You are on page 1of 23

Ocean Engineering 38 (2011) 1031–1053

Contents lists available at ScienceDirect

Ocean Engineering
journal homepage: www.elsevier.com/locate/oceaneng

Review

Hydraulic performance and wave loadings of perforated/slotted


coastal structures: A review
Zhenhua Huang a,b,n, Yucheng Li c,1, Yong Liu d,2
a
School of Civil and Environmental Engineering, Nanyang Technological University, Singapore
b
Earth Observatory of Singapore, Nanyang Technological University, Singapore
c
State Key Laboratory of Coastal and Offshore Engineering, Dalian University of Technology, PR China
d
College of Engineering, Ocean University of China, PR China

a r t i c l e i n f o abstract

Article history: This paper reviews recent progress in the study of perforated/slotted breakwaters, with an emphasis on
Received 24 June 2010 two main groups of such breakwaters: (1) perforated/slotted breakwaters with impermeable back
Accepted 19 March 2011 walls, and (2) perforated/slotted breakwaters without a back-wall. The methods commonly used to
Editor-in-Chief: A.I. Incecik
simulate the interactions between such structures and various linear/nonlinear waves are summarized.
Available online 26 May 2011
The transmission and reflection characteristics of perforated/slotted breakwaters in these two groups
Keywords: are reviewed extensively. Several methods for calculating wave forces on perforated caissons are also
Breakwaters reviewed. Some recent works published in Chinese journals, which are generally not well-known to
Seawalls non-Chinese researchers, are reviewed with a hope that these works can be beneficial to other
Jarlan-type structures
researchers working in this area.
Slotted/perforated wave barriers
& 2011 Elsevier Ltd. All rights reserved.
Wave scattering
Coastal structures

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1032
2. Reflection characteristics of Jarlan-type perforated structures. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1032
2.1. Fully perforated Jarlan-type structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1032
2.2. Partially perforated Jalan-type structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1034
2.3. Multiple perforated front walls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1034
2.4. Perforated caissons with top cover plate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1035
2.5. Perforated breakwaters with rock core . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1036
2.6. Perforated structures with internal horizontal plate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1036
2.7. Wave direction, wave irregularity, wave steepness, and rubble foundation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1037
2.7.1. The direction of incident waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1037
2.7.2. Wave irregularity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1037
2.7.3. Wave steepness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1038
2.7.4. Rubble foundation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1038
3. Slotted barriers without back-wall (wave screens) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1038
3.1. Surface-piercing slotted breakwaters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1038
3.1.1. Regular waves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1038
3.1.2. Irregular waves. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1041
3.1.3. Effects of near-shore currents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1041
3.2. Submerged slotted/perforated breakwaters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1041
4. Energy dissipation mechanisms perforated/slotted walls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1042
4.1. Regular waves in the absence of currents. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1042

n
Corresponding author at: School of Civil and Environmental Engineering, Nanyang Technological University, Singapore. Tel.: þ 65 67904737.
E-mail addresses: zhhuang@ntu.edu.sg (Z.H. Huang), liyuch@dlut.edu.cn (Y.C. Li), liuyong@ouc.edu.cn (Y. Liu).
1
Tel.: þ86 411 84709527; fax þ86 411 84708526.
2
Tel.: þ86 532 66781129; fax þ86 532 66781550.

0029-8018/$ - see front matter & 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.oceaneng.2011.03.002
1032 Z.H. Huang et al. / Ocean Engineering 38 (2011) 1031–1053

4.1.1. Energy loss coefficient for surface-piercing slotted/perforated walls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1042


4.1.2. Porous-effect parameter for surface-piercing slotted/perforated walls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1044
4.1.3. Relation between energy loss coefficient and porous-effect parameter for surface-piercing structures. . . . . . . . . . . . . . . . . 1044
4.1.4. Complex structure parameter for submerged slotted/perforated walls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1044
4.2. Irregular waves in the absence of steady currents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1045
4.2.1. Approach based on regular wave models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1045
4.2.2. Approach based on Gaussian random processes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1045
4.3. Regular waves in the presence of a steady current . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1045
4.4. Solitary waves in the absence of currents. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1046
4.5. Oblique waves in the absence of currents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1046
5. Wave forces on perforated caissons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1046
5.1. Takahashi’s method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1046
5.2. Tabet-Aoul and Lambert’s method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1047
5.3. Research in DUT, China, on wave forces on perforated caissons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1048
5.4. Comments on the three reviewed methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1051
6. Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1051
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1051
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1052

1. Introduction
slotted/perforated walls and submerged slotted/perforated walls.
After that, three methods are reviewed for calculating wave forces
Breakwaters and seawalls are two widely used coastal struc- on perforated caissons. The last section summarizes the main
tures. Breakwaters are normally constructed either parallel or findings in the reviewed literature.
perpendicular to the coast, and mainly serve the purpose of
enabling safe navigation into harbors, providing calm water areas
for loading and unloading of ships or crafts during storms and 2. Reflection characteristics of Jarlan-type
protecting shorelines from wave-induced erosion. Seawalls, on perforated structures
the other hand, are usually constructed along the coastline, and
their main functions are to protect the shore, reduce shoreline After Jarlan (1961) first proposed his design of a breakwater
erosions, or control wave-induced overtopping rates. Tradition- with a chamber between a perforated wall and an impermeable
ally, representative seawalls are vertical walls, curved concrete wall, several modifications have been proposed to improve the
walls, revetments, or rubble-mounds. Typical breakwaters are the reflection characteristics and structure stability. Jarlan-type per-
rubble-mound breakwaters, floating breakwaters, and perforated- forated breakwaters reviewed in this section are shown in Fig. 1,
wall caisson breakwaters. where 6 main types are included for the convenience of present-
Perhaps Jarlan (1961) was the first one to discuss the perfo- ing the review materials in this section.
rated-wall caisson breakwaters (now this type of breakwaters
bears his name). In its simplest form, a Jarlan-type breakwater
consists of a perforated front wall and a vertical impermeable 2.1. Fully perforated Jarlan-type structures
back-wall. (A vertical plate member is called a wall when it is
sitting on the bottom.) Since Jarlan first introduced his design of A fully perforated Jarlan-type structure typically consists of a
the Jarlan-type breakwater, different designs have been proposed vertical front wall perforated along the entire length of the plate,
over the years to improve the hydrodynamic efficiency and to a solid back-wall, and a wave absorbing chamber in between, as
better control the scouring at the toe of the structure. The first shown in Fig. 1(a). For a flat seabed the water depth/wavelength
scientific account of the slotted breakwaters (sometimes called inside the wave chamber is the same as that outside. For this type
slotted wave screens or pile breakwaters) was given by Wiegel of breakwaters, two dominant factors controlling the wave
(1960), even though this type of breakwaters has been in use reflection characteristics are the geometrical porosity of the front
since ancient times. Slotted structures and perforated structures wall and the relative wave chamber width B/L, where B is the
are sometimes used interchangeably in the literatures since both chamber width and L is the incident wavelength.
of them share a similar energy-dissipation mechanism (see Mei, Using linear potential theory and assuming normal incident
1983). The main advantages of the slotted/perforated coastal plane waves, Sahoo et al. (2000) derived the following explicit
structures are the saving in construction cost in relatively deep expression for the reflection coefficient:
water and less disturbance to coastal water environments. This  
 1Gð1icotðkBÞÞ  pffiffiffiffiffiffiffi e
review attempts to summarize the recent research findings on CR ¼  , i ¼ 1,G ¼ ð2:1Þ
1 þ Gð1 þ icotðkBÞÞ kdðf isÞ
perforated/slotted coastal structures.
We shall review a class of perforated/slotted coastal structures where k the wavenumber defined by k¼2p/L and G is the porous-
from a hydrodynamic point of view, with a focus on the reflection effect parameter of the perforated wall (Chwang, 1983; Yu, 1995).
and transmission characteristics. The wave forces on some of e is the front wall porosity and d is the wall thickness. Dimension-
these structures are also reviewed. However, the construction and less parameter f is a linear porous resistance coefficient and s is
stability of such structures are not covered in this review. Also, related to the wall inertial effect. The derivation of the porous-
the porous breakwaters are not reviewed here even though effect parameter G can be found in Section 4.1.2. When the
they are very similar to perforated/slotted breakwaters. The inertial effect of the front wall is not considered, the porous-
remainder of the paper is structured as follows. The next two effect parameter G is real; when the inertial effect of the front
sections discuss various types of perforated/slotted structures wall is considered, G is complex. Typical variations of reflection
that have been extensively researched, which is followed by a coefficient with the relative chamber width B/L are shown in
section on energy dissipation mechanisms for surface-piercing Fig. 2. For a real G, Eq. (2.1) says that the reflection coefficient
Z.H. Huang et al. / Ocean Engineering 38 (2011) 1031–1053 1033

Perforated Partially
front wall perforated wall
Solid back wall
h1

h h
B Wave absorbing
chamber

Multiple perforated walls Top cover

Sc

B1 B2
h h B
B

B
Horizontal porous plate

Rock core h1
h1

h h
B

Fig. 1. Definition sketches for various Jarlan-type perforated wall breakwaters.

takes its minimum values at


1
B=L ¼ 0:25 þ 0:5n, n ¼ 0,1,2,. . . ð2:2Þ

0.8 Due to the limitation in chamber width, only the fundamental


resonant mode (i.e., n ¼0) is of engineering interest when design-
ing Jarlan-type structures.
0.6 0.3 When the reflection coefficient is at a minimum, the incident
CR

waves pass through the front perforated wall without changing


0.4 2.0 its wave phase, while the waves reflected by the impermeable
back-wall and the incident waves at the front wall have opposite
0.2 phases (Chwang and Dong, 1984; Fugazza and Natale, 1992). For a
1.0 + 0.3i G = 1.0 real G, the reflection coefficient attains its minimums at the
values of B/L slightly larger than those for a complex G of the
0 same real part. When the reflection coefficient is at one of its
0 0.25 0.5 0.75 1 1.25 1.5
minimums, the surface elevations on both sides of the front plate
B/L
are found to have the largest difference, resulting in the largest
Fig. 2. Variations of the reflection coefficient for a fully perforated Jarlan-type horizontal wave force on the front wall. In other words, the
structure versus B/L at kh ¼1.3. variation of the reflection coefficient with B/L is opposite to that
1034 Z.H. Huang et al. / Ocean Engineering 38 (2011) 1031–1053

of wave force acting on the front wall (Yip and Chwang, 2000). We inside and outside the breakwater will generate strong evanes-
remark that the inertial effects and the effects of the wave cent waves, which in turn will contribute to the resonance inside
nonlinearity and evanescent waves may cause the minimum the chamber. To obtain a smaller reflection coefficient, the
reflection coefficients to occur at values of B/L slightly different geometrical porosity of the front wall (or the porous effect
from those calculated by Eq. (2.2). According to the published parameter G) should increase with decreasing internal water
experimental data for fully perforated Jarlan-type structures depth. Generally, it is possible to design a perforated breakwater
(Kondo, 1979; Bennett et al, 1992; Ou Yang et al., 1997; Zhu with both a smaller wave absorbing chamber width (low con-
and Chwang, 2001, among others), the fundamental resonant struction cost) and low reflection coefficients by choosing suitable
mode generally occurs in the range of 0.2 oB/Lo0.25. In parti- h1/h and G. Note that the low reflection coefficients can only be
cular, if the thickness of the front wall is large compared to achieved in a narrow range of B/L1 when h1/h is small, which
the water depth, the fundamental resonant mode may occur at makes the designed breakwater less efficient for broad-
B/Lo0.2 (Kakuno et al., 1992), possibly due to the significant banded waves.
inertial effects of the thick perforated front wall. This has also The inertial effects and the effects of strong evanescent waves
been indicated by Zhu and Zhu (2010), who found that the on the minimum reflection coefficients can also be found in
minimum reflection coefficients of a Jarlan-type breakwater could the literature. Tanimoto and Yoshimoto (1982) found that the
occur at smaller values of B/L for perforated walls of large minimum reflection coefficient occurred at B/L1 ¼0.15 0.20 for
thicknesses. h1/h ¼0.83 0.33, where L1 is the internal wavelength. Suh et al.
For a fixed B/L, the reflection coefficient is controlled mainly by (2006b) observed that the minimum reflection coefficient
the geometrical porosity of the front wall, and an optimum occurred at about B/L1 ¼0.20 (B/L is about 0.177) for h1/h¼0.65.
porosity exists to maximize the wave energy dissipation by the Li et al. (2002) and Liu et al. (2007d) found that the value of B/L for
front wall (Yu and Chwang, 1994), thus to minimize the reflection the minimum reflection coefficient was about 0.15 for h1/h ¼0.5,
coefficient as shown in Fig. 2. For perforated walls, it turns out and they also observed that for a fixed value of h1/h¼0.5, the
that for normally incident waves the perforation shapes (e.g., slits, reflection coefficients were slightly smaller for the front wall
circular holes, rectangular holes, etc.) have no significant effects porosity of 40% than for the porosity of 20%. For practical designs,
on the measured reflection coefficients (Park et al., 1993; Li et al., the values of h1/h, B/L and the geometrical porosity should be
2006; Cho and Kim, 2008). However, no definitive conclusion can carefully chosen so as to achieve both the structure stability and
be drawn for obliquely incident waves on the effects of the lower reflection coefficients for a wide range of wave frequency.
perforation shapes from the results published so far. For a fully
perforated Jarlan-type structure with a flexible perforated front 2.3. Multiple perforated front walls
wall, the values of B/L at which the reflection coefficient has its
local minimums increase slightly with increasing flexibility, while Two or more perforated walls (see Fig. 1(c)) may be used to
the corresponding wave force acting on the front wall decreases further reduce the reflection coefficient of perforated breakwaters
with increasing flexibility (Wang and Ren, 1994). (Sawaragi and Iwata, 1978; Kondo, 1979; Fugazza and Natale,
1992; Williams et al., 2000; Bergmann and Oumeraci, 2000; Chen
2.2. Partially perforated Jalan-type structures et al., 2002; Li et al., 2003a; Huang, 2006, among others). Caisson
breakwaters with three perforated walls have been constructed in
In practice, due to the requirements on structural stability, the Porto Torres industrial harbor, Italy (Franco, 1994; Franco
perforated wall breakwaters are generally constructed in a form et al., 1998). Caisson breakwaters with five perforated walls have
of partially perforated caissons, as shown in Fig. 1(b), where the been designed and constructed for the Dalian Chemical Produc-
inner water depth is h1 and the outer water depth is h. For such tion Terminal, China; Fig. 4(a) shows the design of the cross
structures the water depth inside the chamber is smaller than section of this breakwater and Fig. 4(b) is a view of the five-
that outside. Typical variations of reflection coefficient with the perforated-wall breakwater under construction.
relative chamber width B/L1, where L1 is the wavelength inside For a fully perforated breakwater with two perforated front
the chamber, are shown in Fig. 3 for three values of h1/h and two walls (Li et al., 2003a), the typical variations of reflection coeffi-
values of the porous effect parameter G (Li et al., 2002). When cient with the relative total chamber width is shown in Fig. 5,
h1/h decreases, the minimum reflection coefficient will occur at where G1 and G2 are the porous effect parameters of the first and
smaller values of B/L1: the abrupt change of the water depths second perforated walls, respectively; B1 and B2 are the widths of
the first and second wave absorbing chambers, respectively.
1 B1 ¼B2 ¼B/2 was taken in Fig. 5. The addition of the middle
perforated wall has little effects on long waves (or small B/L).
For a perforated single chamber breakwater with a small geome-
0.8
h1/h=0.2 G=1.0 trical porosity (G1 ¼ 0.5), the addition of a middle perforated plate
does not significantly change the reflection coefficient for
0.6 B/Lo0.1, but it can significantly reduce the reflection for short
waves (B/L40.4) and increase the reflection of the intermediate
CR

waves (0.1 oB/Lo0.4). However, when both the geometrical


0.4
porosity of the first wall (G1 ¼1.5) and B/L are large, a middle
perforated wall can significantly reduce the reflection coeffi-
0.2 h1/h=0.5 G=1.0 cients; these predicted results have also been observed in other
experimental tests for breakwaters with two fully perforated
h1/h=0.2 G=2.0 h1/h=1.0 G=1.0 walls (Sawaragi and Iwata, 1978; Kondo, 1979) or breakwaters
0
with two partially perforated walls (Chen et al., 2002). The tests of
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
Sawaragi and Iwata (1978) indicated that the reflection coeffi-
B/L1
cients of a double perforated wall breakwater with a uniform
Fig. 3. Variations of the reflection coefficient for a partially perforated Jarlan-type porosity of 20% were smaller than those of a single chamber
structure versus B/L1 at kh ¼1.3. structure with the same porosity. In the tests of Kondo (1979), a
Z.H. Huang et al. / Ocean Engineering 38 (2011) 1031–1053 1035

Fig. 4. The perforated breakwater with five perforated walls built for Dalian chemical production terminal, China (Courtesy of Mr. Zhongliang Shi, the Chief Engineer of the
Project): (a) The design cross section of the breakwater and (b) a view of the breakwater under construction.

perforated wall with porosity of 20% was installed inside the when the ratio of the water depth to the incident wavelength
chamber of a breakwater with a single perforated wall of porosity varied from 0.05 to 0.5. Under optimum conditions, the theore-
34%; the addition of the middle perforated wall significantly tical reflection coefficient could be as low as 0.04 (Twu and Lin,
reduced the reflection coefficients compared to the original 1991).
structure. Chen et al. (2002) found that adding a middle perfo-
rated wall was helpful in enhancing the structure’s wave absorb- 2.4. Perforated caissons with top cover plate
ing ability when the porosity of the front perforated wall was 40%.
However, the opposite results were observed when the porosity Very often a top cover is constructed on a perforated caisson
of the front perforated wall was 20%. Therefore, caution needs to structure to provide areas for other uses (see Fig. 1(d)). Researchers
be taken when designing breakwaters with two or more in Dalian University of Technology (DUT) have conducted a series
perforated walls. of experimental and numerical studies on this topic (Ma, 2004;
Structures with multiple perforated walls can also be used as Jiang, 2004; Li, 2007; Chen et al., 2007). In their numerical studies,
highly effective wave absorbers in wave flumes or basins (Evans, the equation of the state for air under constant-temperature
1990; Twu and Lin, 1991; Losada et al., 1993, among others). conditions was used in order to simulate the compressibility
According to Twu and Lin (1991), the wave absorbing ability of of the air inside the chamber. When the wave crests interact with
such structures is mainly controlled by the spacing between the a perforated caisson with a top cover, the holes in the perforated
neighboring porous walls and the arrangement of wall porosities: front wall are filled with water, resulting in a pressure increase in
the geometrical porosities of the perforated walls should be the air inside the chamber. Their studies showed that (i) a top
arranged in a gradually decreasing order in the direction of the cover would increase the reflection coefficient of the perforated
incident waves. The optimum ratio of the spacing between the caissons, and (ii) the reflection coefficient generally would decrease
neighboring porous walls to the water depth was found to be 0.88 with increasing Sc, the spacing between the top cover and the still
1036 Z.H. Huang et al. / Ocean Engineering 38 (2011) 1031–1053

1 1
With a surface-piercing
rock core
0.8 G1=1.5 G2=Infinity 0.8

G1=0.5 G2=1.0
0.6 0.6
CR

CR
0.4 0.4
G1=0.5 G2=Infinity
With a submerged
0.2 0.2 rock core

G1=1.5 G2=1.0 B1=B2=B/2 Without rock core


0 0
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
B/L B /L

Fig. 5. Variations of the reflection coefficient for a Jarlan-type structure with two Fig. 6. Variations of the reflection coefficient with B/L for various perforated
perforated front walls versus B/L at kh¼ 1.3. breakwaters with/without rock core with kh¼ 1.3, G ¼ 1.0, e ¼ 0.45, f¼ 2.0, and
s¼ 1.0.

water level, and (iii) the compressibility of the air in the chamber
needs to be considered in the numerical modeling. It should be
noted that no slamming impact force was observed on the top 2.6. Perforated structures with internal horizontal plate
cover plate in all experimental tests of Chen et al. (2007). The ratio
of the reflection coefficient with a top cover to that without is Yip and Chwang (2000) studied a modified Jarlan-type break-
about 1.0–1.3 for regular waves with Sc/H¼ 0.333 2.0, and 1.0–1.2 water, which consisted of a fully perforated front wall, an
for irregular waves with Sc/Hs ¼0.52.0, where Hs is the significant impermeable rear wall and a submerged horizontal solid plate
wave height. in the chamber. They found that the hydrodynamic performances
of the perforated wall breakwaters with and without an internal
horizontal solid plate were similar. One of the advantages of a
2.5. Perforated breakwaters with rock core perforated structure with an internal plate is that the breakwater
can be constructed with a small wave absorbing chamber, with
Isaacson et al. (2000) studied analytically the performance of a the hydrodynamic performance being similar to that without
fully perforated Jarlan-type structure with a surface-piercing rock the internal plate. This is equivalent to reducing the water
core, which had been used in replacement of an old breakwater at depth inside the chamber discussed preciously. Yip and Chwang
Kelsey Bay, Canada. The surface-piercing rock core could increase (2000) also pointed out that the internal plate could reduce the
the reflection coefficient, and reduce the wave force acting on the wave force and moment acting on the perforated front wall,
front perforated wall. which can be very helpful for enhancing the stability of the
Liu and Li (2006a) proposed a further modification to the front wall.
structures studied by Isaacson et al. (2000). They filled the chamber Liu et al. (2007c) proposed another modified Jarlan-type
with a submerged rock core, as shown in Fig. 1(e). Their studies breakwater with an internal submerged horizontal porous plate
revealed that increasing the thickness of the rock core could increase as shown in Fig. 1(f). According to their analytical results, the
the reflection coefficient, but decrease the wave force acting on the internal horizontal perforated plate could significantly enhance
front perforated wall. Therefore, if the wave force acting on the the wave absorbing ability of the perforated wall breakwater.
perforated wall is the main concern of a design, filling the chamber With a suitable design (in general, the front wall porosity should
with a surface-piercing rock core is recommended; if the main be relatively large and the porosity of the horizontal plate should
concern is to have both a smaller wave force acting on the front wall be moderate), it is possible to achieve very small reflection
and a lower reflection coefficient, a submerged rock core in the coefficients for B/L greater than certain small critical value
chamber is encouraged. Liu et al. (2007b) examined the performance (however, more experimental research is needed to quantify this
of a perforated wall breakwater with two layers of rock cores (the critical value). This is very useful from an engineering point of
upper layer pierces the water surface), and found that the perfor- view since the value of B/L may vary significantly in practice. Liu
mance of the perforated breakwaters with two layers of rock cores et al. (2007c) also found that perforating the internal horizontal
was similar to that with a single surface-piercing rock core. plate could significantly reduce the vertical wave force and the
Typical variations of the reflection coefficients with relative moment on the horizontal plate, which is also very helpful for
chamber width B/L for a perforated breakwater with a submerged enhancing the breakwater’s stability.
rock core (Liu and Li, 2006a) are shown in Fig. 6 for normal incident Typical reflection coefficients for breakwaters with/without
regular waves, with a submergence of the rock core being 20% of the internal horizontal plates are shown in Fig. 7 for regular waves
water depth. It should be stressed that the change of B/L in Fig. 6 normal to the breakwaters, where the porous-effect parameter
was achieved by varying the chamber width B with the wavelength G¼ 2.0 was used for the front perforated wall. The internal
fixed. The submerged rock core could significantly increase the horizontal perforated plate has a submergence that is 20% of the
reflection coefficients for wide chambers, but may have insignificant water depth, and the porous-effect parameter G ¼0.5. An internal
effects for narrow chambers. A surface-piercing rock core can further solid plate will reduce the wavelength inside the chamber, which
increase the reflection coefficient for almost all chamber widths, makes the minimum reflection coefficients occur at smaller
except for extremely narrow chambers. Note that the wavelength values of B/L. An internal porous horizontal plate also dissipates
inside the chamber is affected by the rock core, which results in wave energy in addition to reducing the wavelength inside the
a shift of the values of B/L at which the minimum reflection chamber, as a result, low reflection coefficients can be achieved
coefficients occur. when B/L exceeds certain small value.
Z.H. Huang et al. / Ocean Engineering 38 (2011) 1031–1053 1037

2.7. Wave direction, wave irregularity, wave steepness, and et al. (2009) derived the following explicit expression for the
rubble foundation reflection coefficient
 
 cos yGð1i cotðkB cos yÞÞ 
2.7.1. The direction of incident waves CR ¼   ð2:3Þ
The energy loss associated with water flowing through the cos y þ Gð1þ i cotðkB cos yÞÞ
perforated wall is directly related to the velocity component which is similar to Eq. (2.1). For oblique regular waves, the
normal to the breakwater. For a fixed B/L, the normal velocity wavenumber in the direction normal to the breakwater is k cos y,
decreases with increasing wave angle (the angle between the and the corresponding wavelength is L/cos y. The minimum
wave crest and the breakwater), resulting in a decrease in energy reflection coefficients will occur at B cos y/L¼0.25þ0.5n (n ¼0,
loss. When the long-crested waves are parallel to a perforated 1, 2,y) after ignoring the inertial effects associated with the
structure, no energy dissipation is expected. For incident waves perforated wall. Fig. 8 shows the typical reflection coefficients
propagating at an angle y to a fully perforated breakwater, Liu calculated from Eq. (2.3) for kh ¼1.3 and G ¼1.0. For a fixed
B cos y/L, the reflection coefficient increases with increasing
1 incident wave angle y. The perforated caissons are usually
prefabricated with transverse sidewalls in practice, which will
complicate the wave motion inside and outside the caisson
0.8
chambers and cause the reflected waves to leave the breakwaters
in different directions (Teng et al., 2004; Liu et al., 2007a).
0.6 Interested readers can consult Li et al. (2002, 2003a) and Suh
CR

et al. (1995) for more details on oblique wave reflections from


perforated breakwaters without transverse sidewalls.
0.4

0.2 2.7.2. Wave irregularity


With internal With internal Without For irregular waves, a frequency-averaged reflection coeffi-
solid plate porous plate internal plate cient can be defined by (Goda, 2000)
0
0 0.2 0.4 0.6 0.8 1 rffiffiffiffiffiffiffiffiffiffi
m0,R
B /L CR ¼ ð2:4Þ
m0,I
Fig. 7. Reflection coefficients for perforated wall breakwaters with/without
internal horizontal plates. Modified from Liu et al. (2007c) for kh¼ 1.3.
where, m0,R and m0,I are the zeroth moments of the reflected and
incident wave spectra, respectively. Liu et al. (2007d) made a
comparison between irregular and regular waves for partially
1
perforated caissons by defining an equivalent regular waves with
T  Tand HEH1%, where Tand H1% are, respectively, the mean
0.8 wave period and the wave height whose exceedance probability
of occurrence in a train of irregular waves is 1%. (H1% is slightly
0.6 different from the average height of the highest 1% waves in a
train of irregular waves.) As shown in Fig. 9, the reflection
CR

θ coefficients of irregular waves are generally larger than those of


0.4 60
ο
regular waves under practical conditions. Suh et al. (2001) and
ο
45 Ketabdar and Varjavand (2008) studied the spectrum of the
30ο
0.2 0ο reflected waves and found that the reflected wave spectrum could
have multiple peaks and the energy density function might be
close to zero at some frequencies. If the interactions between
0
0 0.25 0.5 0.75 1 1.25 1.5
different component waves are ignored, the reflected wave
B cos θ / L spectrum can be obtained from the results for regular waves
using a linear transfer function between the incident and
Fig. 8. Variations of the reflection coefficient for oblique waves versus B cosa/L at reflected waves. For a large wave chamber with its width fixed,
kh ¼1.3 and G ¼ 1.0. the reflection coefficient of each component wave may vary in

3.5
3 20% porosity

2.5 40% porosity


R
⎯C / C

2
R

1.5
1
0.5
0 0.05 0.1 0.15 0.2 0.25 0.3

B /L ( L )

Fig. 9. Comparison of measured reflection coefficients between regular and irregular waves (after Liu et al., 2007d).
1038 Z.H. Huang et al. / Ocean Engineering 38 (2011) 1031–1053

a wide range with wavelength and near-zero reflection coeffi-


cients may be found under certain wave conditions.

2.7.3. Wave steepness


The experimental results of Park et al. (1993) and Suh et al.
(2006b) showed that increasing wave steepness led to a reduction
in the reflection coefficients of perforated caissons, due mainly to
the increased energy dissipation at the perforated front wall
when waves are steep. Ma (2004) found from his experiments
that the effects of the wave steepness were not significant for
regular waves, and noticeable for irregular waves: wave reflection
coefficients slightly decrease with increasing irregular wave
steepness defined by Hs/Ls, where Hs and Ls are the significant
wave height and the significant wavelength, respectively. It seems
that the effects of wave steepness on the reflection coefficient
may not be of practical significance.

2.7.4. Rubble foundation


Perforated caissons are usually built on a rubble filled founda- Fig. 10. A bottom-sitting, surface-piercing pile breakwater along the coast of
tion or a rubble mound foundation (see Fig. 4(a) for a typical Singapore.

rubble mound foundation design). For a rubble mound founda-


tion, the numerical study of Suh et al. (1995) indicated that the
presence of the rubble mound foundation would slightly increase This section will review some recent works on single slotted
the reflection coefficient under most practical conditions, except breakwaters, double slotted breakwaters, and some variants.
for relatively high mound foundations with steep mound slopes. Several slotted breakwaters reviewed in this section are shown in
The effects of a rubble filled foundation on the reflection coeffi- Fig. 11, where 5 main types are included for the convenience of
cients are insignificant for most practical designs and wave presenting the review materials in this section. Fig. 11(a) shows an
conditions, and thus the studies on perforated caissons sitting original bottom-sitting and surface piercing pile breakwater.
on various rubble filled foundations should focus on the vertical Fig. 11(b) and (c) is two major variants: the suspended slotted
forces acting on the structures (Liu et al., 2006b, 2008). We barrier suitable for deep water conditions and the curtain-wall-pile
remark that future research is needed to understand the effects breakwater, which can further reduce the transmission coeffi-
of the wave energy dissipation due to the motion of water cients. Fig. 11(d) is a sketch of a generic multi-layer surface-
through the pores in the rubble foundation. piercing slotted breakwater, where each wall may have an
impermeable portion (a curtain) of length hf and a slotted/perfo-
rated portion of length hs; each wall may or may not sit on the
3. Slotted barriers without back-wall (wave screens) seabed—if the wall is not bottom-sitting, the bottom clearance
hc 40. If a slotted breakwater has more than one layer, the distance
Near-shore circulations are important to a healthy eco-system between two adjacent layers, B, will be an important parameter
in harbors or marinas. Introducing a natural circulation into a determining the hydrodynamic characteristics of the breakwater.
deteriorating harbor can help improve the water quality inside Fig. 11(e) shows a typical multi-layer submerged perforated
the harbor. To keep the water in harbors from stagnating breakwater.
pollutants, perforated/slotted breakwaters without back-wall
(pile breakwaters, slotted breakwaters, submerged wave screens, 3.1. Surface-piercing slotted breakwaters
etc.), which can be in the form of either one row or multiple rows
of closely spaced cylinders without a back-wall, should be 3.1.1. Regular waves
considered for those harbors or marinas where certain degree of 3.1.1.1. Single-layer slotted wave screens. There is a rich literature
wave activities is permitted inside the harbors. The advantages of in the wave scattering by a single-layer wave screen (a row of
the breakwaters in this group include possible improvement of closely spaced cylinders, see Fig. 11(a)). For long waves, Mei et al.
water circulation inside the harbors, fish passage, sediment (1974) showed that the transmission and reflection coefficients
transport across the breakwaters, and low construction cost when for a single-layer wave screen in a water of depth h could be
the water is relatively deep (Mani and Jayakumar, 1995). Such calculated by
breakwaters have been constructed at several harbors; for exam- U0 o U0
ple, Baie Comeau Harbor and Chandle Harbor in Canada, Roscoff CT ¼ pffiffiffiffiffiffi , CR ¼ 1CT ¼ 1 ð3:1Þ
ðA=hÞ gh gk A
Harbor in France, Half Moon Bay Marina in New Zealand, and
Plymouth harbor in the United States (Mei et al., 1974; where o is the wave angular frequency, and the characteristic
Hutchinson and Raudkivi, 1984; Gardner et al., 1986, etc.). velocity U0 is related to the linearized friction coefficient f and the
Fig. 10 is a photo of a section of a pile breakwater along the coast wave amplitude A by
of Singapore. Some other earlier attempts to understand wave qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
transmission and wave energy loss through vertical slotted/ A pffiffiffiffiffiffi 12bf 1 4 A
U0 ¼ gh and bf ¼ f ð3:2Þ
perforated breakwaters, which are not reviewed in detail here, h bf 3p h
can be found in Terret et al. (1968), Hayashi et al. (1968), Grune
and Kohlahas (1974), Kakuno (1983), Urashima et al. (1986), and The relationship between CR and CT in Eq. (3.1) was derived based
Kriebel (1992). The wave scattering by arrays of cylinders without on the continuity of velocity across the slotted/perforated wall.
energy loss (Martin and Dalrymple, 1988; Linton and Evans, 1990, The theory compares reasonably well with the experiments of
etc.) will not be covered in this review. Hayashi et al. (1966) for a single row of circular cylinders; the
Z.H. Huang et al. / Ocean Engineering 38 (2011) 1031–1053 1039

A bottom-sitting pile breakwater A suspended slotted barrier

A curtain-wall-pile breakwater A multi-layer surface piercing slotted barriers

A multi-layer submerged perforated barriers

Fig. 11. Definition sketches for various slotted/perforated barriers without back-wall.

observed differences between the theoretical and experimental CR and CT, reasonably well, except for wave screens with very
results are due possibly to the fact that some waves in the small porosities
experiments might not be long waves. For single-layer wave    
 2G   
screens, the higher harmonic waves generated by the quadratic CR ¼  , CT ¼  1  ð3:3Þ
1 þ 2G   1þ 2G
head loss have little effects on the predicted reflection and
transmission coefficients of the fundamental waves (Mei et al,
1974). Rectangular cylinders may help dissipate more wave The reflection coefficient increases with decreasing barrier porosity,
energy (Huang, 2007b)—the enhanced energy loss is due to the while the transmission coefficient decreases with decreasing barrier
flow separation around the sharp edges of the rectangular bars. porosity. Also see Yu (1995) for the use of Eq. (3.3) in the wave
Treating the single-layer wave screen as a porous thin wall of scattering by porous walls.
a porous-effect parameter G, Isaacson et al. (1998) found that The reflection and transmission coefficients depend strongly
Eq. (3.3) could predict the reflection and transmission coefficients, more on the porosity of the wave screen than on wavelength or
1040 Z.H. Huang et al. / Ocean Engineering 38 (2011) 1031–1053

height. Reducing the porosity of the wave screen will increase showed that the second barrier had negligible effects on the
wave reflection but reduce the wave transmission (Isaacson et al, waves scattered by the wave screens if the porosity of the second
1998, Huang, 2007a). Energy loss is theoretically zero when either barrier was greater or equal to 0.4; when the porosities of both
the porosity e ¼ 0 or 1, and a maximum energy loss is expected at slotted barriers were less than 0.4, the transmission coefficients
certain porosity between 0 and 1. were not sensitive to both the chamber width and the wave
Under deep water conditions, wave motion near the bottom is period, but controlled mainly by the porosities of the two barriers.
negligible. In this case, suspended wave screens, as shown in As shown in Fig. 12, the reflection coefficients, however, have a
Fig. 11(b), may be preferred for an economic design (Mani and strong dependence on both the chamber width and the wave
Jayakumar, 1995). As far as the measured transmission coeffi- period, with the minimum reflection coefficient being located
cients are concerned, Mani and Jayakumar’s experiments showed generally around B/L¼0.25 and the maximum reflection coeffi-
that a suspended slotted wave screen with a draft to depth cient around B/L¼0.5 (for given wave conditions). For fixed wave
ratio of 0.46 could perform equally well in the range of periods, larger waves tend to produce smaller transmission
0:002 o HI =gT 2 o0:02 when compared with the bottom-sitting coefficients and slightly larger reflection coefficients, meaning
slotted screens (draft to depth ratio is equal to 1). The effects of that the energy dissipation increases with incident wave height
the relative draft (hs/h) has been studied by Isaacson et al. (1998), (Huang, 2007b). By optimizing the chamber width and the barrier
who, using the porous effect parameter G and the eigen-function porosities, it is possible to design a double slotted wave screen
expansion method, devised a theory for suspended single slotted with a near-zero reflection coefficient and a transmission coeffi-
wave screens and compared their numerical results with their cient around 0.2 for most practical wave conditions (Hagiwara,
experiments. 1984).
Recently, Krishnakumar et al. (2010b) carried out an extensive
3.1.1.2. Curtain-wall-pile breakwater. Curtain-wall-pile breakwaters, experimental study of single- and double-wave screens in the
which have impermeable potions near the water surface (see form of a series of horizontally placed circular bars. The double
Fig. 11(c)), have been proposed to improve the performance of tra- wave screens are combinations of bottom-sitting and suspended
ditional slotted breakwaters without sacrificing in water exchange slotted barriers with difference submergences. Based on their
across the breakwaters (Suh et al., 2006a, 2007). As wave energy experiments, design formulae have been derived for the reflection
cannot flux through the impermeable part of the vertical wall, and transmission coefficients and the pressures on the slotted
transmission coefficients/reflection coefficients can be effectively barriers of single- and double slotted breakwaters.
reduced/increased by the curtain in comparison with the traditional
slotted breakwaters. Even though the impermeable portion will
have significant effects on the reflection and transmission coeffi-
cients in a wide range of wave frequency, the influences are more
profound for short waves: if the draft of the curtain is more than one
half of the length of short waves, no wave energy can be transmitted
to the other side of the breakwaters. The impermeable portion will
take much of the wave force acting on the breakwater, thus the
overturning moment can be significantly larger than that on the
traditional slotted breakwaters, especially when such breakwaters
are to be used in relatively deep water.
Recently, Rageh and Koraim (2010) investigated, theoretically
and experimentally, the screen-type breakwaters with horizontal
slots and a curtain (an impermeable part near the water surface).
Reflection, transmission and energy dissipation were studied. In
particular, the effects of the impermeable part on the transmis-
sion coefficients were discussed; as expected, increasing the draft
of the impermeable part would decrease the transmission coeffi-
cients. Rageh and Koraim (2010) also summarized the experi-
mental conditions of experiment facilities and the ranges of key
parameters found in the literature they reviewed, which included
some works published in Egyptian sources.
Besides curtain walls, alternative upper structures supported
by piles have also been proposed. For example, Sundar et al.
(2002) studied a breakwater consisting of closely spaced support-
ing piles and an upper structure with a seaside quadrant solid
front face. Both the reflection coefficient and the wave force
acting on the breakwater were carefully studied in their
experiments.

3.1.1.3. Double-layer slotted wave screens. For a double slotted


wave screen (two rows of closely spaced cylinders or two slotted
barriers separated by certain distance), there is a chamber between
the two slotted wave screens, which may result in a resonant
condition under which the energy dissipation may be maximized
and the transmission coefficient may be minimized.
For bottom-sitting double slotted wave screens, both the
experimental results (Kono and Tsukayama,1981; Huang, 2007b) Fig. 12. Variations of reflection and transmission coefficients with wave height
and the theoretical results (Hagiwara,1984; Huang, 2007b) (a) and chamber width (b). After Huang (2007b).
Z.H. Huang et al. / Ocean Engineering 38 (2011) 1031–1053 1041

3.1.1.4. Multi-layer slotted wave screens. It is possible to further 2006a) if the head loss coefficient of Mei (1983) was used to
improve the performance of the slotted wave screens by including calculate CR(f),CT(f) in Eq. (3.4).
more slotted barriers (see Fig. 11(d)). Kakuno and Nakata (1998)
employed the concept of complex blockage coefficient to derive a 3.1.3. Effects of near-shore currents
linear theory, based on a plane wave approximation, for the scat- There is always a complicated current system near-shore,
tering of regular waves by multi-layer slotted barriers with/without which may affect the performance of the slotted breakwaters.
a back-wall. Their theory predicted the reflection and transmission Experiments for single slotted breakwater in the presence of a
coefficients reasonably well when compared with the experiments current (Huang, 2007a) showed that the current could signifi-
of Hagiwara (1984) for three layers of slotted barriers, which cantly reduce the transmission coefficients but only slightly
showed that a minimum transmission coefficient of 0.55 could be increase the reflection coefficients (see Fig. 13), resulting in an
achieved without sacrificing the reflection coefficients. Recently, increase in wave energy dissipation. The transmission coefficient
Shepsis et al. (2007) reported two case studies of multi-slotted wall was very sensitive to the current strength, but less sensitive to the
breakwaters, including a numerical simulation of waves inside a variations in wave period or chamber width. The increased energy
harbor protected by a slotted breakwater system. More recently, loss is due mainly to the increased drag forces on the cylinders of
Ji and Suh (2010) examined the performance of multiple-row cur- the slotted barriers (Huang and Ghidaoui, 2007). For double
tain wall-pile breakwaters by means of analytical and experimental slotted breakwaters in combined wave-current fields (Huang,
methods. They found that the multi-row curtain wall-pile break- 2008; Huang and Liu, 2008), it seems that a weak current
waters can significantly reduce the transmission coefficients com- may not significantly change the conditions under which the
pared to a single row one, but the difference between two rows and maximum reflection coefficient and the minimum reflection
three rows was marginal. Krishnakumar et al. (2010a) measured the coefficient occur.
pressure and wave forces on a partially immersed solid wall pro- Inside harbors, water level changes periodically with local
tected by one or two slotted walls. They found that the wave tides, producing a current flowing through the slotted break-
pressures and forces on the solid wall decreased significantly with waters. The current, even though weak in most cases, is not only
increasing angle of incident waves. important to maintaining the water quality inside the harbor but
also beneficial to preventing ocean waves from entering the
harbor through slotted breakwaters.
3.1.2. Irregular waves
Ocean waves are random in nature. The power spectra of the
3.2. Submerged slotted/perforated breakwaters
reflected and transmitted waves, SZ,R,SZ,T, can be estimated,
respectively, by
Compared to the surface-piercing slotted/perforated break-
2 2
SZ,Rðf Þ ¼ 9CR ðf Þ9 SZ,I ; SZ,Tðf Þ ¼ 9CT ðf Þ9 SZ,I ð3:4Þ waters, not much research has been done for submerged
slotted/perforated breakwaters. This type of breakwaters might
where SZ,I is the spectral density of the incident waves. The reflection be preferable for the purpose of erosion controls without adverse
and transmission coefficients for the wave component of frequency f impacts on the landscape and environment along recreational
are CR(f) and CT(f),which can be calculated using one of the methods coasts (Oumeraci, 2010). Clauss et al. (2000) explored the idea
described in Section 4.2. After the irregular waves passing through the of using so-called submerged multi-layer systems of permeable
wave screen, the shape of the wave spectrum might be changed, vertical walls for coastal protection. Oumeraci and Koether (2009)
especially the peak frequency and the peak magnitude. Due to energy systematically studied the submerged multi-filter systems, which
dissipation, the peak magnitude will be reduced. But there is no are essentially a series of submerged permeable walls in the form
definite conclusion in the literature on the shift of the peak frequency of horizontal slotted barriers (see Fig. 11(e)); they have studied
of the transmitted waves: the experiments of Isaacson et al. (1998) the wave reflections, transmissions and energy losses for regular
showed a slight downshift of the peak frequency of the transmitted waves, irregular waves and solitary waves. Oumeraci and Koether
waves, which is consistent with the fact that long waves can easily go (2009) pointed out that the porous-effect parameter G, which has
through the slotted barriers than short waves do; however, the
experimental and numerical results of Park et al. (2000) showed a
slight shift of the spectrum peak of transmitted waves towards higher
frequencies. Suh et al. (2006a) examined the scattering of irregular
waves by curtain-wall-pile breakwater; JONSWAP spectrum was
used in their study. For the peak parameter g ¼1.0 a slight peak
frequency downshift was found, while no observable shift of peak
frequency was found for g ¼3.3 and 10.0. More work need to be done
to clarify this discrepancy.
For the scattering of irregular waves by slotted barriers
designed according to peak frequencies, the frequency-averaged
reflection and transmission coefficients are commonly defined by
rffiffiffiffiffiffiffiffiffiffi rffiffiffiffiffiffiffiffiffiffi
m0,R m0,T
CR ¼ , CT ¼ ð3:5Þ
m0,I m0,I

where m0,R,m0,T,m0,I are the zeroth moments of the reflected,


transmitted, and incident wave spectra, respectively. The fre-
quency-averaged reflection and transmission coefficients varied
with component wave frequency in a way similar to what regular
waves did (Isaacson et al., 1998; Park et al., 2000). The frequency- Fig. 13. Variations of wave reflection and transmission coefficients with wave
averaged reflection and transmission coefficients predicted by height in the presence of a weak current. Curves are theoretical predictions. After
Eq. (3.5) agreed generally well with those measured (Suh et al., Huang and Ghidaoui (2007).
1042 Z.H. Huang et al. / Ocean Engineering 38 (2011) 1031–1053

been commonly used for surface piercing slotted/perforated that this type of breakwaters may have potential applications in
walls, could not be used for submerged slotted walls since water using engineering measures to mitigate tsunami hazards.
not only can pass through the gaps of a slotted wall, but also can
flow over the crest of the submerged wall. Two types of wave
energy losses exist: (i) energy loss induced by the flow separation 4. Energy dissipation mechanisms perforated/slotted walls
at the crest of the submerged wall, and (ii) energy loss induced by
the water flowing through the gaps of the submerged slotted/ After having introduced two major types of coastal perforate/
perforated wall. Structures of one, two and three submerged walls slotted structures, we discuss in this section the energy dissipa-
were examined in their experiments. The main findings of their tion mechanisms for (i) waves passing through perforated/slotted
experimental study are summarized in this section; the energy vertical walls that pierce through the water surface and (ii) waves
dissipation mechanism for this type of structures will be dis- passing submerged slotted walls.
cussed in Section 4.1.4. Except within the very narrow regions adjacent to the gaps/
For a single submerged wall, the reflection and transmission holes of the slotted/perforated walls, the wave motion can be
coefficients can be described well by the so-called ‘‘Overall assumed irrotational and a velocity potential f can be defined by
Permeability Parameter (OPP)’’ defined by ! !
u ¼ rF, where u is the wave orbital velocity. In regions close to
Rc 1:4 the gaps/holes, flows are turbulent, and there is a pressure drop
OPP ¼ 2,Rc ¼ ðRc þ h0 Þe dB , e ¼ tanh ð2:8eÞ ð3:6Þ
þ
Zmax across the perforated/slotted wall due to the loss of wave energy.
In the rest of this section, we shall discuss various ways to model
where h0 ¼MWL  SWL (SWL stands for the still water level and
this pressure drop under various conditions. For simplicity, long-
MWL stands for the mean water level above the wall crest ), Rc is
crested waves normal to the wall will be used to demonstrate the
the submergence depth; dB is the wall height; e is the porosity of
basic principles.
the submerged wall, and Zmax þ
is the maximum upward surface
For surface-piercing slotted/perforated structures, there are two
elevation of the incident waves. OPP is negative for submerged
major methods to model the energy dissipation caused by water
walls. In terms of OPP, the measured reflection and transmission
waves passing through a slotted/perforated wall: (1) using an
coefficients correlated well with the following empirical expres-
energy loss coefficient if the perforated/slotted walls are treated
sions for all their tests:
as thin walls with orifices and (2) using a porous-effect parameter if
8 the perforated/slotted walls are treated as thin porous walls. For
CT ¼ arctan3 ð1:49OPP9Þ, CR  1CT ð3:7Þ
p3 submerged slotted/perforated structures, Oumeraci and Koether
For a single wall, the maximum wave energy dissipation (2009) generalized the porous-effect parameter to the so-called
coefficient Cd is estimated by ‘‘complex structure parameter’’, which includes the energy dissipa-
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi tion due to the vortex shedding at the crest of the submerged wall.
Cd ¼ 1CT2 CR2  2CT ð1CT Þ ð3:8Þ

From the above equation, the maximum value Cd E0.7 is obtained 4.1. Regular waves in the absence of currents
at CT E0.5, which can be realized when OPP¼ –2.12. The experi-
ments of Oumeraci and Koether (2009) show that the energy 4.1.1. Energy loss coefficient for surface-piercing
dissipation coefficient varies from 0.4 to 0.7 for all their tests on slotted/perforated walls
single wall structures. One of the widely used approaches to model the frictional
To further increase the energy dissipation coefficient, two- or effects of a perforated/slotted wall (or barrier) is to use a
three-submerged walls (with wall porosities decreasing in the quadratic loss coefficient. Mei et al. (1974) first derived an
wave direction) can be used. The experiments of Oumeraci and expression for the head loss of long waves at a slotted wall based
Koether (2009) show that the dissipation coefficients for two- or on the assumption that the wall is very thin compared to typical
three-submerged walls can reach 0.9 and both the reflection and wavelengths. Several authors (Kriebel, 1992; Bennett et al. 1992;
transmission coefficients can be effectively reduced so that Zhu and Chwang, 2001; Huang, 2006, 2007b, etc.) have extended
CT þCR o1. For multiple-wall structures, it is found that the the work by Mei et al. (1974) to transitional water waves.
relative chamber width and the wall height are two important Referring to Fig. 14, we assume that 9x þ  x  95Lw, where x 
design parameters. For high two-wall structures (each wall and x þ are coordinates immediately up-wave and down-wave of
occupies 98% or more of the still water depth), the minimum the slotted wall, respectively, and Lw is the wavelength. At the
wave transmission is found near the peak energy dissipation, elevation z, the pressure change across the slotted/perforated wall
which corresponds to B/LE0.25 (B is the chamber width); for low can be modeled by
two-wall structures (each wall occupies 79% or less of the still
Cf @uðx þ ,z,tÞ
water depth), the minimum transmission shifts to B/LE0.4. pðx ,z,tÞpðx þ ,z,tÞ ¼ r uðx þ ,z,tÞ9uðx þ ,z,tÞ9 þ rLg
2 @t
A comparison of three-wall structures with two-wall structures
ð4:1Þ
also shows that the overall damping performance of the two-wall
structures can be improved by introducing an intermediate wall where p is the dynamic pressure associated with the water waves
in the chamber. of a surface elevation Z, g the gravitational acceleration, r the
For irregular waves, the reflection and transmission coeffi- density of water, t the time, Cf a quadratic loss coefficient, and Lg
cients calculated using the spectral wave approximation approach an empirical length scale (see Zhu and Chwang, 2001). Note that
agrees well with those measured for two- or three-walls. The the pressure is hydrostatic for long waves, i.e., p ¼ rgZ, which can
reflection and transmission coefficients calculated using ‘‘regular reduce Eq. (4.1) to the form in Mei et al. (1974). The last term
wave approximation’’ approach generally show larger variations in Eq. (4.1) does not dissipate wave energy; it only causes a
(Oumeraci and Koether, 2009). phase shift to the waves passing through the wall. For most
For solitary waves, a single wall that occupies 98.5% of the still practical problems, Mei et al. (1974) argued that the inertial term
water depth may reduce the transmission coefficient to 0.54; in Eq. (4.1) was not important and could be safely ignored. For
however, it is possible to reduce the transmission coefficient to as completeness, we shall keep the inertial term throughout this
small as 0.33 using three submerged walls. These results suggest review.
Z.H. Huang et al. / Ocean Engineering 38 (2011) 1031–1053 1043

proposed the following empirical expression for the discharge


coefficient:
 2  
S0 d
Cc ¼ a þ ð1aÞ , a ¼ 0:6 þ 0:4 tanh ð4:4Þ
S 2e

where e is the gap between two adjacent rectangular bars. For


walls of very thick rectangular bars and with small gaps, a-1 and
Cc-1. For slotted barriers consisting of circular bars, we can still
use the diameter of the cylinder to calculate the porosity, but
mathematically take d ¼0 in Eq. (4.4) to obtain a ¼0.6, which
reduces the expression for Cc in Eqs. (4.4)–(4.3).
Fugazza and Natale (1992) used a constant discharge coeffi-
cient Cc¼ 0.5 in their study and modified Eq. (4.2) in the following
way to better fit their predictions with their experiments:
 2
1 S
Cf ¼ 1 ð4:5Þ
Cc S0

For analytical analysis of wave scattering by perforated/slotted


structures, a common practice is to linearize the quadratic term in
Eq. (4.1). To this end, Lorentz’s principle of equivalent work
(Sollitt and Cross, 1972; Mei et al., 1974; Zhu and Chwang,
2001, etc.) is invoked to introduce a dimensionless linearized
friction coefficient be

Cf 9uðz,tÞ9uðz,tÞ2
be ðzÞ ¼ at the perforated=slottedwall ð4:6Þ
2C uðz,tÞ2

where C is the wave celerity, and the over-bar represents a time


averaging over one wave period. After linearization, the pressure
drop across the perforated/slotted wall is related linearly to the
velocity at the down-wave side of the wall,
@uðx þ ,z,tÞ
pðx ,z,tÞpðx þ ,z,tÞ ¼ rbe ðzÞCuðx þ ,z,tÞ þ rLg ð4:7Þ
@t
Fig. 14. Wave passing through a row of cylinders: (a) circular cylinders and This linearization method ensures that the energy loss calcu-
(b) rectangular cylinders. lated for the linearized problem will be the same as that for the
original nonlinear problem. Note that be is independent of the
water depth for long waves.
For an array of closely spaced circular cylinders in long waves, Referring again to Fig. 14, the continuity equation can be
Mei et al. (1974) showed that Lg had an upper bound scaled by the approximately described by
distance between two adjacent cylinders. Huang and Ghidaoui
(2007) showed that Lg could be related to the added mass uðx þ ,z,tÞ  uðx ,z,tÞ ð4:8Þ
coefficient of the slotted barrier. Kakuno et al. (1992) and
as long as the perforated/slotted wall is very thin compared to the
Kakuno and Liu (1993) related the added mass coefficient (thus
wavelength (Mei et al, 1974). Assuming simple harmonic waves
the length scale Lg) to a blockage coefficient. For rectangular
of angular frequency o, we can write the dynamic pressure p, and
cylinders, Kakuno and Liu (1993) derived an expression for the
the horizontal velocity u as
blockage coefficient in terms of the slotted barrier’s geometrical
pffiffiffiffiffiffiffi
parameters, which was used by Kakuno et al. (1995) to study ^
ðpðx,z,tÞ,uðx,z,tÞÞ ¼ Re½ðpðx,zÞ, ^
uðx,zÞÞexpðiotÞ, i ¼ 1 ð4:9Þ
the wave forces on closely spaced cylinders. All these studies
suggested that Lg should not have a strong dependence on wave where p,^ u^ are the amplitudes of the dynamic pressure and
conditions. horizontal velocity, respectively.
The quadratic loss coefficient, Cf, is related to the orifice Substituting Eq. (4.9) in Eqs. (4.7) and Eq. (4.8), and combining
discharge coefficient Cc by the resulting equations lead to the following compact matching
 2 conditions at the slotted wall:
1 S
Cf ¼ 1 ð4:2Þ ^  ,zÞpðx
pðx ^ þ ,zÞ
Cc S0 ^  ,zÞ ¼ uðx
uðx ^ þ ,zÞ ¼ ð4:10Þ
rðbe ðzÞCioLg Þ
where S is the total area of a perforated/slotted plate and S0 is the
total net opening area in the slotted wall (Mei et al., 1974). Note Note that the linearized friction coefficient be depends on the
that the porosity of the slotted wall is simply e ¼ S0/S. It is well- scattered waves, which is part of the solution. Therefore, a
known that the discharge coefficient for a sharp-edged orifice can numerical iteration procedure is needed to determine be.
be described by the following empirical expression: The quadratic friction term in Eq. (4.1) may generate higher
 2 harmonic waves through wave-structure nonlinear interactions.
S0
Cc ¼ 0:6 þ0:4 ð4:3Þ However, Mei et al. (1974) showed that, for a single row of
S
circular cylinders, the transmission coefficients of the higher
For slotted barriers consisting of rectangular cylinders (bars) of harmonic wave components were much smaller than that of the
thickness d and porosity e, as shown in Fig. 14(b), Huang (2007b) fundamental waves.
1044 Z.H. Huang et al. / Ocean Engineering 38 (2011) 1031–1053

4.1.2. Porous-effect parameter for surface-piercing 4.1.3. Relation between energy loss coefficient and porous-effect
slotted/perforated walls parameter for surface-piercing structures
Another widely used approach to modeling the frictional and Since the porous-effect parameter G, given by Eq. (4.12), is
inertial effects of perforated/slotted walls is to use a porous-effect independent of water depth, a depth-averaged linearized dissipa-
parameter G (Chwang, 1983; Yu, 1995), which is derived origin- tion coefficient, b, can be obtained by averaging be(z), defined by
ally for thin porous walls interacting with simple harmonic water Eq. (4.6), over the water depth. For linear waves, the dynamic
waves. After the drag force being linearized using a linear porous pressure can be found from the linearized Bernoulli equation
resistance coefficient f, the flow inside the porous wall can be p ¼ r@f=@t. For simple harmonic waves, Eq. (4.10) can be
!
described by a velocity potential f defined by u ¼ ð@f=@x,@f=@zÞ. rewritten in the following form:
For a surface-piercing and bottom-sitting perforated/slotted wall
^ ðx,zÞexpðiotÞ, ^ ðx ,zÞ
@f ^ ðx þ ,zÞ
@f ^ ðx ,zÞf
io½f ^ ðx þ ,zÞ
in simple harmonic waves, i.e., fðx,z,tÞ ¼ Re½f ¼ ¼ ð4:15Þ
the following porous boundary conditions have been derived by @x @x ðbCioLg Þ
Yu (1995)
Comparing Eqs. (4.11) and (4.15) leads to
^ ðx ,zÞ
@f ^ ðx þ ,zÞ
@f 1
¼ ^ ðx ,zÞf
¼ ikG½f ^ ðx þ ,zÞ ð4:11Þ G¼ ð4:16Þ
@x @x bikLg
where wave-overtopping is not considered. In this set of equa- In view of Eq. (4.12) , we have the following relations:
tions, k is the wavenumber, and the porous effect parameter G is
defined by eb Lg
f¼ and s¼e ð4:17Þ
kd d
e Thus, length scale Lg is related to the inertial effect coefficient s,
G¼ ð4:12Þ
kdðf isÞ
which in turn is a function of added mass coefficient (see also
Huang and Ghidaoui, 2007).
where d is the thickness of the porous wall, e the porosity of the
porous wall, f the linear porous resistance coefficient, and s the
inertia coefficient, which is related to the added mass coefficient 4.1.4. Complex structure parameter for submerged
Cm by slotted/perforated walls
For a submerged perforated/slotted wall breakwater, the
1e
s ¼ 1 þCm ð4:13Þ energy loss due to the vortex shedding at the crest of the wall
e needs to be considered as well. To account for the energy loss at
Waves passing through a thin porous wall are accompanied by the crest of the submerged wall, the following empirical vortex
both energy dissipation and phase shift: the wave energy dis- loss coefficient Cv was introduced by Oumeraci and Koether
sipation, due to the resistance effect of the porous wall, is related (2009):
to the real part of G; the phase shift, due to the inertia effect of the pffiffiffi
Cv ¼ ð1 eÞCv ð4:18Þ
wall, is related the imaginary part of G. For 9G9¼0, the porous wall
becomes impermeable and no wave can go through the wall; for where Cv is the vortex loss coefficient for a vertically immersed
9G9-þ N, the wall becomes transparent to waves and there will impermeable plate, proposed by Stiassnie et al. (1984). Physically,
be no energy loss and wave scattering. A porous-effect parameter Cv is the ratio of the dissipated wave energy flux due to flow
similar to Eq. (4.12) is described by Chwang and Chan (1998) for separation at the wall crest to the incident wave energy flux. The
regular wave interaction with a thin surface-piercing porous wall. proposed vortex loss coefficient Cv reaches zero when the
In their derivation of their porous-effect parameter, it was porosity e approaches unity and reduces to Cv when the porosity
assumed that the seepage velocity inside the porous wall was approaches zero. We point out that the energy loss at the crest of
linearly proportional to the pressure difference on the two sides the submerged wall may be less significant for submerged
of the wall. Even though Eq. (4.12) was proposed originally for barriers in the form of closely spaced vertical cylinders as no
simple harmonic waves, it can be extended to irregular waves by structured vortex can shedding from the top of the cylinders.
defining equivalent regular waves for irregular wave; details are Oumeraci and Koether (2009) used the Morison equation, with
described in Section 4.2. modified drag and added inertia coefficients, to model the dynamic
The value of G for a given porous wall can be calculated with force on each of the cylinders in a submerged slotted wall. Based on
the known porous resistance coefficient f and the inertial effect their experiments, Oumeraci and Koether (2009) provided empirical
coefficient s (or the added mass coefficient Cm), both of which can expressions for the modified drag coefficient and the modified
only be obtained through physical experiments. Generally, the inertia coefficient, and proposed the following complex structure
inertial effect of a thin porous plate is not significant: a value of parameter to replace the porous-effect parameter G,
Cm ¼0, i.e., s¼1, may be used safely for most practical problems
1
(Yu, 1995; Isaacson et al., 1998; Li et al., 2006); some authors f¼ ð4:19Þ
ðfD þ fv ÞifI
even set s to zero (Chwang, 1983; Cho and Kim, 2000). As for the
porous resistance coefficient f, Li et al. (2006) developed the where fD is the modified drag loss coefficient, fI the modified inertia
following empirical formula for the models they used: coefficient, and fV the modified vortex loss coefficient which is
related to the vortex loss coefficient CV . A procedure was outlined in
 2  
d d Oumeraci and Koether (2009) to determine the three coefficients in
f ¼ 3338:7 þ 82:769 þ 8:711 ð4:14Þ
h h Eq. (4.19) using force measurements. Formally, Eq. (4.19) is similar
to Eq. (4.16) for G; however, the expression for G is not valid for
where h is the water depth at which the slotted wall of thickness submerged walls as the energy dissipation due to the vortex
d is located. They stated that Eq. (4.14) could be used for shedding from the crest of submerged walls is not considered in G.
0.0094r d/hr0.05. For large value of d/hZ0.1, f¼ 2.0 may be Note that the new complex structure parameter reduces to
used (Yu, 1995). Eq. (4.16) for emergent walls (fV ¼0). For impermeable submerged
Z.H. Huang et al. / Ocean Engineering 38 (2011) 1031–1053 1045

walls, the new structure parameter reduces to and the experimental data of Bennett et al. (1992) and Suh et al.
f ¼ 1=fV a0 ð4:20Þ (2001). We remark here that using a constant porous resistance
coefficient for all wave components is similar to using Method-II
Therefore, for submerged slotted/perforated breakwaters, the in that the energy dissipation is controlled by wave frequency
new structure parameter should be used. alone.

4.2. Irregular waves in the absence of steady currents 4.2.2. Approach based on Gaussian random processes
Massel and Mei (1977) first applied this approach to a problem
For irregular waves, the spectrum of the reflected waves of random waves interacting with a perforated wall. They exam-
SZ,R(o) and the spectrum of the transmitted waves, SZ,T(o), can ined the JONSWAP spectral waves under the assumption of long
be estimate by Eq. (3.4), where the reflection and transmission waves. Recently, Suh et al. (2001) extended the method to the
coefficients can be obtained using a regular wave model in which water waves in intermediate depth. Assuming that the horizontal
the linearized dissipation coefficient is estimated based on certain velocity of a train of irregular waves was a Gaussian process with
assumptions for the energy dissipation process. Two approaches a zero mean and a standard deviation s(z), Suh et al. (2001)
have been proposed to calculate the linearized dissipation coeffi- introduced an error function j defined by
cient: one is based on regular wave models and another is based  
on Gaussian random processes; the former is in extensive use in f
j ¼ be ðzÞC uðz,tÞ uðz,tÞ ð4:21Þ
practice because it is easier to implement than the latter. 2
The linear dissipation function b(z)is determined by minimizing
4.2.1. Approach based on regular wave models the expectation of the squared error j2 with respect to b(z), i.e.,
According to Suh and Son (2003), there are three methods in
@j2
this approach: (1) the first method (Method-I) approximates the ¼0 ð4:22Þ
@b
irregular waves by an equivalent regular wave train whose period
and height are the significant wave period and root-mean-square/ Denoting the expectation of the random variable X by E(X), the
significant wave height, respectively. With the equivalent waves following expression was derived from Eq. (4.22)
defined, the linearized dissipation coefficient be(z) can be calcu- rffiffiffiffi
f Eðjuj3 Þ f 8
lated using Eq. (4.6) with u(x, z, t) being provided by the bðzÞ ¼ ¼ sðzÞ ð4:23Þ
2 Eðu2 Þ 2 p
equivalent waves; (2) the second method (Method-II) calculates
the reflection and transmission coefficients for each component For a given spectrum of incident waves, SZI(o), the spectrum of
wave, with the same root-mean-square wave height being used the horizontal velocity u is Su(o), which is formally related to
for all component waves; and (3) the third method (Method-III) SZI(o) by
calculates the reflection and transmission coefficients for each 2
Su ðon ,zÞ ¼ 9Tn ðo,zÞ9 SZI ðon Þ ð4:24Þ
component wave whose height is obtained from the energy
spectrum of incident waves. Since the frequency-averaged reflec- with Tn(o,z)being a transfer function between the velocity and
tion and transmission coefficients can be calculated by Eq. (3.5), the surface elevation for the problem under investigation. It then
Method-I assumes constant reflection and transmission coeffi- follows that the standard deviation s(z) can be determined by
cients for all component waves. For Method-II and -III, the sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Z ffi sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Z ffi
1 1
2
reflection and transmission coefficients vary with component sðzÞ ¼ Su ðo,zÞ do ¼ 9Tn ðo,zÞ9 SZI ðoÞ do ð4:25Þ
wave frequency. Method-II assumes that the energy dissipation 0 0

is controlled only by the frequency of the component waves, For perforated-wall caisson breakwaters of chamber width B,
while Method-III assumes that each wave component dissipates Suh et al. (2001) gave the following expression for the transfer
wave energy independently, neglecting the non-linear interac- function Tn(o,z)
tions among the component waves.
cosh kn ðz þhÞ Wn
Using the experimental data of Bennett et al. (1992) and Suh et al. Tn ðo,zÞ ¼ on qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi ð4:26Þ
sinh kn h
(2001), Suh and Son (2003) used the root-mean-squared height in Wn ð1þ Rn Þ2 þ Cn2
2

their Method-I and made a comparison among the three methods


(Methods I–III); their comparison shows that the performance of where
Method-II is better than both Method-I and -III for surface-piercing Wn ¼ tanðkn BÞ, Rn ¼ bkn =on , Rn ¼ bkn =on , Cn ¼ 1Pn Wn ,
perforated/slotted structures: Method-III always over-predicts the Pn ¼ 2kn Lg ð4:27Þ
frequency-averaged reflection coefficients and their Method-I over-
predicts/under-predicts the frequency-averaged reflection coefficients The transfer functions for other perforated/slotted structures
for strong/weak reflections. Oumeraci and Koether (2009) used the can be derived following the procedure described in Suh et al.
significant wave height in their Method-I and made a comparison (2001) or Huang (2006). From Eq. (4.23), a depth-integrated linear
between the Method-I and the Method-III for submerged slotted dissipation coefficient b can be defined by
"Z # "Z #
walls using their experimental data; their comparison supports that 0 0
Method-III performs better than their Method-I. bE u2 dz ¼ E bðzÞu2 dz ð4:28Þ
h h
Even though Eqs. (4.11) and (4.12) were originally derived for
simple harmonic waves, Liu et al. (2007d), suggested that the
porous resistance coefficient f in Eq. (4.12) be simply treated as
a constant for all component waves of irregular waves. (The 4.3. Regular waves in the presence of a steady current
porous effect parameter still varies with wave frequency due to
its dependence on wavenumber.) They obtained a linear transfer Energy loss associated with the flows through the gaps in
function between the incident and reflected wave spectra perforated/slotted walls can be enhanced by steady currents
using the equivalent regular wave model. Subsequently the (wind-driven currents or tides). For long waves, Huang (2007a)
frequency-averaged reflection coefficients were calculated and showed experimentally that the transmission coefficients could
the results agreed reasonably well with their experimental data be reduced significantly by wave-opposing currents, while the
1046 Z.H. Huang et al. / Ocean Engineering 38 (2011) 1031–1053

reflection coefficients could be increased slightly by the same


wave-opposing currents. Huang and Liu (2008) modified Eq. (4.1)
by including the effects of a weak current. In the presence of a
weak, uniform current, the pressure change across the slotted
plate is modeled by
Cf
pðx ,z,tÞpðx þ ,z,tÞ ¼ r ~ þ ,z,tÞ þuÞ9uðx
ðuðx ~ þ ,z,tÞ þ u9
2
~ þ ,z,tÞ
@uðx Cf
þ rLg r ðuðx~ þ ,z,tÞ þ uÞ9uðx
~ þ ,z,tÞ þ u9
@t 2
ð4:29Þ Fig. 15. Approximation of a solitary wave as a train of truncated solitary waves
(after Huang and Yuan, 2009).
where u is the mean current velocity, tildes represent the wave
part, and the long over-bar represents a time average over one series of solitary waves (long waves) under the assumption that
wave period. Part of the nonlinear interaction between the the equivalent wavelength was not affected by the solitary wave
current and waves is implied in the dispersion equation where height.
the wavelength is changed by the current: the incident waves and
the reflected waves will have different wavelengths. Eq. (4.29) is 4.5. Oblique waves in the absence of currents
applied at x¼ x þ , the immediate down-wave side of the perforate/
slotted wall. The first nonlinear term in Eq. (4.29) can be Li et al. (2002, 2003a) have successfully applied the porous-
linearized again by using Lorentz’s principle of equivalent work effect parameter method to the interactions of oblique regular
to obtain the following linear dissipation coefficient: waves with perforated/slotted structures; they concluded that the
Cf ju þ ujðu
~ þ uÞ ~ u~ porous-effect parameter obtained by the experiments for normal
be ðzÞ ¼ at x ¼ x þ ð4:30Þ waves could be used for the cases of oblique regular waves under
2C u~ u~
most practical conditions. However, the methods based on the
For a single slotted barrier in long waves, Huang and Ghidaoui quadratic loss coefficient are not easy to use when modeling
(2007) found that the reflection and transmission coefficients oblique waves interacting with slotted/perforated structures.
predicted for relatively weak currents agreed well with the
measurements reported in Huang (2007a). For double slotted
barriers in intermediate water waves, Huang and Liu (2008) 5. Wave forces on perforated caissons
compared the reflection and transmission coefficients calculated
using Eq. (4.30) with those measured by Huang (2008), and found In addition to minimizing the reflection coefficients, the
good agreements with experiments only for those cases of weak horizontal and vertical forces acting on perforated caisson break-
currents. For strong currents, there is a significant change of the waters with back-wall must also be carefully analyzed when
mean water level across the barrier, which makes the matching designing such structures. Unfortunately, there are not much
condition Eq. (4.8) invalid. published results on this topic in the literature, lacking especially
in the theoretical or numerical studies. In this section three
4.4. Solitary waves in the absence of currents different methods for calculating wave forces on perforated
caisson breakwaters are reviewed. Limited also by the availability
Huang and Yuan (2009) and Liu et al. (2010) recently studied of experimental data, no direct comparison among the reviewed
the interactions between a solitary wave and a row of closely methods is attempted in this review.
spaced circular cylinders. Huang and Yuan (2009) also proposed a
simple theory to model the energy dissipation. In their theory the 5.1. Takahashi’s method
solitary wave was replaced by a train of truncated solitary waves
with the equivalent length L and period T defined by Takahashi’s method (Takahashi et al., 1992, 1996; Tanimoto and
rffiffiffiffiffiffiffi Takahashi, 1994) is a modification to the well-known Goda’s
h3 L pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
L ¼ 4p and T ¼ , C ¼ gðh þHÞ ð4:31Þ formula (see, for example, Goda (2000) for a description), which
3H C can be used to calculate wave forces on an upright solid wall. In this
where H is the height of the solitary wave in water of depth h. They method, it is assumed that the wave pressure along the front wall
approximated this train of truncated solitary waves as a train of has a trapezoidal distribution both above and below the still water
long waves riding on a mean current, as shown in Fig. 15. In this level; it is also assumed that the uplift pressure on the breakwater
method, the surface elevation and the velocity of this train of bottom has a triangular distribution, as shown in Fig. 16, where the
solitary waves are decomposed into a mean part and a fluctuation wave pressure p1, p2, p3, p4, and pu can be calculated by the
around the mean at any given point x, i.e., following formulas (Goda, 2000; Takahashi et al., 1992):
Zðx,tÞ ¼ ZðxÞ þ Z~ ðx,tÞ,uðx,tÞ ¼ uðxÞ þ uðx,tÞ
~ ð4:32Þ p1 ¼ 0:5ð1 þ cos yÞðl1 a1 þ l2 a cos2 yÞgHd ð5:1Þ
where the mean parts are defined by p1
Z T Z T p2 ¼ ð5:2Þ
cos hð2pd=LÞ
ZðxÞ ¼ Zðx,tÞ dt,uðxÞ ¼ uðx,tÞ dt ð4:33Þ
0 0
p3 ¼ a3 p1 ð5:3Þ
After Eq. (4.30) is used to calculate the linear dissipation
coefficient with the mean current being defined by Eq. (4.33), p4 ¼ a4 p1 ð5:4Þ
the problem of solitary wave scattering by slotted barriers can be
solved as if the waves were regular waves. pu ¼ 0:5ð1 þ cos yÞl3 a1 a3 gHd ð5:5Þ
It is interesting to note that the method of Huang and Yuan  2
(2009) is similar to that of Mei (1985), who studied the scattering 4pd=L
a1 ¼ 0:6 þ0:5 ð5:6Þ
of solitary wave at a step and approximated the solitary wave by a sinhð4pd=LÞ
Z.H. Huang et al. / Ocean Engineering 38 (2011) 1031–1053 1047

p4
0
p1
Crest I Crest II a Crest II b
hc

Caisson
d1 d2 Trough I Trough II Trough III
d p3
Fig. 17. Different situations for wave action on perforated caissons defined by
pu
Takahashi (1996).

p2
caisson. Takahashi (1996) have provided the values of the
correction coefficients for all six wave phases. These values are
Fig. 16. Pressure distributions on the upright solid wall by the modified Goda’s shown in Table 1, where L0 is the wavelength at water depth d1,
formula.
and B0 equals to the wave chamber width B plus the front wall
(   ) thickness, and the first subscript (S, L, R, M, U) of lj (j¼ 1, 2, 3)
hb d2 Hd 2 2d2 denotes the position where the correction factor lj is used. The
a2 ¼ min , ð5:7Þ
3hb d2 Hd sketch of pressure distributions on the perforated caisson at Crest
  IIb is shown in Fig. 18. Other details of Takahashi’s method can be
d1 1 found by Tanimoto and Takahashi (1994).
a3 ¼ 1 1 ð5:8Þ
d coshð2pd=LÞ
5.2. Tabet-Aoul and Lambert’s method
hc 
a4 ¼ 1 ,hc ¼ min Z0 ,hc ð5:9Þ
Z0 Based on the experimental data, Tabet-Aoul and Lambert
(2003) concluded that the maximum horizontal force acting on
Z0 ¼ 0:75ð1 þ cos bÞl1 Hd ð5:10Þ
the perforated caisson generally would occur between the three
a ¼ maxfa1 , aip g ð5:11Þ crest wave phases defined by Takahashi (1996). They developed a
tentative formula to calculate the maximum horizontal force
where y is the angle between the incident wave direction and the on perforated caissons by introducing a new phase adjustment
normal to the caisson breakwater; g is the specific weight of sea factor. Shown in Fig. 19 is the pressure distribution in Tabet-Aoul
water; hb denotes the water depth at the offshore location by a and Lambert (2003), which is calculated in a way similar to the
distance of five times the significant wave height Hs; Hd and L are modified Goda’s formula (Goda, 2000; Takahashi et al., 1992). The
the design wave height and wavelength, respectively. The symbols formulas for wave pressures are
of l1, l2, and l3 are the correction coefficients for the wave    
B
pressures recommended by Takahashi, and all are usually set to pp1 ¼ ð1 þ cos bÞ 0:21a1 þ ð1 þ a Þcos2 y gHd ð5:12Þ
unity for ordinary upright walls. The factor l1 denotes the variation 4L
of the slowly varying wave pressure, the factor l2 the change of the
pp3 ¼ a3 pp1 ð5:13Þ
breaking pressure component, and the factor l3 the change in the
uplift pressure. The factor aip in Eq. (5.11) is the so-called impulsive
pp4 ¼ a4 pp1 ð5:14Þ
pressure coefficient. The method to calculate aip, omitted here, can
be found by Takahashi et al. (1992) or Tanimoto and Takahashi "    #
B 2 B
(1994). Other symbols (Z0, hc, d, d1, and d2) in Eqs. (5.1)–(5.11) are pr1 ¼ 0:5ð1 þcos yÞ 0:7 a1 þ 0:43 ð1þ a Þcos2 y gHd
L L
defined in Fig. 16. The design wave height in this method is taken as
Hd ¼Hmax ¼H1/250 ¼1.8H1/3 seaward of the surf zone, and the largest ð5:15Þ
wave height is the height of breaking waves at the breaking water
depth hb within the surf zone. The wavelength of the largest wave is pr3 ¼ a3 pr1 ð5:16Þ
generally calculated using the significant wave period Ts.
pr4 ¼ a4 pr1 ð5:17Þ
Takahashi (1996) demonstrated that, unlike wave forces acting
on upright solid walls, the maximum horizontal and vertical Zp0 ¼ 0:32ð1þ cos yÞHd ð5:18Þ
forces acting on perforated caisson did not occur at the time
when the wave crest was attacking the caisson’s front barrier. He  2
B
defined six different wave phases labeled by Crest I, Crest IIa, Zr0 ¼ 0:75ð1 þcos yÞ 0:7 Hd ð5:19Þ
L
Crest IIb, Trough I, Trough II and Trough III, as shown in Fig. 17
(see also Tanimoto and Takahashi, 1994). The wave force on the where, pp1, pp3, and pp4 denote the resultant pressures on the still
perforated front wall reaches a peak at Crest I, and the wave force water level, the bottom and the top of the perforated front wall,
on the back-wall may have two peaks—a primary peak associated respectively; pr1, pr3, and pr4 denote the pressures on the still water
with the impulsive force occurred at Crest IIa, followed by a level, the bottom and the top of the rear wall, respectively; Zp0 and
secondary peak of smaller magnitude occurred at Crest IIb. The Zr0 are the proposed wave run-up on the perforated front wall
wave force on the front wall reaches its minimum at Trough I. The and the rear wall, respectively. The definitions of other symbols in
wave trough arrives at the front wall at Trough II, while the Eqs. (5.12)–(5.19) are the same as those in Eqs. (5.1)–(5.11). It is
lowest water level occurs in the wave absorbing chamber at noted that when calculating a4, the wave run-up Z0 in Eq. (5.10)
Trough III. Takahashi (1996) suggested that all three crest wave should be replaced by the present Zp0 for the perforated front wall
phases should be considered for engineering designs. The correc- and by Zr0 for the rear wall.
tion coefficients l1, l2, and l3 are no longer equal to unity, but The formulas for the total horizontal peak wave forces on a
they are rather different at different parts of the perforated perforated caisson, as proposed by Tabet-Aoul and Lambert
1048 Z.H. Huang et al. / Ocean Engineering 38 (2011) 1031–1053

Table 1
Values of correction coefficients for perforated caissons given by Takahashi (1996).

CrestI CrestIIa CrestIIb

Perforated part of the front wall lS1 0.85 0.7 0.3


lS2 0.4 (an r 0.75) 0 0
0.3/an (an 40.75)
Solid part of the front wall lL1 1.0 0.75 0.65
lL2 0.4 (an r 0.75) 0 0
0.2/an (an 40.75)
Wave chamber rear wall lR1 0 20B0 /3L0 (B0 /L0 r 0.15) 1.0 (B0 /L0 4 0.15) 1.4 (Hd/d r 0.1)
1.6–2Hd/d (0.1 oHd/d o 0.3)
1.0 (Hd/d Z0.3)
lR2 0 0.56 (an r 25/28) 0
0.5/an (an 4 25/28)
Wave chamber bottom slab lM1 0 20B0 /3L0 (B0 /L0 r 0.15)1.0 (B0 /L0 4 0.15) 1.4 (Hd/d r 0.1)
1.6–2Hd/d (0.1 oHd/d o 0.3)
1.0 (Hd/d Z0.3)
lM2 0 0 0
Uplift lU3 1.0 0.75 0.65

wall, and Ftot is the total horizontal peak force acting on the caisson.
R The symbol e denotes the front wall porosity and the factor w is the
so-called ‘‘phase adjustment factor’’, which is the ratio of total
maximum horizontal force to the sum of maximum forces acting
M
on the perforated front wall and the rear wall. Tabet-Aoul and
s d'2 Lambert (2003) suggested that Eq. (5.22) should be safely used for
d1 0oB/Lo0.35.
d d2
L
B 5.3. Research in DUT, China, on wave forces on perforated caissons

Crest II b
The Dalian University of Technology (DUT), China, has carried
out a series of numerical, analytical and experimental studies on
U
perforated caissons since 1998, and their proposed method has
been included in a new version of Chinese Code of Design and
Fig. 18. Pressure distributions on the perforated caisson at Crest IIb given by Construction of Breakwaters.
Takahashi (1996). After Tanimoto and Takahashi (1994). In their numerical studies (Chen et al., 2003, 2007), 2D
Reynolds averaged Navier–Stokes equations were solved by using
a finite-difference method. The nonlinear free surface was treated
pp4 p r4 by means of the VOF method. The reflection coefficients and the
p0 horizontal wave forces were calculated for both regular and
hc r0
irregular waves. In the analytical studies (Li et al., 2002, 2003a;
Teng et al., 2004; Liu et al., 2006b, 2007d, 2008) they used eigen-
p r1 function expansion methods and studied the reflection coeffi-
p p1 cients and the horizontal wave forces for both normally and
d2
p r3 obliquely incident waves.
p p3
Both 2D and 3D physical experiments have been carried
out (Liu, 2003; Jiang, 2004; Ma, 2004; Li et al., 2003b, 2005a,
B
2005b) in DUT labs. The majority of 2D model tests were
Fig. 19. Pressure distributions on the perforated caisson given by Tabet-Aoul and conducted in a wave-current flume, which was 56 m long, 0.7 m
Lambert (2003). After Tanimoto and Takahashi (1994). wide and 1.0 m deep, under the test conditions shown in Table 2.
In their studies, the caisson models were placed on rubble-filled
foundations consisting of gravels. The caisson chambers under
(2003), are as follows: the perforated part of the front perforated wall were filled with
  fine gravels. Both regular and irregular waves were examined in
d2 hc
Fp ¼ ðpp1 þ pp3 Þ þ ðpp1 þ pp4 Þ ð1eÞ ð5:20Þ these tests. The modified JONSWAP spectrum with the peak
2 2
enhancement factor being fixed at g ¼3.3 (Goda, 1988) were used
 in the irregular wave tests. No wave breaking and overtopping
d2 hc
Fr ¼ ðpr1 þ pr3 Þ þðpr1 þ pr4 Þ ð5:21Þ were observed in all their model tests.
2 2
The numerical, analytical and the experimental results have
Ftot ¼ wðFp þFr Þ ð5:22Þ been validated against each other. Fig. 20 shows the comparisons
between numerical and experimental results of the total hor-
   3   izontal force on caissons and the vertical force on the top cover
9 B 11 B 2 B 10 B 4
w ¼ 1 þ  4 þ ð5:23Þ plate; Fig. 21 shows the comparisons between analytical and
25 L 4 L L 3 L
experimental results of the total horizontal and vertical wave
where Fp is the total horizontal peak force acting on the perforated forces on perforated caissons without top cover, where all the
front wall, Fr is the total horizontal peak force acting on the rear forces have been normalized by rghHs. It can be seen from these
Z.H. Huang et al. / Ocean Engineering 38 (2011) 1031–1053 1049

Table 2
Experimental conditions and geometric parameters in the tests of DUT.

Height of top cover above the Geometrical porosity of Water depth in Depth of Width of wave- Width of
still water Sc (cm) front wall e front of model h perforated portion absorbing perforated
(cm) h1 (cm) chamber B (cm) caisson (cm)

4, 8, 12, 16, 20, 24, no top cover 0.0, 0.2, 0.4 40 20 15, 20, 30 45
Regular wave tests Wave period T (s) 0.86 1.0 1.2 1.4
Wave height H (cm) 6, 8 6, 8, 10, 12 6, 8, 10, 12 6, 8, 10, 12
Irregular wave tests Significant wave period Ts (s) 0.99 1.15 1.38
Significant wave height Hs (cm) 5.3, 6.7, 8.0 5.3, 6.7, 8.0 5.3, 6.7, 8.0

0.4 1
Numerical results (KN/m)

0.8
0.3

Calculated results
0.6
0.2
0.4

0.1
0.2

0 0
0 0.1 0.2 0.3 0.4 0 0.2 0.4 0.6 0.8 1
Test data | F xs| / ρ ghH s
Test data (KN/m)
0.3 0.6
Numerical results (KN/m)

Calculated results

0.2 0.4

0.1 0.2

0 0
0 0.1 0.2 0.3 0 0.2 0.4 0.6
Test data (KN/m) Test data | F zs| / ρ ghH s

Fig. 20. Comparisons between numerical results and experimental data for Fig. 21. Comparisons between predicted and measured significant values of total
perforated caissons with top cover (Chen et al., 2007): (a) Total horizontal wave wave forces for perforated caissons without top cover (Liu et al., 2008): (a) total
force on perforated caissons and (b) total vertical wave force on top cover. horizontal force and (b) total vertical force.

comparisons that both the numerical and theoretical predictions wall reaches its peaks, the corresponding total vertical forces are
agree well with the measurements. From the numerical, analy- in general small, sometimes even negative (acting downward).
tical and experimental results, a simplified method was devel- This phenomenon is beneficial to the stability of a perforated
oped to calculate the total wave forces acting on a perforated caisson under the attack of wave crests. Therefore, the phase
caisson breakwater with or without a top cover plate. This difference between the total horizontal and vertical peak forces
method has been adopted in the latest version of Chinese Code should be considered for economical designs.
of Design and Construction of Breakwaters. The DUT’s method (Jiang et al., 2004; Li et al., 2005a, 2005b;
The DUT’s experimental and numerical results (Jiang et al., Chen et al., 2005; Li, 2007) can predict the following: (1) the total
2004; Li et al., 2005a, 2005b; Chen et al., 2005; Li, 2007; Liu et al., horizontal force, the total vertical force, the arm of force, and the
2008) showed that when the total vertical peak force was acting moment; (2) the phase difference between total horizontal and
on the perforated caisson with or without a top cover plate, there vertical force peaks; (3) the pressure difference on the perforated
was always an obvious phase delay of the total horizontal front wall , the pressures on the caisson’s bottom plate, solid rear
peak force. Selected time histories of total wave forces are wall, and top cover; and (4) the reflection coefficient. They noted
plotted in Fig. 22(a) for Sc ¼4 cm, e ¼0.4, B ¼15 cm, Hs ¼6.7 cm, that the measured wave forces for irregular waves were generally
and Ts ¼ 1.15 s. The results for the corresponding solid caisson are larger than those for the corresponding regular waves. Therefore,
shown in Fig. 22(b). When the total horizontal force on the front the formulas for irregular waves are recommended for
1050 Z.H. Huang et al. / Ocean Engineering 38 (2011) 1031–1053

0.3

0.2 Vertical force

0.1
KN/m
0

-0.1
Horizontal force
-0.2

-0.3
t (s)

0.3
Vertical force
0.2

0.1
KN/m

-0.1

-0.2 Horizontal force


-0.3
t (s)

Fig. 22. Measured time histories of total wave forces on caisson breakwaters: (a) perforated caisson breakwater and (b) solid caisson breakwater.

     
conservative designs. For irregular waves, the simplified formulas lv1 Hs h B
¼ 1:011:917 þ 1:063 2:023
for the maximum total forces, the arms of the total forces, the lv0 Ls Ls Ls
phase differences, and the force reduction coefficient due to phase  2
B
differences are listed below: þ2:532 þ0:429e ð5:30Þ
Ls
     
P1 Hs h B
¼ 0:9971:515 0:804 1:312 þ0:25e ð5:24Þ lv2
P0 Ls Ls Ls ¼ 1:0 ð5:31Þ
lv1
   
P2 h B    
¼ 1:247 þ 0:648 0:573 Dt1 B h
P1 Ls Ls ¼ 0:009 þ 0:477 þ 0:099 þ0:324e ð5:32Þ
Ts Ls Ls
   2
Sc Sc      
0:349 þ0:082 þ0:215e ð5:25Þ Dt2 B h Sc
Hs Hs ¼ 0:237 þ 0:304 þ0:08 þ 0:299
Ts Ls Ls Hs
     2
Pv1 Hs h Sc
¼ 0:088þ 11:154 2:084 0:088 þ0:347e ð5:33Þ
Pv0 Ls Ls Hs
   2
B B  
þ 8:273 19:508 þ 0:832e ð5:26Þ P10 Dt1
Ls Ls ¼ cos 2p ð5:34Þ
P1 Ts
     
Pv2 Hs h P20 Dt2
¼ 2:1347:056 þ 0:961 ¼ cos 2p ð5:35Þ
Pv1 Ls Ls P2 Ts
   2
B B  
0:296 þ 4:3656 0
Pv1 Dt1
Ls Ls ¼ cos 2p ð5:36Þ
   2 Pv1 Ts
Sc Sc
0:811 þ 0:184 0:703e ð5:27Þ 0  
Hs Hs Pv2 Dt2
¼ cos 2p ð5:37Þ
Pv2 Ts
     2  
l1 Hs h h B where L
¼ 1:0632:56 þ 1:019 0:88 1:432
l0 Ls Ls Ls Ls s is the significant wavelength. The symbols P and Pv

 2 denote, respectively, 1% cumulative probabilities of the maximum


B total horizontal and the vertical forces on a caisson of unit length.
þ2:848 þ0:091e ð5:28Þ
Ls The symbols l and Dt denote, respectively, the arms of the total
wave forces and the phase shift (time delay) between the total
l2 horizontal and the vertical peak forces. The arm of total horizontal
¼ 1:0 ð5:29Þ
l1 wave force is equal to the distance between the total horizontal
Z.H. Huang et al. / Ocean Engineering 38 (2011) 1031–1053 1051

force action point and the caisson front toe, and the arm of total of the maximum vertical forces and the phase relations between
vertical force is equal to the distance between the total vertical the horizontal and vertical forces are not discussed in this
force action point and the caisson back heel. The symbol P0 method. In addition to the information provided by Takahashi’s
denotes the total horizontal force corresponding to the maximum method and Tabet-Aoul and Lambert’s method, DUT’s method can
total vertical force, and the symbol Pv0 denotes the total vertical also provide the information on the force on the top cover as well
force corresponding to the maximum total horizontal force. Sub- as the information on the phase difference between the total
scripts 0–2 refer to solid caissons, perforated caissons without top horizontal and the total vertical forces. Using the DUT’s method,
cover plates, and perforated caissons with top cover plates, the structure can be designed such that the horizontal and
respectively. Other parameters are defined in Table 2. The main vertical forces are nearly out of phase, which helps enhance the
dimensionless parameters used in the above simplified formulas stability against sliding of perforated caissons and may lead to an
were selected based on theoretical analysis and numerical calcu- economic design.
lations, and all coefficients in the above formulas were obtained Even though Li (2007) has presented a comparison among the
from the experimental data. These formulas can be used for three reviewed methods using DUT’s experimental data and the
e ¼ 0.2–0.4, B/Ls ¼0.064–0.209, d/Ls ¼0.171–0.278, and d/Hs Z2.0. DUT’s method has been adopted in a newly revised Chinese
The effects of the top cover plate can be ignored when Sc/Hs Z2.1 design code, third party experiments are useful for making an
in Eqs. (5.25) and (5.27), and Sc/Hs Z1.7 in Eq. (5.33). A typical independent evaluation and recommendation of the existing
comparison between the experimental data and the predictions methods for calculating wave loads.
by Eq. (5.25) is given in Fig. 23. It can be seen that the agreement
between predicted and measured results is reasonably well. For
obliquely incident waves, a simplified method has also been given 6. Concluding remarks
to calculate the reflection coefficient and the horizontal force on a
perforated caisson, but is not reviewed here. Major types of coastal perforated/slotted structures with or
Instead of directly calculating the wave forces acting on a without a back-wall, energy dissipation mechanisms and three
perforated caisson, the DUT’s method calculates the ratios of the methods for calculating wave forces on perforated caissons have
force on a perforated caisson to the force on a solid caisson. The been reviewed. For Jarlan-type perforated structures, small reflec-
wave forces acting on a solid caisson can be calculated using other tion coefficients over a wide range of wave frequency can be
widely accepted methods in the literature or design codes, for achieved by using multiple perforated walls, submerged rock
example, the Goda’s formula. Both the effects of a top cover plate core, or internal porous plates. For fully perforated Jalrlan-type
and the phase difference between the horizontal and vertical peak structures, the frequency-averaged reflection coefficients of irre-
forces are considered in the DUT’s method as well. gular waves are found to be generally larger than those of regular
waves under practical wave conditions, and oblique waves may
result in larger wave reflection than normal incident waves. For
5.4. Comments on the three reviewed methods surface-piercing slotted breakwaters without a back-wall, small
transmission coefficients over a wide range of wave frequency can
Maximum horizontal and vertical wave forces, as well as the be achieved by using multi-slotted walls with the porosity of each
phases at which they occur, are important design parameters. wall carefully chosen. Impermeable curtain/crown can also effec-
Takahashi’s method calculates both maximum horizontal and tively reduce the transmission coefficients. Coastal currents may
maximum vertical wave forces for normally/obliquely incident help to reduce the transmission coefficients. For submerged
waves; three wave phases were also identified at which the perforated/slotted walls, the energy dissipation associated with
horizontal force may reach local maximum values. The designs the vortex shedding at the crest of the submerged wall needs to
based on Takahashi’s method tend to be conservative as only the be considered. Submerged slotted walls may have potential
magnitudes of the maximum forces are considered. Tabet-Aoul engineering applications for mitigating tsunami hazards and
and Lambert’s method calculates various horizontal forces and reducing shoreline erosion. Three kinds of methods for calculating
the phase differences among them, pointing out that the max- wave forces on caisson breakwaters have been reviewed for the
imum horizontal forces may occur at the phases different from structural designs of surface-piercing slotted/perforated struc-
those identified by Takahashi’s method; however, the calculation tures, but more experimental data are needed for a convincing
comparison among these methods.
1.6 Aspects other than hydrodynamic coefficients and structural
stability also need to be considered. For the protection of coastal
environments, the flushing time of harbors and the water exchange
1.4
across the slotted/perforated breakwaters are worth consideration
Predictions by Eq.(5.25)

in further studies. Scouring around the pile breakwaters also deserve


1.2 more research. Possible oscillations and contaminant transport in
harbors protected by slotted/perforated breakwaters are also chal-
lenging topics.
1

Acknowledgments
0.8
The first author would like to thank the Nanyang Technologi-
0.6 cal University for supporting part of his research through
0.6 0.8 1 1.2 1.4 1.6 Project no. SUG 3/07. The second author would like to thank the
Test data (P2/ P1) National Natural Science Foundation of China for supporting
his study through Chinese national innovation Project no.
Fig. 23. Comparison between experimental results and calculated results from 50921001. The third author would like to thank the National
Eq. (5.25). Natural Science Foundation of China for supporting his study
1052 Z.H. Huang et al. / Ocean Engineering 38 (2011) 1031–1053

through Project no. 50909086. The authors would like to appreci- Huang, Z.H., Yuan, Z., 2009. Transmission of solitary waves through slotted
ate the following DUT research team members for their contribu- barriers: a laboratory study with analysis by a long wave approximation.
Journal of Hydro-environment Research 3, 179–185.
tions to this work: Professor Bin Teng, Professor Dapeng Sun, Hutchinson, P.S., Raudkivi, A.J., 1984. Case history of a spaced pile breakwater at
Professor Ningchuan Zhang, Doctor Xuefeng Chen, Doctor Hongjie Half Moon Bay Marina, Auckland, New Zealand. In: Proceeding of the Coastal
Liu, Doctor Lu Sun, Mr. Junjie Jiang, and Mr. Baolian Ma. We would Engineering, vol. 84, pp. 2530–2533.
Isaacson, M., Baldwin, J., Allyn, N., Cowdell, S., 2000. Wave interactions with
like to thank the three anonymous reviewers for their valuable
perforated breakwater. Journal of Waterway, Port, Coastal, and Ocean Engi-
comments, which have greatly improved the quality of this paper. neering 126 (5), 229–235.
Isaacson, M., Premasiri, S., Yang, G., 1998. Wave interactions with vertical slotted
barrier. Journal of Waterway, Port, Coastal, and Ocean Engineering 124 (3),
118–126.
References
Jarlan, G.E., 1961. A perforated vertical wall breakwater. Dock and Harbour
Authority XII (486), 394–398.
Bennett, G.S., McIver, P., Smallman, J.V., 1992. A mathematical model of a slotted Ji, C.H., Suh, K.D., 2010. Wave interactions with multiple-row curtain wall-pile
wavescreen breakwater. Coastal Engineering 18, 231–249. breakwaters. Coastal Engineering. doi:10.1016/j.coastaleng.2009.12.008.
Bergmann, H., Oumeraci, H., 2000. Wave loads on perforated Caisson breakwaters. Jiang, J.J., 2004. Experimental Study on Vertical Wave Forces Acting on Perforated
In: Proceedings of the 27th Coastal Engineering Conference. ASCE, Sydney, Caissons. Master Thesis. Dalian University of Technology (in Chinese with
Australia, pp. 1622–1635. English abstract).
Chen, X.F., Li, Y.C., Ma, B.L., Jiang, J.J., Lu, G.R., 2005. Calculating method of irregular Jiang, J.J., Li, Y.C., Sun, D.P., Ma, B.L., 2004. Experimental study of the vertical wave
wave pressures on components of perforated caissons with top cover. China forces acting on perforated caisson by irregular waves. China Offshore Plat-
Offshore Platform 20 (4), 1–9 (in Chinese with English abstract). form 19 (5), 1–8 (in Chinese with English abstract).
Chen, X.F., Li, Y.C., Sun, D.P., 2002. Regular waves acting on double-layered Kakuno, S., 1983. Reflection and transmission of waves through vertical silt-type
perforated caissons. In: Proceedings of the 12th (2002) ISOPE, USA, vol. 3, structure, Proceedings of the Coastal Structures ’83. ASCE, pp. 939–953.
pp. 736–743. Kakuno, S., Liu, P.L.-F., 1993. Scattering of water waves by vertical cylinders.
Chen, X.F., Li, Y.C., Teng, B., 2007. Numerical and simplified methods for the Journal of Waterway, Port, Coastal, and Ocean Engineering—ASCE 119 (3),
calculation of the total horizontal wave force on a perforated caisson with a 302–322.
top cover. Coastal Engineering 54, 67–75. Kakuno, S., Nakata, Y., 1998. Scattering of water waves by rows of cylinders
Chen, X.F., Li, Y.C., Wang, Y.X., Dong, G.H., Bai, X., 2003. Numerical simulation of without a backwall. Applied Ocean Research 20, 191–198.
wave interaction with perforated caissons breakwaters. China Ocean Engi- Kakuno, S., Nakata, Y., Liu, P.L.-F., 1995. Wave forces on an array of vertical
neering 17 (1), 33–43. cylinders. Journal of Waterway, Port, Coastal, and Ocean Engineering 122 (3),
Cho, I.H., Kim, M.H., 2000. Interactions of horizontal porous flexible membrane 147–149.
with waves. Journal of Waterway, Port, Coastal, and Ocean Engineering 126 Kakuno, S., Oda, K., Liu, P.L.F., 1992. Scattering of water waves by vertical cylinders
(5), 245–253. with a backwall, Proceedings of the 23th Coastal Engineering Conference.
Cho, I.H., Kim, M.H., 2008. Wave absorbing system using inclined perforated ASCE, Venice, Italy, pp. 1258–1271.
plates. Journal of Fluid Mechanics 608, 1–20. Ketabdari, M.J., Varjavand, I., 2008. Reflected energy spectrum from slotted
Chwang, A.T., 1983. A porous-wavemaker theory. Journal of Fluid Mechanics 132, breakwaters due to irregular waves. Journal of Coastal Research 24 (6),
395–406. 1529–1535.
Chwang, A.T., Dong, Z., 1984. Wave-trapping due to a porous plate. In: Proceedings of Kondo, H., 1979. Analysis of breakwaters having two porous walls, Coastal
the 15th Symposium on Naval Hydrodynamics, Session 6, Hamburg, pp. 32–42. Structures ’79. ASCE, Alexandria, VA, pp. 962–977.
Chwang, A.T., Chan, A.T., 1998. Interaction between porous media and wave Kono, T., Tsukayama, S., 1981, Wave deformation on a barrier reef. In: Proceedings
motion. Annual Review of Fluid Mechanics 30, 53–84. of the Japan Society of Civil Engineering, vol. 307 (in Japanese).
Clauss, G.F., Habel, R., 2000. Artificial reefs for coastal protection—transient Kriebel, D.L., 1992. Vertical wave barriers: wave transmission and wave forces, The
viscous computation and experimental evaluation. In: Proceedings of the 23th Coastal Engineering Conference. ASCE, Venice, pp. 1313–1326.
27th International Conference on Coastal Engineering, pp. 1799–1812.
Krishnakumar, C., Sundar, V., Sannasiraj, S.A., 2010a. Pressures and forces due to
Evans, D.V., 1990. The use of porous screens as wave dampers in narrow wave
directional waves on a vertical wall fronted by wave screens. Applied Ocean
tanks. Journal of Engineering Mathematics 24, 203–212.
Research 32, 1–10.
Franco, L., 1994. Vertical breakwaters: the Italian experience. Coastal Engineering
Krishnakumar, C., Sundar, V., Sannasiraj, S.A., 2010b. Hydrodynamic performance
22, 31–55.
of single- and double-Wave screens. Journal of Waterway, Port, Coastal, and
Franco, L., de Gerloni, M., Passoni, G., Zacconi, D., 1998.Wave forces on solid and
Ocean Engineering 36 (1), 59–65.
perforated caisson breakwaters: comparison of field and laboratory measure-
Li, Y.C., 2007. Interaction between waves and perforated-caisson breakwaters. In:
ments. In: Proceedings of the 26th Coastal Engineering Conference, Copenha-
Proceedings of the Fourth International Conference on Asian and Pacific
gen, Denmark, pp. 1945–1958.
Coasts, 2007, Nanjing, China, pp. 1–16.
Fugazza, M., Natale, L., 1992. Hydraulic design of perforated breakwaters. Journal
Li, Y.C., Chen, X.F., Sun, D.P., Liu, Y., Jiang, J.J., Ma, B.L., 2005a. The calculation of
of Waterway, Port, Coastal, and Ocean Engineering—ASCE 118 (1), 1–14.
horizontal wave forces on perforated caisson with top cover. China Offshore
Gardner, J.D., Townend, T.H., Fleming, C.A., 1986. The design of a slotted vertical
Platform 20 (1), 1–6 (in Chinese with English abstract).
screen breakwater, The 20th Coastal Engineering Conference. ASCE, Taipei
Li, Y.C., Dong, G.H., Liu, H.J., Sun, D.P., 2003a. The reflection of oblique incident
pp. 1881–1893.
waves by breakwaters with double-layered perforated wall. Coastal Engineer-
Goda, Y., 1988. Statistical variability of sea state parameter as a function of
ing 50, 47–60.
spectrum. Coastal Engineering in Japan 31 (2), 39–52.
Goda, Y., 2000. Random Seas and Design of Maritime Structures. World Scientific Li, Y.C., Jiang, J.J., Ma, B.L., Sun, D.P., Liu, Y., 2005b. Calculation of irregular wave
Publishing, Singapore. forces acting on perforated caisson. China Offshore Platform 20 (2), 12–19
Grune, J., Kohlhas, S., 1974. Wave transmission through vertical slotted walls. 14th (in Chinese with English abstract).
Coastal Engineering Conference, vol. III. ASCE, New York, NY, pp. 1906–1923. Li, Y.C., Liu, H.J., Sun, D.P., 2003b. Analysis of wave forces induced by the interaction of
Hagiwara, K., 1984. Analysis of upright structure for wave dissipation using oblique incident waves with partially-perforated caisson structures. Journal of
integral equation, Proceedings of the 19th Coastal Engineering Conference. Hydrodynamics Series A 18 (5), 553–563 (in Chinese with English abstract).
ASCE, Houston, pp. 2801–2826. Li, Y.C., Liu, Y., Teng, B., 2006. Porous effect parameter of thin permeable plates.
Hayashi, T., Hattori, M., Kano, T., Shirai, M., 1968. Closely spaced pile breakwater as Coastal Engineering Journal 48 (4), 309–336.
a protect structure against beach erosion. Coastal Engineering in Japan 11, Li, Y.C., Liu, H.J., Teng, B., Sun, D.P., 2002. Reflection of oblique incident waves by
149–160. breakwaters with partially-perforated wall. China Ocean Engineering 16 (3),
Hayashi, T., Kano, T., Shira, M., 1966. Hydraulic research on the closely spaced 329–342.
piled breakwater, Proceedings of the 10th Coastal Engineering Conference. Linton, C.M., Evans, D.V., 1990. The interaction of waves with arrays of vertical
ASCE pp.728–884. circular cylinders. Journal of Fluid Mechanics 215, 549–570.
Huang, Z.H., 2006. A method to study interactions between narrow-banded Liu, H.W., Ghidaoui, M.S., Huang, Z.H., Yuan, Z.D., Wang, J., 2010. Numerical
random waves and multi-chamber perforated structures. Acta Mechanica investigation of the interactions between solitary waves and pile breakwaters
Sinica 22 (4), 285–292. using BGK-based methods. Computers and Mathematics with Applications.
Huang, Z.H., 2007a. An experimental study of wave scattering by a vertical slotted doi:10.1016/j.camwa.2010.06.012.
barrier in the presence o f a current. Ocean Engineering 34 (5–6), 717–723. Liu, Y., Li, Y.C., Teng, B., 2007a. The reflection of oblique waves by an infinite
Huang, Z.H., 2007b. Wave interaction with one or two rows of closely spaced number of partially perforated caissons. Ocean Engineering 34, 1965–1976.
rectangular cylinders. Ocean Engineering 34, 1584–1591. Liu, Y., Li, Y.C., Teng, B., 2007b. Wave interaction with a new type perforated
Huang, Z.H., 2008. Wave scattering by double slotted barriers in a steady current: breakwater. Acta Mechanica Sinica 23, 351–358.
experiments. China Ocean Engineering 22 (2), 205–214. Liu, Y., Li, Y.C., Teng, B., 2007c. Wave interaction with a perforated wall breakwater
Huang, Z.H., Ghidaoui, M.S., 2007. A model for the scattering of long waves by with a horizontal porous plate. Ocean Engineering 34, 2364–2373.
slotted breakwaters in the presence of currents. Acta Mechanica Sinia 23, 1–9. Liu, Y., Li, Y.C., Teng, B., Ma, B.L., 2007d. Reflection of regular and irregular waves
Huang, Z.H., Liu, C.R., 2008. A linear theory for wave scattering by double slotted from a partially-perforated caisson breakwater with a rock-filled core. Acta
barriers in weak steady currents. China Ocean Engineering 22 (2), 215–226. Oceanologica Sinica 26 (3), 129–141.
Z.H. Huang et al. / Ocean Engineering 38 (2011) 1031–1053 1053

Liu, Y., Li, Y.C., 2006a. Wave interaction with a modified Jarlan-type perforated Suh, K.D., Jung, H.Y., Pyun, C.K., 2007. Wave reflection and transmission by
breakwater. In: Proceedings of the Seventh ISOPE Pacific/Asia Offshore curtainwall-pile breakwaters using circular piles. Ocean Engineering 34,
Mechanics Symposium, Dalian, pp. 274–281. 2100–2106.
Liu, Y., Li, Y.C., Teng, B., Jiang, J.J., Ma, B.L., 2008. Total horizontal and vertical forces Suh, K.D., Shin, S., Cox, D.T., 2006a. Hydrodynamic characteristics of curtain-wall-
of irregular waves on partially perforated caisson breakwaters. Coastal pile breakwaters. Journal of Waterway, Port, Coastal, and Ocean Engineering
Engineering 55, 537–552. 132 (2), 83–96.
Liu, Y., Li, Y.C., Teng, B., Jiang, J.J., 2006b. Experimental and theoretical investiga- Suh, K.D., Son, S.Y., 2003. On the calculation of irregular wave reflection from
tion of wave forces on a partially-perforated caisson breakwater with a rock- perforated-wall caisson breakwaters using a regular wave model. Journal of
filled core. China Ocean Engineering 20 (2), 179–198. Korean Society of Coastal and Ocean Engineers 15 (1), 1–20.
Liu, H.J., 2003. The Interaction between oblique incident waves and caisson Suh, K.D., Park, J.K., Park, W.S., 2006b. Wave reflection from partially perforated-
structures with perforated wall. Doctoral Thesis. Dalian University of Technol- wall caisson breakwater. Ocean Engineering 33, 264–280.
ogy (in Chinese with English abstract). Suh, K.D., Park, W.S., 1995. Wave reflection from perforated-wall caisson break-
Liu, H.J., Liu, Y., Li, Y.C., 2009. The theoretical study on diagonal wave interaction
waters. Coastal Engineering 26, 177–193.
with perforated-wall breakwater with rock fill. Acta Oceanologica Sinica 28
Suh, K.D., Choi, J.C., Kim, B.H., Park, W.S., Lee, K.S., 2001. Reflection of irregular
(6), 103–110.
waves from perforated-wall caisson breakwaters. Coastal Engineering 44,
Losada, I.J., Losada, M.A., Baquerizo, A., 1993. An analytical method to evaluate the
141–151.
efficiency of porous screens as wave dampers. Applied Ocean Research 15,
Sundar, V., Subba rao, B.V.V., 2002. Hydrodynamic pressures and forces on
207–215.
quadrant front face pile supported breakwater. Ocean Engineering 29,
Ma, B.L., 2004. Wave interaction with perforated vertical wall breakwater. Master
Thesis. Dalian University of Technology (in Chinese with English abstract). 193–214.
Mani, J.S., Jayakumar, S., 1995. Wave transmission by suspended pile breakwaters. Tabet-Aoul, E., Lambert, E., 2003. Tentative new formula for maximum horizontal
Journal of Waterway, Harbor, Coastal, and Ocean Engineering 121 (6), 335–338. wave forces acting on perforated caisson. Journal of Waterway, Port, Coastal,
Martin, P.A., Dalrymple, R.A., 1988. Scattering of long waves by cylindrical and Ocean Engineering—ASCE 129 (1), 34–40.
obstacles and grating using matched asymmetric expansions. Journal of Fluid Takahashi, S., 1996. Design of Vertical Breakwaters, Reference Document No. 34.
Mechanics 188, 465–490. Port and Harbour Research Institute, Japan.
Massel, S.R., Mei, C.C., 1977. Transmission of random wind waves through Takahashi, S., Tanimoto, K., Shimosako, S., 1992. Experimental study of impulsive
perforated or porous breakwaters. Coastal Engineering 1, 63–78. pressures on composite breakwaters. Report of Port and Harbour Research
Mei, C.C., 1985. Scattering of solitary wave at abrupt junction. Journal of Water- Institute 31 (5), 35–74.
way, Port, Coastal and Ocean Engineering 111 (2), 319–328. Tanimoto, K., Yoshimoto, Y., 1982. Theoretical and experimental study of reflec-
Mei, C.C., Liu, P.L.-F., Ippen, A.T., 1974. Quadratic loss and scattering of long waves. tion coefficient for wave dissipating caisson with a permeable front wall.
Journal of Waterway, Harbors and Coastal Engineering Division—ASCE 100 Report of the Port and Harbour Research Institute 21 (3), 44–77 (in Japanese
(WW3), 217–239. with English abstract).
Mei, C.C., 1983. The Applied Dynamics of Ocean Surface Waves. Wiley, New York Tanimoto, K., Takahashi, S., 1994. Design and construction of caisson
740 pp. breakwaters—the Japanese experience. Coastal Engineering 22, 57–77.
Oumeraci, H., Koether, G., 2009. Hydraulic performance of a submerged wave Teng, B., Zhang, X.T., Ning, D.Z., 2004. Interaction of oblique waves with infinite
absorber for coastal protection. In: Lynett, P. (Ed.), Non-linear Wave Dynamics. number of perforated caissons. Ocean Engineering 31, 615–632.
World Scientific Publishing, pp. 31–65. Terret, F.L., Osorio, J.D.C., Lean, G.H., 1968. Model studies of a perforated
Oumeraci, H., 2010. Nonconventional wave damping structures. In: Kim, Y.C. (Ed.), breakwater, Proceedings of the 11th Coastal Engineering Conference. ASCE,
Handbook of Coastal and Ocean Engineering. World Scientific. pp. 1104–1120.
Ou-Yang, H.T., Huang, L.H., Hwang, W.S., 1997. The interference of a semi- Twu, S.W., Lin, D.T., 1991. On a highly effective wave absorber. Coastal Engineering
submerged obstacle on the porous breakwater. Applied Ocean Research 19, 15, 389–405.
263–273. Urashima, S., Ishizuka, K., Kondo, H., 1986. Energy dissipation and wave force at
Park, W.S., Chun, I.S., Lee, D.S., 1993. Hydraulic experiments for the reflection
slotted wall, Proceedings of the 20th Coastal Engineering Conference. ASCE,
characteristics of perforated breakwaters. Journal of Korean Society of Coastal
pp. 2344–2352.
and Ocean Engineers 5 (3), 198–203 (in Korean with English abstract).
Wang, K.H., Ren, X., 1994. An effective wave-trapping system. Ocean Engineering
Park, W.S., Kim, B.H., Suh, K.D., Lee, K.S., 2000. Scattering of irregular waves by
21 (2), 155–178.
vertical cylinders. Coastal Engineering Journal 2 (2), 253–271.
Wiegel, R.L., 1960. Transmission of waves past a rigid vertical thin barrier. Journal
Rageh, O.S., Koraim, A.S., 2010. Hydrulic performance of vertical walls with
of Waterway, Harbor, Division—ASCE 86 (1), 1–12.
horizontal slots used as breakwater. Coastal Engineering 57 (8), 745–756.
Williams, A.N., Mansour, A.E.M., Lee, H.S., 2000. Simplified analytical solutions for
Sahoo, T., Lee, M.M., Chwang, A.T., 2000. Trapping and generation of waves by
vertical porous structures. Journal of Engineering Mechanics—ASCE 126 (10), wave interaction with absorbing-type caisson breakwaters. Ocean Engineering
1074–1082. 27, 1231–1248.
Sawaragi, T., Iwata, K., 1978. Wave attenuation of a vertical breakwater with two Yip, T.L., Chwang, A.T., 2000. Perforated wall breakwater with internal horizontal
air chambers. Coastal Engineering in Japan 21, 63–74. plate. Journal of Engineering Mechanics—ASCE 126 (5), 533–538.
Shepsis, V., Carter, J.D., Fenical, S., Tirindelli, M., 2007. Recent experience in Yu, X.P., 1995. Diffraction of water waves by porous breakwaters. Journal of
analysis and design of perforated vertical breakwaters. In: Proceedings of Waterway, Port, Coastal, and Ocean Engineering—ASCE 121 (6), 275–282.
the Ports 2007, San Diego, California, pp. 1–10. Yu, X.P., Chwang, A.T., 1994. Water waves above submerged porous plate. Journal
Sollitt, C.K., Cross, R.H., 1972. Wave transformation through permeable break- of Engineering Mechanics—ASCE 120 (6), 1270–1282.
waters. In: Proceedings of the 13th International Conference on Coastal Zhu, D., Zhu, S., 2010. Impedance analysis of hydrodynamic behaviors for a
Engineering, pp. 1827–1846. perforated-wall caisson breakwater under regular wave orthogonal attack.
Stiassnie, M., Naheer, E., Boguslavsky, I., 1984. Energy losses due to vortex Coastal Engineering 57 (8), 722–731.
shedding from the lower edge of a vertical plate attached by surface waves. Zhu, S., Chwang, A.T., 2001. Investigation on the reflection behavior of a slotted
In: Proceedings of the Royal Society of London A, vol. 396, pp. 131–142. wall. Coastal Engineering 43, 93–104.

You might also like