave loads on the crown wall, e.g., Iribarren and Nogales wIribarren, R., Nogales, C., 1964. Obras Martimas. Dossat Ed.., Madrid, 376 pp.x, Jensen wJensen, O.J., 1984. A Monograph on Rubble Mound Breakwaters. Danish Hydraulic wGunbak, Institutex and Gunbak and Gokce A.R., Gokce, T., 1984. Wave screen stability of rubble-mound breakwaters. International Symposium of Maritime Structures in the Mediterranean Sea. Athens, Greece, pp. 2.992.112x. In this paper, a new method based on those previous results, and on further experimental work, using monochromatic waves, is presented. The application of the new method requires waves breaking on the armour layer; i.e., only broken waves will reach the crown wall. The method is extended to irregular waves via the hypothesis of equivalence introduced by Saville wSaville, T., 1962. An approximation of the wave run-up frequency distribution. Proc. 8th International Conference on Coastal Engineering, Mexico Cityx and is applied to the crown walls of Gijon and Bilbao breakwaters in Spain. The comparison of the probability force distributions obtained by the present method to that measured by Burcharth et al. wBurcharth, H.F., Frigaard, P., Berenguer, J.M., Gonzalez, B., Uzcanga, J., Villanueva, J., 1995. Design of the Ciervana breakwater, Bilbao. In: T. Telford Ed.., Proc. 4th Coastal Structures and Breakwaters, Chap. 3. Institution of Civil Engineersx and Jensen 1984. is relatively good. q 1999 Elsevier Science B.V. All rights reserved.
Keywords: Wave loads; Crown walls; Mound breakwaters

Corresponding author. Fax: q34-42-20-18-60; E-mail: martin@puer.unican.es

0378-3839r99r$ - see front matter q 1999 Elsevier Science B.V. All rights reserved. PII: S 0 3 7 8 - 3 8 3 9 9 9 . 0 0 0 1 9 - 8

150

F.L. Martin et al. r Coastal Engineering 37 (1999) 149174

1. Introduction In Mediterranean countries, mound breakwaters are often built with a concrete parapet resting on the mound layer, and being partially protected by the armour layer. In engineering practice, this parapet is known as a crown wall, wave wall, wave screen, etc. Although the primary function of the crown wall is to reduce wave overtopping, there are several reasons for topping the breakwater with a crown wall, e.g., i. protection of breakwater rear slope if the breakwater is overtopped, ii. facilitation of some construction procedures, and iii. reduction of required volume of quarry material and thus reduction of construction costs, etc. There are a few methods for the calculation of wave forces on crown walls: Iribarren and Nogales 1964., Jensen 1984. and Gunbak and Gokce 1984. are some. However, it is known that the first method is pessimistic, yielding conservative design. The second is not reliable since the influence of wave period is not represented adequately, the influence of the armour geometry in reducing wave loading has not been addressed and, therefore, calculated wave forces deviate from measurements up to "30% Bradbury et al., 1988.. Moreover, Pedersen and Burcharth 1992. tried to verify Jensens parameterisation by using experimental measurements from different authors finding a large scatter in the results. The third method is difficult to apply for design purposes. In this paper a new semi-empirical method, based on these previous investigations and on additional experiments using monochromatic waves, is proposed. First, the crown wall problem is discussed from a design point of view. Next, the formulation of the wave pressure on a vertical wall induced by broken waves is presented. After the introduction of the experimental results, the new method is extended to irregular waves via the hypothesis of equivalence introduced by Saville 1962. and 1978. for run-up on rough permeable slopes. empirically proven by Bruun and Gunbak Finally, the method is applied to actual breakwaters and the results are compared to empirical data from Burcharth et al. 1995. and Jensen 1984..

2. Definition of the problem 2.1. The crown wall problem In this section the main factors involved in the design of a crown wall are discussed. Moreover, the physical background for the derivation of the present method is given. The procedure for calculating a crown wall usually includes the following steps: i. The rate of wave overtopping determines the crest level of the crown wall. ii. The construction procedure and costs governs the crown wall foundation level. And finally, iii. a stability analysis determines the width and the other dimensions of the crown wall. If the upper berm of the armour layer is very low, the crown wall has to withstand most of the wave actions, including those of wave breaking at the wall. Traditionally, this type of structure is denoted composite breakwater. On the other hand, if the berm is higher than the maximum wave run-up level, then the design is not dominated by

F.L. Martin et al. r Coastal Engineering 37 (1999) 149174

151

wave actions and its overall dimensions are essentially dictated by functional requirements. Among these extreme cases, there are several alternatives ranging from high berm and small crown wall to see Hamilton and Hall, 1992. low berm and large crown wall. A very convenient solution is to build the upper berm high enough so that wave breaking always occurs on the armour layer; i.e., the crown wall will have to withstand only the pressures induced by broken waves. From an engineering point of view, the crown wall problem may be described as follows Fig. 1.: to determine crown wall geometrical dimensions crest elevation, foundation level and width. for a given design water level and wave characteristics as a function of the height and width of the armour layer upper berm. These dimensions must satisfy the functional requirements safely and economically. To solve this problem, it is necessary to define: 1. the geometry of the armour layer which guarantees wave breaking onto the slope, and 2. the pressure distribution of broken waves on a vertical wall, including uplift pressure. Next, the stability of the upright section has to be verified. 2.2. Wae breaking on the slope of rubble mound breakwaters Descriptions of pressure distribution when waves are impinging on vertical structures may be found in several papers. Nagai 1973. analysed wave pressure on structures induced by monochromatic standing waves, partially standing waves and breaking or broken waves. For non-breaking waves, the main feature of the time pressure distribution is the occurrence of a symmetrical double peak around the wave crest. Fig. 2 shows the time evolution of the wave pressure on a vertical wall, under different wave steepnesses. For waves with slight steepness reaching the wall, the pressuretime series induced by the standing wave show a sinusoidal shape. Increasing the wave steepness and keeping the wave period constant, the peak pressure at the bottom of the wall fluctuates with twice the wave frequency, Fig. 2a. As the wave steepness is further increased, the fluctuation expands up to the water surface. The

Fig. 1. Overall dimensions of the crown wall problem.

152

F.L. Martin et al. r Coastal Engineering 37 (1999) 149174

Fig. 2. Time evolution of wave pressure distribution on a vertical wall under increasing wave steepness after Losada et al., 1995..

double peak induced by the standing wave system is symmetric, Fig. 2b. The maximum wave pressure is always around the still water level. Further increasing of the wave steepness, being close to breaking conditions, the double peak of the pressuretime curve becomes asymmetric, with the former being shorter and higher, Fig. 2c. Oumeraci et al. 1993. pointed out that the asymmetry of the double peak indicates that a transition from a standing wave to a breaking wave system is taking place. Grilli et al. 1992. described the flow velocities and accelerations for waves close to the breaking conditions. When an incident wave breaks on the wall, the first peak may increase extraordinarily and may even split into two peaks with a very short duration, Fig. 2d. Bagnold 1939. called it Shock Pressure. This pressure and the effect on the structure stability has been studied by Chan and Melville 1988., Oumeraci et al. 1991., Peregrine 1994. and Topliss 1994.. The subsequent peak, denoted Secondary Pressure or Reflecting Pressure Topliss, 1994. has a relatively slow time variation, and a larger duration than the shock pressure, Fig. 2d. When a broken wave hits the wall, the double peak pattern of the time pressure distribution is still apparent. Their relative magnitude and duration depend on the distance between the breaking point and the hit wall, Fig. 2e. For the cases where the wave does not break directly on the wall, Fig. 2a, b and c, there are several theoretical solutions which provide the pressure and forces on the structure see Fenton, 1985 for references.. For the cases where impact forces occur

F.L. Martin et al. r Coastal Engineering 37 (1999) 149174

153

Fig. 2d, no theoretical approximation is valid and only impulse methods Cooker and Peregrine, 1990; Losada et al., 1995. or empirical approximations Nagai, 1973; Goda, 1985, revisited by Takahashi et al., 1992; Tanimoto and Takahashi, 1994; Oumeraci and Kortenhaus, 1997. are available. When the wave impinges the wall after breaking, empirical methods for a bore hitting a wall Ramsden and Raichlen, 1990. may be applied. In the case of a crown wall, the wave breaks on the armour layer, the wave pressure distribution on the crown wall is produced by the broken wave, and its characteristics depend on the wave evolution after breaking. For the application of the forthcoming equations it is necessary that the waves hit the crown wall as broken waves. If the wave does not break on the slope and it can break onto the crown wall, impulsive forces can occur which are not taken into account in the present method. The most frequently used breakwater slopes are in the range 1.5 - cotan b - 2.5. For rough seas, large wave height and period, or swell conditions, large wave period and moderate wave height, this range of slopes produces essentially collapsing wave breaking. It is well known that waves under swell conditions occur in groups and large waves are followed by other large waves with similar characteristics. Generally, for long period swell waves ) 15 s. under collapsing wave breaking conditions, the incoming wave generates minor interaction with the run-down of the previous one. Another characteristic of this type of breaker is that wave breaking always occurs around the SWL. Thus, the waves will hit the crown wall as broken waves if the slope extends from the order of the wave height Ac ) 0.8 to 0.9 H , see Fig. 1., measured vertically from the SWL. In this case, the wave breaks on the slope, and hits the crown wall during the run-up process. This is a common practice since, in Mediterranean countries, most of the crown walls built were designed following Iribarren recommendations and thus they have the level of the upper berm of the armour layer around Ac s H.

3. Description of model tests Scale model tests were conducted in the 70 m long, 2 m wide, 2 m high wave flume at the Ocean and Coastal Engineering Lab at the University of Cantabria. The test model Fig. 3. consisted of a 1r90 scale section of the Prncipe de Asturias breakwater at Port of Gijon Spain.. The Principe de Asturias breakwater is the main protective structure of El Musel Port, located at Gijon harbour in the North of Spain. The crown wall base level is 0.0 m over the low tide level zero datum., the level of the rubble berm is q13.5 m q12.2 m before 1995. and the level of the crown wall top is q18.35 m. The width of the wall is F s 18.72 m and the berm width is B s 3.75 m, which means that the berm is built of one unit of 120 tons. The armour layer slope is 1:1.5. The water depth at the breakwater toe is 21.0 m at LLWL, and the maximum tidal range in the zone is 4.5 m. In the tests, the water depth was set to correspond to high tide level in the prototype q4.0 m.. Monochromatic waves 36 tests. and random waves 15 tests. were generated

154

F.L. Martin et al. r Coastal Engineering 37 (1999) 149174

Fig. 3. Gijon breakwater cross-section.

F.L. Martin et al. r Coastal Engineering 37 (1999) 149174

155

by a piston-type wavemaker. The free surface in front of the structure was measured by three capacitance wave gauges and a reflection analysis of the free surface time series was performed. By using this technique it is possible to obtain the incident and the reflected wave time series. The transmitted wave height was measured by one free surface gauge located 1 m from the lee side toe of the breakwater. Four strain-gauge type pressure gauges were installed in the crown wall basement while eight gauges were fixed along the vertical structure front face. Pressures at the front face and under the crown wall were integrated by a rectangular method to obtain the forces induced by waves hitting the structure. The logging data rate was 120 Hz. One of the main targets of the tests was to identify and quantify the protective effect of the berm on the resulting pressures. Therefore, three berm widths were tested, corresponding to the length of 1 mound unit, 2 units and 3 units. Two types of armour were used corresponding to 90 and 120 ton blocks in the prototype.

4. Semi-empirical procedures for monochromatic waves In this section, the proposed method to calculate wave-induced pressures on crown walls is introduced. The method allows the calculation of i. wave pressure distribution on the crown wall front face Section 4.1. and ii. uplift pressures Section 4.2.. Since a single wave generates two peaks of pressure pressure pattern described in Section 4.1.1., there are two loading cases for each of the previous pressure distributions front and uplift. called dynamic and reflecting loads. Therefore, Sections 4.1 and 4.2 have three sub-sections: i. observed characteristics, where the overall patterns of the pressure distributions are described; ii. first peak: dynamic pressures, where the methodology to calculate dynamic pressure distributions is given; and iii. second peak: reflecting pressures, similar to ii. for the reflecting pressure case. In Section 4.3 the method is verified for monochromatic waves by comparison of calculated forces to empirical measurements. Finally, a practical application of the method is shown and the results from the proposed method and from other calculating methods are compared. 4.1. Horizontal pressure distribution at crown wall 4.1.1. Obsered characteristics From the results of the experimental study conducted at the Ocean and Coastal Engineering Laboratory of the University of Cantabria, Losada et al., 1995; Martin, 1995., it can be concluded that when the wave impinges the crown wall after breaking in the armour slope, the first peak is generated during the abrupt change of direction of the bore front due to the crown wall, while the second peak occurs after the instant of maximum run-up and is related to the water mass down-rushing the wall. The distinct nature of these peaks, denoted by A and B, may be observed in Fig. 4 where a force time series recorded in the lab is given and the pressure distributions at the wall at the times A and B, are shown. For the pressure distribution produced by the first peak A., dynamic pressure, two regions may be distinguished: the upper one, not

156

F.L. Martin et al. r Coastal Engineering 37 (1999) 149174

Fig. 4. Experimental dynamic and reflecting pressure distributions for broken waves.

protected by the rubble-mound layer, and the lower one, protected by the rubble-mound layer. In both regions, the pressure is almost constant, but higher in the upper region. The pressure profile due to the secondary peak B., reflecting pressure, linearly increases downwards. The crown wall and the armour layer are functionally-dependent elements: i. The hydrodynamics of the running-up water is modified by the presence of the crown wall. This modification of the flux affects the resulting forces on the armour units; ii. the armour layer characteristics slope, permeability, roughness, berm width, etc.. determine the characteristics of the run-up water tongue which hits the wall. Therefore, it is obvious that the design of both structures must be related. In the proposed method, the relation between the armour layers geometry, porosity, etc.. and the resulting pressures on the crown wall is taken into account. Following other authors, a semi-empirical approach to formulate the pressure distribution during the first and second peak is developed. The crown wall is partially protected by the armour layer, Fig. 1. Thus, for the pressure distribution, two regions may be distinguished. In the upper region, run-up water hits the wall directly. In the lower region, protected by the armour units, waves reach the wall after flowing through the porous armour layer. The lower region extends from the wall foundation level, z s wf , up to the berm level, z s Ac, where z is the vertical coordinate measured from the still water level, positive upwards. The upper region extends from z s Ac, up to the wall crest level, z s wc . In Fig. 5 the proposed pressure distributions are schematically summarised. 4.1.2. First peak: dynamic pressures Following Nagai 1973., Jensen 1984. and Tanimoto and Takahashi 1994. it may be established that the dynamic pressure, PS0 at the berm crest level z s Ac, in the non-protected region, is linearly related to the water tongue thickness at that level, S z s Ac. s S0 , and may be evaluated by the following expression Martin, 1995.: PS0 s a r g S0

1.

F.L. Martin et al. r Coastal Engineering 37 (1999) 149174

157

Fig. 5. Schematic representation of the pressure distribution on the wave screen.

where a is a non-dimensional parameter that will be analysed further in this paper, r is the water density and g is the gravity acceleration. In Fig. 5, the geometrical definition of S0 is shown. Because of the experimental evidence of the constancy of the dynamic pressures in positive z-direction Jensen, 1984; Martin, 1995. the pressures profile Pd z ., can be represented by, Pd z . s PS0 for z ) A c . 2. Moreover, the dynamic pressures in the lower region are also almost vertically constant, Fig. 4, and it was experimentally verified Martin, 1995. that it can be related to PS0 through an empirical parameter l., analysed further in this section, where l is smaller than one, for wf - z - A c . Pd z . s l PS0 3. The determination of S0 and a is based on the experimental evidence that the dynamic and reflecting pressures occur shortly before and shortly after the instant of maximum run-up on the wall, respectively. Assuming that: 1. S0 and a may be evaluated during the occurrence of the maximum run-up, and that 2. for collapsing breakers, the water tongue kinematics and dynamics on the edge of the berm during the maximum run-up event is approximately the same as for slopes with or without a crown wall. Then, it may be concluded that the values of S0 and a depend only on the maximum run-up, Ru, on an infinite slope and on the water particle velocity at the jet tip. Losada and Gimenez-Curto 1981., based on experimental work under monochromatic waves and normal incidence, proposed the following expression for Ru on an infinite slope: Ru s Au 1 y e Bu Ir . 4. H

158

F.L. Martin et al. r Coastal Engineering 37 (1999) 149174

where, Au and Bu Fig. 6, after Losada 1992.. are empirical coefficients, which depend on the type of armour unit and Ir is the Iribarren number defined by, Ir s tan b

H L0

5.

where b is the slope angle, H is the local wave height and L0 is the deep water wavelength. In order to verify the applicability of Eq. 4., an experimental evaluation of Ru on a mound breakwater with a crown wall was carried out. The armour layer was built of concrete parallelepipedic blocks a = a = 1.25 a where a s 3.8 cm. with a 1:1.5 slope and the berm was built of 2 units. Because of experimental uncertainty, the same test same structure and same waves. was repeated three times. In Fig. 7 the best fit curve to the values of Ru is shown. The best fit values for Au and Bu are 1.2 and y0.7, respectively. These values are very close to the values proposed by Losada and Desire 1984. Au s 1.2, Bu s y0.65, for parallelepipedic blocks., giving support to assumption 2.. 1984. and Yamamoto and Horikawa 1992. showed that the Gunbak and Gokce water surface at the instant of maximum run-up may be approximated by a straight line

Fig. 6. Au and Bu coefficients for run-up calculation as a function of armour layer porosity.

F.L. Martin et al. r Coastal Engineering 37 (1999) 149174

159

Fig. 7. Comparison of measured and calculated run-up.

as indicated in Fig. 5.. Then, the water tongue thickness, S, of the impinging bore at the level z . may be evaluated by: z S z . s Sw 1 y 6. Ru where S w is the water tongue thickness at the SWL. For 1.5 - cotan b - 2.5, and following the theoretical and experimental work of Yamamoto and Horikawa 1992., it may be assumed without significant error that the water tongue thickness at the SWL is on the same order of the wave height, S w ; H. Hence, the thickness, S0 , of the bore at the berm crest, z s Ac, is given by, Ac S0 s H 1 y . 7. Ru

Assuming no energy loss due to friction above Ac level, the alongslope bore tip celerity can be calculated from gravity acceleration as: C b z . s 2 Ru y z . g .

8.

Therefore, the horizontal component of the bore tip celerity, C bx , at any level z G Ac, may be approximated by the following expression Martin, 1995.: C bx z . s 2 Ru y z . g cos b .

9.

Next, the averaged horizontal velocity of the water particles near the bore front, x , can be considered to be equal to the celerity of the tip see Ramsden and Raichlen, 1990.. Thus, x at the berm crest level z s Ac. can be obtained by Eq. 9.. For the calculation of a , results from Cross 1967. and Wiegel 1970. for the maximum pressure induced by a bore hitting a vertical wall pmax . are considered: Pmax s r Cf Nf2
2 C bx

10 .

160

F.L. Martin et al. r Coastal Engineering 37 (1999) 149174

Thus, the maximum pressure is defined as Cf Nf2 times the stagnation pressure due to the bore celerity, where Nf is a dimensionless parameter defined as: Nf s Cb

'gS

11 .

where S is the bore thickness, and Cf is a coefficient that depends only on the bore front geometry Cumberbatch, 1960.. Using Eqs. 7. and 8. into Eq. 11. for z s Ac, the resulting value for Nf is Nf s

Ru H

12 .

By using Eqs. 1., 7., 9., 10. and 12., for z s Ac, the dimensionless parameter a is given by:

a s 2 Cf

Ru H

cos b

13 .

Note that Eq. 13. is valid if the horizontal component of the bore celerity C bx . is not affected by the berm width. This assumption has been experimentally verified by the authors for berm widths up to BrL s 0.1 Martin et al., 1998.. Cf represents the short-duration pressure oscillations induced by the impact of the water tongue front to the vertical wall. Cumberbatch 1960. showed that it depends only on the bore wedge angle, Q . For wedge angles of 22.58 and 458, Cumberbatch 1960. found Cf to be 1.4 and 2.1, respectively. Cross 1967. computed the coefficient Cf for wedge angles between 08 and 758 and found the following relation agreed with his results: Cf s 1 q tan u .
1.2

14 .

Fig. 8 shows the values of Cf calculated from the maximum pressures PS0 measured in the tests described in Section 3, using Eqs. 1., 7. and 13.. If Cf is calculated from pressures of which the persistence is larger than Tr100 instead of maximum pressures., the dispersion of the calculated Cf is smaller than the dispersion shown in Fig. 8, and the best fit value for Cf is 1.0. To define Cf for design purposes the dynamic response of the structure must be taken into account. For small structures low inertia. andror rigid foundations, maximum pressures must be taken as the design pressures and a value of Cf s 1.45 Q s 278. is proposed see Fig. 8.. Finally, a can be calculated as:

a s 2.9

Ru H

cos b

15 .

For large structures large inertia. andror elastic foundations, Cf s 1.0 could represent better the equivalent static loading situation for design purposes. Further research on dynamic response of crown walls and on scale effects is required for better assessment of the design value of Cf .

F.L. Martin et al. r Coastal Engineering 37 (1999) 149174

161

Fig. 8. Values of Cf calculated from PS0 measured and Eqs. 1., 7. and 13..

The parameter l was evaluated experimentally from monochromatic wave tests in the flume. A description of the experimental set up is given in Section 2. The experimental values obtained for l are shown in Fig. 9. The range of the measured l values 0.25 - l - 0.65., is in agreement with those given by Jensen 1984. and Gunbak and 1984.. Notice, that the experimental wave steepness range is 0.03 - HrL at Gokce breakwater toe. - 0.075. The best fit curve to the empirical results is:

l s 0.8ey1 0.9 B r L .

16 .

4.1.3. Second peak: reflecting pressures The second peak occurs shortly after the occurrence of the maximum wave run-up. The horizontal and vertical velocities at this instant are very small as are the accelera-

Fig. 9. Experimental variation of l vs. B r L H and L measured at the toe of the breakwater..

162

F.L. Martin et al. r Coastal Engineering 37 (1999) 149174

tions. Thus, the resulting pressure field around the crown wall may be considered almost hydrostatic. Furthermore, the pressure distribution recorded on the wall, at the instant of the occurrence of the peak, is continuous along the protected and unprotected regions. Then, the reflecting pressure, Pr , may be evaluated by the following linear expression: Pr z . s mr g S0 q Ac y z . for wf - z - Ac q S0

17 .

where the dimensionless parameter, m F 1, was evaluated experimentally from monochromatic wave tests. From the tests, it is clear that m depends on the wave steepness HrL. and on the non-dimensional berm width, Brle. The parameter le is the equivalent size of the rubble units, and is calculated by, le s

gr

18 .

where W and gr are the total weight and the specific weight of the armour unit, respectively. The experimental values for m are shown in Fig. 10. For wave steepness, HrL - 0.02, the reflecting pressures are r gz m s 1., decaying to 0.5 r gz m s 0.5., approximately, for HrL ; 0.04. By increasing the wave steepness to 0.075, an asymptotic trend is obtained, which depends on the number of units building the berm. m takes values of 0.45, 0.37 and 0.3 for one, two and three armour units on the berm, respectively.

Fig. 10. Experimental variation of m vs. H r L and B rle as a parameter H and L measured at the toe of the breakwater..

F.L. Martin et al. r Coastal Engineering 37 (1999) 149174 Table 1 B rle 1 2 3 a 0.446 0.362 0.296 b 0.068 0.069 0.073 c 259.0 357.1 383.1

163

The trend of experimental values for m can be well represented by an exponential 2 curve of the type m s ae c H r Ly b . . The best fit parameters for these curves are shown in Table 1: The experiments in which l and m are determined were performed with large armour units 1200 kN parallelepipedic blocks in the prototype. and a large porous core. Further experiments are needed to analyse in more detail the effect of unit size and core permeability on these parameters. Meanwhile, the proposed values of l and m can be used as a first approach for design purposes. 4.2. Uplift pressure distribution at crown wall 4.2.1. Obsered characteristics It is common practice in maritime engineering Iribarren and Nogales, 1964. to consider a linear variation of wave pressure under the crown wall. Losada et al. 1993., applying linear wave theory, obtained a parabolic pressure distribution under an impermeable crown wall resting on a porous media, with porosity ranging from 20% to 40%. However, their findings do not differ significantly from the linear trend. In this paper, the linear law is assumed. To define this linear distribution, the pressures at the toe and at the heel of the crown wall were experimentally recorded with the same experimental set-up described in Section 3. 4.2.2. First peak: dynamic pressures Dynamic pressure beneath the seaward edge of the structure is approximately equal to l PS0 . At the heel dynamic pressures are negligible. 4.2.3. Second peak: reflecting pressures Reflecting pressure beneath the seaward edge of the structure is equal to the pressure at the front. Reflecting pressures at the heel are only significant if the crown wall is founded below the transmitted wave amplitude. Therefore, heel pressures depend on the wave transmission process. Thus, the following values are adopted see Fig. 5.: Seaward edge: ( dynamic pressure s l PS0 z s wf . ( reflecting pressure s Pr z s wf . s Pre Heel: ( dynamic pressure negligible, Pra s 0 ( reflecting pressure s Pra , from Losada et al. 1993..

164

F.L. Martin et al. r Coastal Engineering 37 (1999) 149174

Fig. 11. Reflecting pressure at the heel vs. F r L L calculated at the toe of the breakwater.. Dots are experimental data from Gijon prototype measured on 10 February 1996 n s 0.4 approximately..

Fig. 11 shows the reflecting pressure at the heel, Pra , non-dimensionalised with the reflecting pressure at the seaward edge, Pre , vs. FrL, where F is the crown wall width. Each plotted curve corresponds to a different mound porosity, n. The numerical model employed to generate the curves Losada et al., 1993. was applied for a single overall porosity of the rubble mound. For design purposes the porosity selected must represent the porosity of the material on which the crown wall is founded. The proposed curves neither depends on the wave height nor on the water depth. This method for calculating uplift pressures must be regarded as a first engineering approach to the problem since: i. lines in Fig. 11 were obtained from linear theory, and ii. the experimental values were obtained from a low crown wall, where air entrainment is very low. Additional experiments for high crown walls based on less porous core are required to complete the method. Some results from the pressure gauges placed under Gijons breakwater crown wall prototype. are included dots. see Fig. 11.. This breakwater is built of an armour of 120 tons resting on a core of 90-ton parallelepipedic blocks and the overall porosity is approximately n s 0.4. 4.3. Application and erification 4.3.1. Comparison to test data As a first verification of the proposed method, experimental pressure and forces on the scale model of Gijons crown wall were compared to their corresponding analytical values for regular waves and random waves. As an example, Fig. 12a shows the comparison of maximum forces measured in the lab for regular waves tests described in Section 3. and those calculated by the proposed method Cf s 1.45.. Experimental forces per linear meter of structure are integrated from measured pressures on the wall. Furthermore, the values of the forces were obtained as average values of the three highest measured forces in each test for monochromatic waves.

F.L. Martin et al. r Coastal Engineering 37 (1999) 149174

165

Fig. 12. a. Comparison between experimental and calculated forces. Tests done with monochromatic waves. b. Comparison between experimental and calculated dynamic pressures. Tests done with random waves, calculation done by splitting of single waves by zero-upcrossing.

In Fig. 12b the dynamic pressures PS0 . measured in the lab for random waves are compared to the calculated dynamic pressures for each individual single wave. Single waves H ,T . are identified from measured free surface time series by zero-upcrossing. The method is applied to each individual wave. PS0 is calculated and compared to the

166

F.L. Martin et al. r Coastal Engineering 37 (1999) 149174

correspondent vertically averaged measured dynamic pressure in the unprotected region of the crown wall. The averaging procedure is as follows: i. the pressure time series are integrated by a rectangular method to force time series, ii. the maximum dynamic force and maximum dynamic force time are identified for each individual wave, iii. the pressure measurements at this time in the unprotected zone is integrated by a rectangular method for each individual wave., iv. the resulting force in the unprotected region is divided by the unprotected region height to give a vertically-averaged pressure in this zone. The comparison gives relatively good results. The crown wall of Gijon Breakwater has been instrumented to supply full-scale wave pressure records under storm conditions. This information will help validate the present method when severe storm data become available. Some preliminary results of the measurements performed for uplift pressures were presented in Section 4.2.3. 4.3.2. Comparison to other methods In this section, forces on the crown wall of the Gijon Breakwater calculated by 1984. and the present method are Iribarren and Nogales 1964., Gunbak and Gokce compared. An individual wave height of H s 12.0 m, a period of T s 16.0 s and a high tide level q4.0 over LLWL. is adopted. The proposed method is applied with the following values of the main parameters: Ru s 13.2 m calculated by Eq. 4. for Au s 1.2 and Bu s y0.65., Cf s 1.45, BrL s 0.016, l s 0.66, Fig. 9., HrL s 0.052, Brle s 1, m s 0.48, Fig. 10., FrL s 0.08, n s 0.4, Pra rPre s 0.375 Fig. 11. and Ac s 9.5 m. A value of 0.6 was assumed for the friction coefficient between the crown wall and the core material. The weight of the crown wall when the SWL is 4.0 m above LLWL is 3460 kNrm. The resulting value for S0 is 3.37 m and for a is 2.43. In Table 2, the net horizontal force per unit length as well as the uplift forces are given. The sliding and overturning safety coefficients were investigated and for all cases the lower safety factors occurred for the sliding of the crown wall. The safety conditions for sliding calculated from experiments on the wave flume are also presented in Table 2. The pressure distribution in the reflecting pressure condition is linear, growing downwards, thus maximum loads occur at the bottom of the crown wall. Therefore the lower the crown wall foundation is, the larger the pressures at the bottom and the uplift pressures are. Then, the large uplift force in the reflecting pressure condition in this

Table 2 Method Iribarren Gunbak Present method Dynamic pressure condition Reflecting pressure condition Horizontal force kNrm. 2285 1231.5 1044 695 Uplift force kNrm. 1919 1259 512 1055 Calculated safety coefficient 0.41 1.08 1.69 2.08 Lab. safety coefficient 1.72 2.01

F.L. Martin et al. r Coastal Engineering 37 (1999) 149174

167

particular case is due to the low foundation level of the Gijon breakwaters crown wall 4 m below SWL.. By comparing the calculated and measured forces and safety coefficients it can be concluded that the Iribarren and Nogales method and the Gunbak method are pessimistic when analysing the static stability of the crown wall. Wave records over the last 10 years show that storms with significant wave heights greater than 8 m and maximum wave heights greater than 12 m have attacked the breakwater several times. After a visual inspection of the breakwater it can be stated that the crown wall along the breakwater trunk has not experienced displacement.

5. Force distribution for random waves The extension of the previous method to irregular waves is based on the following considerations: 1. The reference parameter for the application of the formulae for calculating the dynamic pressures, see Eqs. 1., 7. and 15., is the run-up on a straight slope. 2. The hypothesis of equivalence introduced by Saville 1962. can be applied for computing run-up distribution on a rough, permeable slope. Given a sea state defined by the significant wave height, Hs, the zero crossing averaged wave period, Tz, the value of the total forces on the wall generated by the dynamic and reflecting pressures may be considered random variables which can give different values for each individual wave H ,T . of the sea state. The hypothesis of equivalence proposes that the distribution function of a random variable may be obtained by assigning to each individual irregular wave the same phenomenon value which would be produced by a periodic train of the same wave height and period. This hypothesis was empirically proven by Bruun and Johannesson 1977. and Bruun 1978. for run-up on rough, permeable slopes. and Gunbak Thus, the distribution function of the forces on a crown wall, under a given sea state, may be obtained by assigning the same value of the force that would be produced by a periodic wave train of the same height and period, to each individual irregular wave. It is important to note the statistical nature of this hypothesis. It does not necessarily imply that each individual wave produces the same forces as the equivalent regular wave, but is less restrictive as it refers to average rather than to individual values. 5.1. Wae force distribution The procedure to compute the distribution function of the forces on the crown wall under an irregular sea state is as follows. 1. Given Hs, Tz, the water depth at the toe of the breakwater h., and the spectral shape parameter, a theoretical TMA spectrum is computed. 2. From this spectrum a synthetic free surface time series is generated.

168

F.L. Martin et al. r Coastal Engineering 37 (1999) 149174

Fig. 13. Ciervana breakwater cross-section.

3. The following breaking criteria are applied to each individual wave: H - 0.142 L tanh 2p drL . Miche criterion. .

19 . 20 .

Hrd . max s 0.55 q 0.88 exp y0.012 cot w . .

where w is the bed slope angle. Since results will be compared to laboratory data, the breaking criterion Eq. 19.. is established for waves in the laboratory with mild bottom slopes Nelson, 1997.. For steeper slopes andror field conditions there are several formulations in the literature e.g., Komar and Gaughan, 1972.. 4. Waves which break due to limited depth are regenerated with a wave height equal to half the water depth, upper limit of values proposed by Dally et al. 1985.. 5. Eqs. 1., 3., 7., 15. and 17. and uplift pressures are calculated for each individual wave height and, after integration, the total force per unit length of wall produced by the dynamic and reflecting pressures are obtained. 6. The force sample data are analysed statistically and the best distribution function is ascribed to the forces under the dynamic and reflecting pressure conditions. 5.2. Application for random waes and its comparison to data from other authors The described methodology has been applied to two cases previously tested in the laboratory: Ciervana Breakwater in Bilbaos Harbour Spain. tested by Burcharth et al. 1995. and a model test carried out by Jensen 1984..

Table 3 Sea state parameters for Ciervana breakwater Hs m. 8 9 10 11 12 Tp s. 15 16 17 19 20

F.L. Martin et al. r Coastal Engineering 37 (1999) 149174 Table 4 Simulation parameters Peak enhancement factor Number of waves generated 1.4 3000

169

5.2.1. Cierana breakwater The present method is applied to Ciervana Breakwater crown wall Bilbao, Spain., following the procedure described in Section 5.1. The need of space to support new developments in Bilbao harbour led to the building of the Ciervana Breakwater to protect the land reclamation between the inner port and Punta Lucero Breakwater. It was built south-east of the famous Punta Lucero Breakwater 150-ton armour units. and it is partially protected by this breakwater from NW storms. Its total length is 3.15 km, and the armour layer is built of parallelepipedic 100-ton blocks a = a = 1.25 a. in a 1:2 slope. The crown wall base level is q1.5 m over the zero datum, the crest level of the rubble berm is q14.0 m and the crown wall top level is q18.0 m see Fig. 13.. The berm width is B s 9.0 m and the width of the wall is F s 29.0 m. The water depth at the breakwater toe is 26.0 m at LLWL, and the maximum tidal range in the zone is 4.5 m. The significant wave heights Hs. and peak periods Tp. and the characteristics of the wave simulation for the test are given in Tables 3 and 4, respectively. Fig. 14 shows the wave height distributions: i. the distribution obtained by the computation described in Section 5.1, ii. the distribution measured in the laboratory Burcharth et al., 1995. and iii. the Rayleigh distribution. In Fig. 15 the 0.1% exceedance probability force obtained from the present method is compared to the experimental results described by Burcharth et al. 1995.. The agreement is good.

Fig. 14. Wave height distribution computed by the present method, measured in the lab by Burcharth et al. 1995. and Rayleigh distribution.

170

F.L. Martin et al. r Coastal Engineering 37 (1999) 149174

Fig. 15. The 0.1% probability horizontal force, FH , measured by Burcharth et al. 1995., and computed from present method.

Moreover, the 0.1% exceedance probability pressure measured at level q1.5 m at the foot of the crown wall for the case Hs s 11 m, is 120 kPa while the computed pressure is 132 kPa. Finally, from measurements performed by CEDEX Madrid. for the case Hs s 11 m, also described by Burcharth et al. 1995., the centre of application of the total horizontal forces corresponding to the 0.3% probability is estimated to be between 11.5 and 12.5 m above SWL. The present method locates the application point 12.4 m above SWL. 5.2.2. Comparison to the data of Jensen (1984) Jensen 1984. published a monograph for the design of rubble mound breakwaters. Data on crown wall forces are also included. For one test the author shows the full distribution of forces which makes a comparison possible. The cross-section of the breakwater is shown in Fig. 16. The armour layer is built of rectangular blocks 2.9 = 2.9 = 4.2 m 82 ton. in a 1:2 slope, the berm width is 6.0 m 2 units. and the berm height is Ac s 10.9 m, the crown wall foundation is wf s 4.30 m, the crown wall height

Fig. 16. Cross-section of breakwater tested by Jensen 1984..

F.L. Martin et al. r Coastal Engineering 37 (1999) 149174

171

Fig. 17. Comparison of Jensens experimental results to calculations from the present method.

is 16.7 m. The SWL in the test is at level q5.3 and the peak period is 18 s. Three significant wave heights were tested: Hs s 8, 11 and 14 m. For the calculation, a TMA spectrum peak enhancement factor s 2. is used, following the procedure defined in Section 5.1. For the calculation of run-up, Eq. 4. was used for Au s 1.2 and Bu s y0.7. The comparison of results is shown in Fig. 17. Differences in the maximum values of the forces for Hs s 14 m could be due to the breaking criterion used in the method. The agreement is relatively good.

6. Concluding remarks and future work A new method to calculate forces on crown walls based on previous works and on new experiments, carried out under monochromatic waves, is presented. The method is extended to irregular waves via the hypothesis of equivalence and checked against two sets of lab tests by: i. Burcharth et al. 1995. and ii. Jensen 1984.. The comparison is fairly good. The application of the new method requires that the waves break onto the main slope; i.e., only broken waves will reach the crown wall. Under such conditions, the time pressure evolution on the wall has two peaks: the first pressure peak which is usually the largest and the secondary peak which is smaller but lasts longer. In some cases HrAc small andror BrL large. the reflecting pressure peak can be larger than the dynamic pressure peak. The magnitude of these peak are related to the armour berm geometry: berm width B . and berm height Ac.. The proposed dynamic and reflecting pressure distribution

172

F.L. Martin et al. r Coastal Engineering 37 (1999) 149174

depends on two parameters, l and m , which are experimentally evaluated and depend on the relative berm width BrL., local wave steepness HrL. and the number of armour units constituting the berm nb s Brle.. Dynamic pressures are generated during the abrupt change of direction of the bore front due to the crown wall. Short duration pressure oscillations are produced in this instant, which depend on the bore front wedge characteristics parameter Cf .. These oscillations could be filtered by the dynamic response of the crown wall and its foundation. Moreover, scale effects could affect these oscillations due to aeration, saltrfresh water testing, water compressibility, etc. Further research on dynamic response of crown walls and on scale effects is being carried out by the authors for better assessment of the design value of Cf for design purposes.

Acknowledgements This research was partially funded by the European Community Research Programme, MAST III Marine Science and Technology., Project PROVERBS Probabilistic Design Tools for Vertical Breakwaters. under EU contract MAS3-CT95-0041. This financial support is very much appreciated. The first author wishes to thank the Ministerio de Educacion y Ciencia for his funding during part of the research F.P.I. Research Grant.. The authors want to thank Autoridad Portuaria de Gijon for its continuous technical support, Prof. C. Vidal for his unconditional help and Prof. Burcharth and Prof. Oumeraci for their very interesting and useful comments in the review of the paper.

References
Bagnold, R.A., 1939. Interim report on wave pressure research. J. Inst. Civil Eng. 12, 201226, 19381939. Bradbury, A.P., Allsop, N.W.H., Stephens, R.V., 1988. Hydraulic performance of breakwater crown walls. Report SR146. H.R. Wallingford. Bruun, P., Gunbak, A.R., 1978. Stability of sloping structures in relation to j s tan a r H r L0 , risk criteria in design. Coastal Engineering 1, 287322. Bruun, P., Johannesson, P., 1977. Parameters affecting stability of rubble mounds. Closure ASCE Journal of Waterway, Port, Coastal and Ocean Division 103 WW4., 553566. Burcharth, H.F., Frigaard, P., Berenguer, J.M., Gonzalez, B., Uzcanga, J., Villanueva, J., 1995. Design of the Ciervana breakwater, Bilbao. In: Telford, T. Ed.., Proc. 4th Coastal Structures and Breakwaters, Chap. 3. Institution of Civil Engineers. Chan, E.S., Melville, W.K., 1988. Deep-water plunging wave pressures on a vertical plane wall. Proc. R. Soc. London, Ser. A 417, 95131. Cooker, M.J., Peregrine, D.H., 1990. A model for breaking wave impact pressures. Proc. 22nd Int. Conference on Coastal Engineering. ASCE, Delft, Holland, pp. 14731486. Cross, R.H., 1967. Tsunami surge forces. Journal of Waterway and Harbour Division, ASCE 93 4., 201231. Cumberbatch, E., 1960. The impact of a water wedge on a wall. J. Fluid Mech. 7 3., 353373. Dally, W.R., Dean, R.G., Dalrymple, R.A., 1985. Wave height variation across beaches of arbitrary profile. J. Geophys. Res. 90, 1191711927.

'

F.L. Martin et al. r Coastal Engineering 37 (1999) 149174

173

Fenton, J.D., 1985. Wave forces on vertical walls. Journal of Waterway, Port, Coastal and Ocean Division, ASCE 111 4., 693717. Goda, Y., 1985. Random Seas and Design of Maritime Structures. University of Tokyo Press. Grilli, S.T., Losada, M.A., Martin, F.L., 1992. Wave impact forces on mixed breakwaters. Proc. 23rd Int. Conference on Coastal Engineering. ASCE, Venice, Italy, pp. 11611174. Gunbak, A.R., Gokce, T., 1984. Wave screen stability of rubble mound breakwaters. International Symposium of Maritime Structures in the Mediterranean Sea. Athens, Greece, pp. 2.992.112. Hamilton, D.G., Hall, K.R., 1992. Preliminary analysis of the stability of rubble mound breakwater crown walls. Proc. 23rd Int. Conference on Coastal Engineering. ASCE, Venice, Italy, pp. 12171230. Iribarren, R., Nogales, C., 1964. Obras Martimas. Dossat Ed.., Madrid, 376 pp. Jensen, O.J., 1984. A Monograph on Rubble Mound Breakwaters. Danish Hydraulic Institute. Komar, P.D., Gaughan, M.K., 1972. Airy wave theory and breaker height prediction. Proc. 13th Int. Conference on Coastal Engineering. ASCE, Vancouver, Canada, pp. 405418. Losada, M.A., 1992. Recent developments in the design of mound breakwaters. In: J.B. Herbich Ed.., Handbook of Coastal and Ocean Engineering, Chap. 21, pp. 9391050. Losada, M.A., Desire, J.M., 1984. Functional comparison of breakwater armour units. International Symposium of Maritime Structures in the Mediterranean Sea. Athens, Greece, pp. 2.332.44. Losada, M.A., Gimenez-Curto, L.A., 1981. Flow characteristics on rough, permeable slopes under wave action. Coastal Engineering 4, 187206, Elsevier. Losada, I.J., Darlrymple, R.A., Losada, M.A., 1993. Water waves on crown breakwaters. Journal of Waterway, Port, Coastal and Ocean Engineering, ASCE 199r4, 367380. Losada, M.A., Martin, F.L., Medina, R., 1995. Wave kinematics and dynamics in front of reflective structures. Wave Forces on Inclined and Vertical Wall Structures. Task Committee on Forces on Inclined and Vertical Wall Structures, ASCE, pp. 282310. Martin, F.L., 1995. Estudio hidrodinamico de la interaccion de ondas de gravedad con estructuras reflejantes. PhD Dissertation. Univ. de Cantabria, Spain in Spanish.. Martin, F.L., Medina, R., Teixeira, L., 1998. Estudio experimental del rebase sobre el dique del Puerto de Llanc a. Internal Report. Universidad de Cantabria, 66 pages in Spanish.. Nagai, S., 1973. Wave Forces on Structures. Advances in Hydroscience, Vol. 9. Academic Press, New York, pp. 253324. Nelson, R., 1997. Height limits in top down and bottom up wave environments. Coastal Engineering 32 23., 247254, Elsevier. Oumeraci, H., Kortenhaus, A., 1997. Wave impact loadingtentative formulae and suggestions for the development of final formulae. Paper 1.0.2. Proceedings of 2nd PROVERBS Task 1 Workshop, Edinburgh. H.R. Wallingford. Oumeraci, H., Partenscki, H.W., Tautenhain, E., 1991. Large scale model investigations: a contribution to the revival of vertical breakwaters. Conf. on Coastal Structures and Breakwaters 91. Inst. of Civil Engineers, London, UK, pp. 207220. Oumeraci, H., Klammer, P., Partenscky, H.W., 1993. Classification of breaking wave impact loads on vertical structures. Journal of Waterway, Port, Coastal and Ocean Engineering 119r4, 381397. Pedersen, J., Burcharth, H.F., 1992. Wave force on crown walls. Proc. 23rd Int. Conference on Coastal Engineering. ASCE, Venice, Italy, pp. 14891502. Peregrine, D.H., 1994. Pressure on breakwaters: a forward look. Proc. of Int. Workshop on Wave Barriers in Deepwaters. Yokoshuka, Japan, pp. 553571. Ramsden, J.D., Raichlen, F., 1990. Forces on vertical wall caused by incident bores. Journal of Waterway, Port, Coastal and Ocean Division, ASCE 116r5, 592613. Saville, T., 1962. An approximation of the wave run-up frequency distribution. Proc. 8th International Conference on Coastal Engineering. ICCE, Mexico City. Takahashi, S., Tanimoto, K., Shimosako, S., 1992. Experimental study of impulsive pressures on composite breakwaters. Rep. Port Harbour Research Institute 31 5., 3574. Tanimoto, K., Takahashi, S., 1994. Design and construction of caisson breakwaters: the Japanese experience. Coastal Engineering 22, 5778, Elsevier.

174

F.L. Martin et al. r Coastal Engineering 37 (1999) 149174

Topliss, M.E., 1994. Water wave impact on structures. PhD Dissertation. School of Mathematics, Univ. of Bristol, UK. Wiegel, R.L., 1970. Tsunamis. In: Wiegel, R.L. Ed.., Earthquake Engineering. Prentice-Hall, Englewood Cliffs, NJ, pp. 253306. Yamamoto, Y., Horikawa, K., 1992. New methods to evaluate wave run-up height and wave overtopping rate. Proc. 23rd Int. Conference on Coastal Engineering. ASCE, Venice, Italy, pp. 17341747.

Potrebbero piacerti anche