You are on page 1of 15

Coastal Engineering 136 (2018) 1–15

Contents lists available at ScienceDirect

Coastal Engineering
journal homepage: www.elsevier.com/locate/coastaleng

Large impulsive forces on recurved parapets under non-breaking waves. A


numerical study
M. Castellino a, *, P. Sammarco b, A. Romano a, L. Martinelli c, P. Ruol c, L. Franco d,
P. De Girolamo a
a
“Sapienza” University of Rome, DICEA, Via Eudossiana, 18, Rome, 00184, Italy
b
University of Rome “Tor Vergata”, DICII, Via del Politecnico, 1, Rome, 00133, Italy
c
Padua University, ICEA Department, Via Ognissanti, 39, Padua, 35129, Italy
d
Roma Tre University, Engineering Department, Via V. Volterra, 62, Rome, 00146, Italy

A R T I C L E I N F O A B S T R A C T

Keywords: This paper describes 2-D numerical simulations of velocity and pressure fields generated by non-breaking waves
Non-breaking waves on a vertical breakwater with a recurved parapet wall. The influence of the geometrical characteristics of the
Vertical breakwaters parapet is investigated. An impulsive pressure force is identified and discussed with respect to the pure vertical
Recurved parapet wall wall case. This force is generated by the seaward flow confinement induced by the surging wave crest. We refer to
Impulsive forces this impulsive impact as “confined-crest impact”.
A large part of the vertical wall is affected by an impulsive increase in pressure caused by pulsating wave,
compared to the case where the parapet is completely vertical. The maximum values of the impulsive pressures
are localized under the recurved parapet. The total force increase on the entire structure may be significant when
compared to the pure vertical wall case.

1. Introduction and state of the art scientific literature, which however has mainly dealt with the use of
curved parapets or overhangs for seawalls. Seawalls are characterized by
Composite vertical breakwaters are a type of coastal structure often shallow water conditions at the toe of the structure and therefore the
used in deep water conditions to protect harbours from incoming waves incoming waves are generally broken or nearly breaking. In these con-
(typically non-breaking). In order to safely use the port-side of these ditions the wave impact contribution of the overhang is not easily
breakwaters, it is crucial to limit wave overtopping. To this purpose, a detectable.
high vertical crown parapet wall is a very efficient and economical design The aim of the present work is to analyze the wave impact on a
solution. To improve the hydraulic efficiency of the crownwall, the recurved parapet of a vertical breakwater placed at a large water depth.
seaward side of the top profile may be shaped to form an overhang Therefore, in these conditions the presence of broken or breaking waves
(recurved parapet/wave return wall/bullnose) aimed at further reducing can be excluded.
wave overtopping by deflecting back seaward the uprushing water (Fig. 1 The scientific literature dealing with impulsive pressures and forces
illustrates the parameters definition from EurOtop, 2016). on vertical or inclined walls with or without recurved parapets or over-
Curved parapets at breakwater crownwalls and quay walls have been hangs is very extensive.
efficiently used since the ancient Roman times (see Fig. 2). A first review concerning wave impacts on bodies and coastal struc-
An overhang or a curved parapet located in the upper part of a tures is provided by Ramkema (1978) who highlighted that the devel-
crownwall yields two effects: (i) reduces wave overtopping if compared opment of mathematical wave impact and slamming models was
to a pure vertical wall characterized by the same value of the crest originated by Von Karman (1929) and Wagner (1932). They were
freeboard Rc (see Fig. 1); (ii) generates a wave impact on the underside interested in the impact of seaplane floats during landing and taking off.
part of the overhang. This wave impact produces impulsive pressures Ramkema (1978) developed the so-called Bagnold's piston model (Bag-
acting both on the overhang and on the vertical wall. nold, 1939). The model was used to describe the wave impact caused by
These two effects have been largely considered by the technical and standing waves against protruding structural elements inserted in the

* Corresponding author.
E-mail address: myrta.castellino@uniroma1.it (M. Castellino).

https://doi.org/10.1016/j.coastaleng.2018.01.012
Received 27 May 2017; Received in revised form 18 January 2018; Accepted 28 January 2018

0378-3839/© 2018 Published by Elsevier B.V.


M. Castellino et al. Coastal Engineering 136 (2018) 1–15

Fig. 1. Parameters definition for assessment of overtopping


at structures with parapet/wave return wall (EurOtop,
2016).

Fig. 2. Left panel: curved seawall in a


Roman mosaic of 2nd century AD in
Rimini (Italy). Right panel: cross section
of the rock carved breakwater of the
Roman port of Ventotene built under
Augustus on 14 BCE and still operational
(from Franco, 1996).

caissons of the storm surge barrier of the Eastern Scheldt. In the geometry water is addressed by Gibson (1970) while the influence of salinity in the
considered by Ramkema (1978) a layer of air between the protruding bubble size in water is studied by Scott (1975a, b). The importance of
element and the water mass was present. scale effects and aeration in simulating violent breaking impacts are
The basic concepts of wave impact beneath a horizontal surface are considered by Blenkinsopp and Chaplin (2011) and Bredmose et al.
given by Wood and Peregrine (1999). They considered the wave impact (2015).
on the underside of a projecting surface, e.g. a flat deck, close to the mean Recent applied research concerning breaking wave loads and their
water level. The horizontal surface of finite length is confined on at one effects at vertical seawalls and at caisson breakwaters may be found in
side by a vertical wall. Wood and Peregrine (1999) showed that the Cuomo et al. (2010a, b) and Elsafti and Oumeraci (2017). Chen (2011);
pressure-impulse (e.g. the integral of pressure over the duration of the Chen et al. (2015) and Van Doorslaer et al. (2017) considered the impact
impact) generated by the wave action under the horizontal surface affects on stormwalls induced by overtopping flow.
also the vertical wall. An extensive review of curved seawall shapes can be found in Anand
Most of research concerning pressures and forces have addressed the et al. (2010). Such curved shapes have been extensively used for onshore
violent breaking wave impacts on vertical sea walls placed in shallow seawalls under breaking and broken wave attack.
water, where breaking of wind waves is induced by bottom effects One of the earliest studies on the application of curved crownwalls on
(Bagnold, 1939; Cooker and Peregrine, 1990, 1992; Peregrine, 2003). composite vertical breakwaters was conducted by De Gerloni et al.
Hattori et al. (1994) observed that pressure waves generated by the wave (1989) for the deep water perforated caissons at Porto Torres, testing
impacts, propagate away from the impact zone at the speed of sound, i.e. three different recurve shapes with excellent overtopping reduction.
as acoustic waves in the fluid. Other findings regarding the characteris- Even field measurements of pressures acting on the Porto Torres caissons
tics of violent breaking wave impacts may be found in the works of Wood (see Fig. 3) were then carried out by De Girolamo et al. (1996), but no
et al. (2000), Bullock et al. (2007), Bredmose et al. (2009), Plumerault large events were recorded in the limited measurement period. General
et al. (2012) and Bredmose et al. (2015). These works (see Bredmose formulations for overtopping reduction factors, as compared to the pure
et al., 2015, for a comprehensive state of art) deal with the strong vertical wall, were then proposed by Franco et al. (1995), while recent
sensitivity of the impact pressures on the wave shape and the oscillatory effective applications were described by Di Risio et al. (2007, 2009) and
temporal pressure variations due to compression and expansion of trap- Franco et al. (2013). A systematic investigation on chamfered parapets
ped and entrained air. In these situations, compressibility effects need to was conducted by Cornett et al. (1999). Early work for curved seawalls
be taken into account. The effect of air bubbles on the speed of sound in was reported by Owen and Steele (1993). An extreme Flaring Shaped

2
M. Castellino et al. Coastal Engineering 136 (2018) 1–15

measured by using tactile sensors (Ravindar et al., 2016).


Kisacik et al. (2012a, b, 2014) describe and analyze the results of
physical model tests reproducing an existing structure on the Belgian
coast. The structure, which may be classified as a seawall, is made of a
vertical wall with an overhanging horizontal cantilever slab whose
function is to reduce wave overtopping. The structure is exposed to vi-
olent wave impacts, including waves running up against the vertical wall
and slamming beneath the horizontal deck. The experimental tests
included a wide range of incoming wave conditions and water levels,
covering all possible types of breaking conditions and considered also the
flip-through impact (Cooker and Peregrine, 1990). Most of the tests were
characterized by shock pressures recorded on the two parts of the
structure. The first impact occurs on the vertical part, while the second
one occurs beneath the horizontal part. The authors excluded from their
analysis waves hitting first the horizontal part.
Despite significant research work on forces acting on vertical and
recurved walls, the mechanisms determining the effectiveness of a
recurved parapet and the associated forces are complex and not yet fully
described. There is still a lack of design guidelines and consolidated
engineering standards for the relatively frequent case of recurved walls,
especially in deep water applications at the crown of vertical breakwa-
ters. This may lead to a significant understimations of wave forces
induced by non-breaking waves and to the consequent damages, as that
occurred recently at the Civitavecchia Harbour (see Fig. 4).
Indeed in this work we infer that the impact of a non-breaking wave
Fig. 3. Picture of the main breakwater of Industrial Port of Porto Torres,
Sardinia (Italy) made by r.c. precast perforated caissons with a on the recurved part of the wall generates pressure shock waves which
recurve crownwall. propagate all along the vertical structure with the sound speed, drasti-
cally increasing the forces acting on the entire wall. Unlike the case of a
Seawall with thin overhanging crest was first presented by Murakami pure vertical wall, when a recurved parapet is considered, the pressures
et al. (1997). and forces reach their maximum values at the top of the structure where
To date, the formulae for estimating the magnitude of impulsive and the recurve is located. These effects may be important for the design of
non-impulsive pressure fields, acting on composite vertical breakwaters the structure and in particular for the design of the top parapet wall.
and generated by breaking and non-breaking waves, are largely derived Indeed such walls may be slender, i.e. characterized by a small inertia
from results of laboratory tests. The most widely used design formulae and high stiffness, and therefore may be much more sensible to impulsive
are summarized by Goda (2010). These formulae include impulsive forces than the rest of the vertical structure which is normally made by
breaking conditions studied by Takahashi (1996). massive precast caissons. As previously mentioned, this case has not been
Oumeraci et al. (2001) gave guidelines for the design of composite analyzed by Kisacik et al. (2012a, b, 2014).
vertical breakwaters and seawalls under breaking and non-breaking
conditions within the framework PROVERBS (PRObabilistic design
tools for VERtical BreakwaterS). PROVERBS also includes guidelines to
estimate wave impact magnitude and duration, and their effects on
caisson breakwaters (Martinelli and Lamberti, 2011). No recurved
parapet walls were considered.
The first works that highlighted the effects of recurved walls on the
total force acting on a coastal structure, due to breaking and non-
breaking waves, are those of Kortenhaus et al. (2002, 2003). They
investigated experimentally the effects of parapets and recurves on wave
overtopping and wave loading. The authors found that the horizontal
force increase is about 1.7 for impulsive breaking wave conditions and
2.0 for non-breaking conditions. However, they state that “it is not
possible to distinguish whether the load increase is due to the increase of
wall height or to the shape of the parapet installed at the wall”.
Pearson et al. (2004) studied experimentally a promenade wall with
an “unusual parapet” made by an overhang to reduce wave overtopping.
They concluded that the overhang increases the horizontal loads of a
factor around two as compared to the loads on a vertical wall. Further-
more, they pointed out that the upward loads on the underside of the
parapet are comparable in terms of magnitude, and even larger in some
cases, when compared to the horizontal ones.
More recently, Stagonas et al. (2014) performed an extensive exper-
imental campaign within the Hydralab IV project investigating both
wave overtopping and forces acting on a recurved seawall in breaking
waves conditions. They pointed out that “the force acting on the recurved
parapet increases as the seaward extent of the wall increases”. The spatial
distribution of impact pressures on the recurved part of the seawall was Fig. 4. Picture of the damages occurred at the Civitavecchia Harbour,
Lazio (Italy).

3
M. Castellino et al. Coastal Engineering 136 (2018) 1–15

In the present work numerical simulations of non-breaking wave order to focus on the hydrodynamic process related to the parapet wall, a
conditions on recurve and vertical parapet walls are carried out in order pure vertical breakwater on a flat bottom is assumed, i.e. no rubble
to analyze the hydrodynamics generating the impulsive forces. The mound foundation is considered. The flat bottom assumption also allows
maximum forces on a recurve wall are evaluated and compared to the to use the Sainflou analytical solution (Sainflou, 1928) to validate the
forces acting on a pure vertical wall characterized by the same freeboard. numerical setup under regular wave conditions. No sea level variations
The trade-off between overtopping reduction and impulsive forces is also are considered. The water depth at the toe of the structure is constant and
investigated. equal to 20.0 m with respect to S.W.L. This water depth reproduces the
The paper is structured as follows: first, the numerical experiment typical deep water conditions of vertical caisson breakwaters along the
assumptions and the implementation and validation of the numerical Mediterranean coasts, where flip-through impacts, breaking waves or
simulations are given, then the numerical results are shown and broken waves at the breakwater toe may be statistically excluded during
discussed. their lifetime.
The sea bottom and the structures are assumed to be impermeable.
2. Basic assumption Entrained or trapped air is simulated by the applied numerical code, even
though water and air are assumed to be incompressible. These assump-
A vertical breakwater with a parapet wall located on its crown is tions (i.e. water and air incompressibility) are justified by the following
considered. The seaside of the parapet wall and of the vertical breakwater reasons: (i) it has been numerically verified that there is no trapped air in
are vertically aligned. As shown in Fig. 5, the sea side of the considered the fluid or between the fluid and the structure during the phenomenon;
parapet wall has two shapes: (i) a recurved shape (upper panel); (ii) a (ii) the fluid velocity is always much smaller than the speed of sound in
vertical shape (lower panel). Fig. 5a shows the main characteristics of the the fluid (Korobkin and Pukhnachov, 1988; Cooker and Peregrine, 1992;
recurved parapet: the curve, a 90∘ sector of circumference, is placed at the Mei et al., 2005).
top of the parapet with a radius of 1.0 m. A comparison between the two different parapet wall geometries is
The elevation of the parapet wall crown (crest freeboard Rc ) is equal carried out in order to estimate the influence of the crown's recurve of the
to þ6:5 m (above the Still Water Level - S.W.L.) for both geometries. In parapet wall on the hydrodynamics. With reference to a recurved parapet

Fig. 5. Sketch of the two vertical breakwaters ge-


ometries. a) recurved wall, b) vertical wall. Rc is the
same for both walls. Elevations are in meters. The
numerical flume length is 100.0 m and the depth is
constant and equal to 20.0 m.

4
M. Castellino et al. Coastal Engineering 136 (2018) 1–15

and non-breaking regular waves, two types of hydrodynamic flows Table 1


induced by the interaction between incoming waves and the vertical Simulated incident wave conditions and crest freeboard. First column: name of the simu-
lations. Second column: nominal wave height H. Third column: nominal wave period T.
breakwater have been observed:
Fourth column: nominal dimensionless freeboard Rc =H. Fifth column: nominal steepness
s ¼ H=L0 . Sixth column: effective wave height H. Seventh column: effective wave period T.
 Case 1: the formation of a wave “clapotis” (standing wave) induced by Eighth column: effective dimensionless freeboard Rc =H. Ninth column: effective steepness
a reflection phenomenon; in this case, the incoming wave crest does s ¼ H=L0 .
not reach the recurved part of the parapet wall; Wave Nominal Effective
 Case 2: the formation of impulsive pressures (i.e. shock pressure name
H T Rc =H s H T Rc =H s
waves) and forces originating from the top recurved part of the wall;
(m) (s) () () (m) (s) () ()
the shock pressures are induced by the sudden confinement caused by
the incoming wave crest. We refer to this case as “confined-crest W1 1.0 8.0 6.50 0.010 0.98 8.13 6.63 0.009
W2 2.0 8.0 3.25 0.020 1.86 8.13 3.49 0.018
impact”.
W3 3.0 8.0 2.16 0.030 2.93 8.13 2.21 0.029
W4 4.0 8.0 1.62 0.040 3.97 8.19 1.63 0.039
Cases 1 and 2 have been sequentially obtained by gradually W5 5.0 8.0 1.30 0.050 4.95 8.19 1.31 0.049
increasing wave height for fixed period of the incident regular waves. For W6 6.0 8.0 1.08 0.060 6.06 8.19 1.07 0.060
a vertical parapet wall, Case 2 is never observed. W7 7.0 11.0 0.92 0.037 7.17 10.83 0.90 0.038

3. Implementation and validation of the numerical simulation reported and in the fifth the wave steepness, where the wave length L0
has been calculated using the linear theory in deep water condition. In
Numerical simulations are performed using OpenFOAM®, which in- the eighth and ninth columns effective dimensionless freeboard and
cludes the incompressible three-dimensional RANS conservation equa- steepness are reported. In the following analysis, we refer to the effective
tions. These equations are based on the Reynolds decomposition that values of the wave characteristics. The effective incoming wave heights
identifies an average and a fluctuating component. The k - ε turbulence have been computed by means of the method proposed by Isaacson
model is used to close the system equations, with k the turbulent kinetic (1991) which is suitable for regular reflected waves. The effective wave
energy and ε the turbulence dissipation rate. periods have been calculated by means of a spectral analysis.
The free surface is tracked by the Volume of Fluid method (VOF, Further decrease of the mesh size have been studied only for the W5
Berberovic et al., 2009). We used IHFOAM, a specific numerical solver wave condition on the recurved configuration. By analyzing the results
well suited for coastal hydrodynamic processes. The solver is able to for increasing mesh resolution we have not noticed any improvement in
reproduce wave generation and active absorption (Higuera et al., 2013a, accuracy.
b). The active absorption system of the reflected waves from the structure The incident wave conditions W1, W2 and W3 produce a wave cla-
is crucial, since vertical breakwaters are characterized by high reflection potis in front of the structure for both vertical and recurved parapet (see
coefficients. This tool has been successfully used in several coastal en- Fig. 6). Starting from incident wave condition W4 the confined-crest
gineering applications (e.g. Higuera et al., 2014a, b). impact takes place on the recurved parapet. This phenomenon becomes
The IHFOAM model has been used to numerically reproduce an more evident in terms of the increase of pressure and force with respect
experimental wave flume. The dimensions of the numerical wave flume to the vertical parapet, when the wave conditions W5, W6 and W7 are
are 100.0 m in length, 35.0 m in height and 0.10 m in width (see again considered.
Fig. 5). We use subscripts R and V respectively for the Recurved wall and the
Different mesh setups have been considered to assess the influence of Vertical wall. Subscript max indicates the maximum value attained by the
the mesh resolution on the results. The chosen mesh is characterized by forces, while the apex v and h respectively indicate the vertical and the
an initial cell resolution equal to 0.20 m along x-direction and 0.15 m horizontal components of the forces. By using this notation we list the
along z-direction, gradually decreasing to 0.03 m along x and z directions following symbols:
near the wall.
All the 2D simulations carried out are characterized by regular non-  FVmax maximum value of the total force on the structure with the
breaking waves to enucleate the physics of the phenomenon. The inte- vertical parapet;
gration time step is automatically chosen by the code to ensure numerical
stability and, for the tests here presented, ranges between 4:0  104 s
and 1:0  103 s. This aspect is crucial both for numerical models and
physical experiments because impulsive forces occur in a very short time.
The Courant number was set to a maximum value of 0.45. Numerical
simulations were performed by generating on the West side of the nu-
merical flume (see Fig. 5) regular waves of period T equal to 8.0 s and
11.0 s with varying wave heights H (minimum wave height of 1.0 m and
maximum of 7.0 m). Few additional tests up to H ¼ 8:0 m and T ¼ 13 s
were also performed in the final analysis described in chapter 5. An active
absorption boundary condition has been imposed on the West side of the
flume. Along the solid impermeable boundaries (walls and bottom) the
numerical code assumes a no-slip velocity condition.
The mean computational time for each simulation is of the order of
5 h (some 15 zerocrossing regular waves have been simulated) on an Intel
Core i7-6700K computer equipped with 64 GB of RAM and 8 processors.
The simulated regular and non-breaking incoming wave character-
istics are summarized in Table 1. The second and third columns of
Table 1 contain the nominal wave heights and periods, while the sixth
and the seventh columns contain the corresponding effective values. In Fig. 6. Time evolution of the free surface elevation η for the numerical
the fourth column of the table, the nominal dimensionless freeboard is simulation W3. Values of the free surface elevation are reported on the z-axis.
Non-breaking wave condition is assured.

5
M. Castellino et al. Coastal Engineering 136 (2018) 1–15

 FRmax maximum value of the total force on the structure with the The values of the above forces are computed: (i) for the part of the
recurved parapet; vertical breakwater located above the S.W.L. (z 2 [ 0.0; þ6.5 ] m); and
 FRv max vertical component of FRmax ; (ii) for the entire vertical structure (z 2 [20.0; þ6.5 ] m).
 FRhmax horizontal component of FRmax ; To verify the numerical implementation, the wave forces acting on
 FR total attained force on the structure with the recurved parapet wall the vertical wall have been compared with those given by the exact
(time-dependent quantity); Sainflou approximated second order wave theory (Sainflou, 1928) for the
 FRv vertical component of FR (time-dependent quantity). W1, W2 and W3 wave conditions. For these three conditions, indeed
 F  ¼ FRmax =FVmax instantaneous normalized force. there is no wave overtopping on the vertical wall, nor the wave crests
reach the recurved part of the parapet (see Table 2). The fourth column of
Table 2 shows the ratio between the maximum total force computed on
Table 2
the entire wall in crest condition (FVmax ) and the corresponding maximum
Characteristics and main results of numerical simulations for z 2 [20.0; þ6.5 ] m. First force by Sainflou's theory (FSainflou ; Sainflou, 1928). The values of this
column: name of the simulation. Second and third columns: regular non-breaking wave ratio, varying between 0.94 and 0.97, imply a good agreement. The
characteristics (height and period). Fourth: ratio of the maximum total forces calculated at clapotis time evolution of the free surface elevation η for W3 wave con-
the vertical wall FVmax with respect to FSainflou .
dition is shown in Fig. 6. Increasing further the incoming wave steepness,
Wave H T FVmax /FSainflou the agreement between the computed and theoretical forces decreases.
name
(m) (s) () This may be explained by observing that Sainflou theory becomes less
precise as wave nonlinearity increases. Similar results have been ob-
W1 1.0 8.0 0.94
W2 2.0 8.0 0.97 tained when the wave trough takes place at the vertical wall.
W3 3.0 8.0 0.96 A further check of the numerical implementation has been carried out

Fig. 7. Vertical wall: comparison between the Goda pressure diagrams and the numerical pressure distribution at the time at which the maximum force occurs.
Each panel refers to the different wave conditions W1 - W7 of Table 1.

6
M. Castellino et al. Coastal Engineering 136 (2018) 1–15

Table 3
α ¼ 90∘ . Characteristics and main results of numerical simulations for z2[0.0;þ6.5]m and z2[20.0;þ6.5]m. First column: name of the simulation. Second to sixth columns: maximum total
forces FRmax , maximum horizontal forces FRhmax , maximum vertical forces FRv max calculated for the recurved wall, maximum force FVmax calculated for the vertical wall, forces ratio
F  ¼ FRmax =FVmax for z2[0.0;þ6.5]m. Seventh to eleventh columns: maximum total forces FRmax , maximum horizontal forces FRhmax , maximum vertical forces FRv max calculated for the recurved
wall, maximum force FVmax calculated for the vertical wall, forces ratio F  ¼ FRmax =FVmax for z2[20.0;þ6.5]m.

Wave z2[0.0;þ6.5]m z2[20.0;þ6.5]m


name 
FRmax FRhmax FRv max FVmax F FRmax FRhmax FRv max FVmax F

(KN/m) (KN/m) (KN/m) (KN/m) () (KN/m) (KN/m) (KN/m) (KN/m) ()

W1 3.29 3.29 0 3.29 1.00 107.30 107.30 0 107.30 1.00


W2 17.39 17.39 0 17.39 1.00 228.10 228.10 0 228.10 1.00
W3 35.13 35.13 0 35.13 1.00 338.30 338.30 0 338.30 1.00
W4 112.50 112.33 6.01 98.92 1.14 419.50 419.46 6.01 415.10 1.01
W5 364.20 352.05 93.25 127.40 2.86 884.90 879.99 93.25 546.60 1.62
W6 479.90 462.26 128.92 169.20 2.84 1179.50 1172.44 128.92 705.70 1.67
W7 719.10 685.95 215.81 252.20 2.85 2051.00 2039.66 215.81 1228.00 1.67

by comparing the forces provided by the standard Goda's formulae Table 3, where the maximum instantaneous values of the forces FRmax
(Goda, 2010) with the forces obtained numerically. We remark that this (recurved wall) and FVmax (vertical wall) acting on the upper part of the
is not a theoretical comparison due to the empirical origin of the Goda's walls (pressure integral from z ¼ 0:0 m to z ¼ Rc ) and the entire walls
formulae. Nevertheless, it is interesting for engineering applications. The (pressure integral from z ¼ 20:0 m to z ¼ Rc ) are given together with
comparison has been made by using for the Hmax and T1=3 of the Goda's their ratio F  ¼ FRmax =FVmax . The horizontal and the vertical components
formulae the H and T values respectively, as reported in Table 1 for each FRhmax and FRv max of the total forces acting on the recurved wall are also
simulation. The comparison has been made only for the vertical parapet shown in the table.
case taking into account both the wave conditions which do not induce Time series of the total force FR and the vertical component of it FRv
wave overtopping (W1, W2, W3 and W4) and those which do (W5, W6 computed on the recurved wall and the total force FV on the vertical wall
and W7). The comparison of the pressure diagrams and of the corre- are represented for the W5 condition in Fig. 8.
sponding maximum forces ratio (FVmax /FGoda ) acting on the entire vertical Forces values in Fig. 8, normalized with respect to the maximum force
breakwater is shown in Fig. 7 for all the simulated wave conditions. The value FVmax attained in the vertical case, are computed on the wall above
ratio FVmax /FGoda resulted between 0.94 and 1.15. This agreement is really the still water level. The black line indicates FR =FVmax , the dashed black
good, considering the Goda's formulae empirical nature. line indicates FRv =FVmax and the dashed red line indicates FV =FVmax .
Fig. 9 shows a zoom of the curves of Fig. 8 around the time t2 (see
4. The nature of the impulsive force Fig. 8) at which the maxima occur. The capability of the automatic time
step selection to resolve rapid pressures and forces time variations is
Numerical simulations have shown that the presence of the recurved highlighted by the curves of Fig. 9 which exhibit a regular profile. Fig. 8
wall induces a very large increase of wave pressure acting on quite a large shows an impulsive force peak of FR (with FRmax =FVmax ¼ 2.86) in a purely
part of the entire structure, compared to the case where the parapet is non-breaking wave condition. The pressure increase has impulsive
completely vertical. characteristics and occurs when the free surface elevation of the wave
The results for the seven different wave conditions are summarized in crest reaches, along the wall, the terminal part of it. We show that this
pressure increase has the characteristics of a pressure shock wave due to a
sudden stop of a water mass in motion.
First of all, the pressure peak occurs at the numerical probes almost at

Fig. 8. Time series of the normalized force on the wall above the mean sea
level (z 2 [ 0.0;6.5 ] m) for the W5 wave condition. Normalized forces ratios
are all taken with respect to the maximum force attained for the vertical wall
FVmax : a) the total force acting on recurved wall FR =FVmax (black line), b) the Fig. 9. Zoom of Fig. 8. Dimensionless time series (W5 wave condition) of the
vertical component of the total force acting on the recurved wall FRv =FVmax forces on the walls for t 2 [ 20.3; 20.6 ] s. Black line: ratio of FR with respect to
(dashed black line), c) the force acting on the vertical wall FV =FVmax (dashed FVmax . Dashed black line: ratio of FRv with respect to FVmax . Dashed red: line ratio
red line). (For interpretation of the references to colour in this figure legend, of FV with respect to FVmax . (For interpretation of the references to colour in
the reader is referred to the Web version of this article.) this figure legend, the reader is referred to the Web version of this article.)

7
M. Castellino et al. Coastal Engineering 136 (2018) 1–15

the same instant (this is due to the assumed fluid incompressibility) all the recurved wall (Fig. 11e) configurations.
along the full vertical structure down to the bottom, as it is shown in Instant t2: Fig. 12d shows how the wave crest, surging from seaward,
Fig. 10. The pressure peak reduces progressively its intensity, moving blocks the water flow deflected offshore by the recurve. The mass of the
from the top of the wall (probe D located on the recurve at þ5:54 m) incident wave, and therefore its momentum flux, confine the returning
toward the base of the wall (probe A located at 10:0 m). This is due to flow, whose maximum negative speed value is in the order of 16 m=s,
the spatial dispersion of the pressure shock wave. introducing a sudden “no flow” constraint. Therefore, a blockage effect
Second, the impulsive pressure occurs at the same time when a may be observed. An entirely new finite area with zero horizontal ve-
massive flow confinement induced by the incoming wave crest suddenly locity suddenly appears at and below the free surface just off the recurved
stops the returning flow induced by the recurved wall. We further clarify edge. Associated to this sudden instantaneous zero horizontal mo-
this aspect. mentum, a large and widespread pressure increase hits the recurved wall
When the wave crest reaches the top recurved part, a high-speed flow extending its influence on part of the vertical wall with a maximum value
runs upward to exit seaward of the recurved part. It is expected that this right under the recurve (see Fig. 12f). This excess pressure will expel the
momentum deflection, induced by the recurve, increases the forces (both flow through the free surface causing the jet of Fig. 13c and d. Under the
vertical and horizontal) acting on the recurved part of the wall. Somehow recurve a smaller vertical velocity component (Fig. 12e) with respect to
less expected is the impulsive force shown in Fig. 8. the vertical wall (Fig. 12b) is also evident. The volume of fluid involved
The impulsive force is generated by the flow confinement induced by in the confined-crest impact is shown by the white and blue area of
the wave crest. This phenomenon, identified in this paper as “confined- Fig. 12d.
crest impact”, can be shown by considering the three instants indicated Instant t3: Fig. 13c shows that now almost all momentum is negative
on the top of Fig. 8: t1 ¼ 20:29 s, t2 ¼ 20:43 s and t3 ¼ 20:83 s. (no horizontal velocity field) and a jet appears. The pressure field is back
At the instant t2 the force peak takes place, while during the time to normal conditions (compare Fig. 13b and d).
interval t3  t1 ¼ 0:54 s the normalized force acting on the recurved wall A minor drop of the total force with respect to the vertical wall case
is higher than the normalized force acting on the vertical wall. The force follows at the instant identified as t4 in Fig. 8. This can be explained as
increases take places in the very short time interval t2  t1 ¼ 0:14 s. follows.
Figs. 11–14 show what happens. To outline the phenomenon we Fig. 14a and c shows two different conditions. In the vertical wall case
consider the horizontal velocity component of the flow. Positive values of the water falls back down along the wall, causing a larger hydrostatic
horizontal velocity component (red) are directed towards the wall, while pressure in the second peak, while the recurve has deflected water
negative values (blue) are directed seaward. In the following analysis seaward, leading to a minor drop.
also vertical velocity component is provided (positive values refer to Fig. 15 shows, at a larger scale, the spatial shape of the free surface at
upward components, while negative values refer to downward ones). several instants for W5 condition, starting from the wave trough. The
Instant t1: Fig. 11a and d shows the horizontal velocity field, while dashed blue line of Fig. 15 corresponds to instant t2 of Figs. 8 and 12,
Fig. 11c and f shows the pressure field for the vertical and recurved wall when the confined-crest impact induces the shock pressure in the fluid
respectively. In Fig. 11d the flow is climbing the curve and a significant originated by the recurved part of the parapet wall. Fig. 15 highlights that
negative horizontal momentum is locally achieved. Over the seaward neither wave breaking nor trapped air occur during the development of
area, the rest of the velocity field is identical to the vertical case (compare the phenomenon. This result has been confirmed for all the wave con-
with Fig. 11a). As for the pressure field, there is only a local increase ditions considered in this paper and also confirm that this phenomenon
accompanying the flow deflection; other than that the two pressure fields can be treated assuming an incompressible fluid.
are identical (see Fig. 11c and f). No substantial differences of the vertical Looking at the maximum vertical forces FRv max of Table 3, the first wave
velocity components can be noticed between the vertical (Fig. 11b) and condition at which a vertical component of the force appears on the

Fig. 10. Time series of the pressure on four probes


along the wall for the W5 wave condition.

8
M. Castellino et al. Coastal Engineering 136 (2018) 1–15

Fig. 11. Instant t1 ¼ 20:29s. Diagrams of the pressure and velocity fields for the W5 wave condition for vertical and recurved wall. a) horizontal velocity field -
vertical wall; b) vertical velocity field - vertical wall; c) pressure field - vertical wall; d) horizontal velocity field - recurved wall; e) vertical velocity field - recurved
wall; f) pressure field - recurved wall.

Fig. 12. Instant t2 ¼ 20:43s. Diagrams of the pressure and velocity fields for the W5 wave condition for vertical and recurved wall. a) horizontal velocity field -
vertical wall; b) vertical velocity field - vertical wall; c) pressure field - vertical wall; d) horizontal velocity field - recurved wall; e) vertical velocity field - recurved
wall; f) pressure field - recurved wall.

9
M. Castellino et al. Coastal Engineering 136 (2018) 1–15

Fig. 13. Instant t3 ¼ 20:83s. Diagrams of the pressure and


velocity fields for the W5 wave condition for vertical and
recurved wall. a) horizontal velocity field - vertical wall; b)
pressure field - vertical wall; c) horizontal velocity field -
recurved wall; d) pressure field - recurved wall.

recurved wall is W4, even though the W4 FRv max is much smaller than the not a function of wave height and period. Apparently the impulsive force
corresponding value attained for W5 condition. increase is, for a given bathymetry, mainly a function of the recurved wall
The analysis of W4 condition, shown in Fig. 16, reveals that also for geometry (i.e. angle).
this condition a small pressure shock wave may be identified because a Finally, Fig. 18 shows a comparison between the asynchronous en-
small pressure peak occurs at the same time at probes D, C and B. As for velopes of the pressure peaks diagrams for the recurved and vertical wall.
W5 condition, the pressure peak reduces progressively its intensity The figure shows the large pressure peak localized at the recurve and the
moving from the curved part of the wall to the base of the wall. At the spread of the impulsive pressures over a large part of the vertical wall.
probe A, located at 10:0 m, the pressure peak almost disappears. Moreover, the recurved shape induces also a vertical component of
Nevertheless, it can be argued that the deflection alone does not give a the force. As shown in Fig. 18, most of the pressure increase induced by
dramatic impulsive force, as shown in Fig. 17. In the upper panels of the the recurved shape of the parapet is localized near the top of the struc-
figure the horizontal velocity field is reported at three different time ture, where the pressure values may be more than ten times the values
steps. In the lower panels, at the same instants, the pressure field is shown obtained for the pure vertical wall. This has a significant influence on the
and it ultimately determines that the W4 condition exhibit the highest structural design of the parapet wall, which may be characterized by
runup of a deflected wave that does not cause any significant pressure small natural period, which in turn might may be comparable to the
(and therefore force) increase. duration of the impulse, so enhancing the induced stress field. Further-
These results show that the confined-crest impact appears for the more, the presence of the upward vertical component of the force, which
threshold value Rc =H ¼ 1:62 (see W4 condition of Table 1) and becomes induces vertical tensile stresses in the parapet structure, may consider-
evident for Rc =H ¼ 1:30 (W5 condition). ably reduce the compression stresses due to the structure weight, thus
The increase of the total forces on the recurved wall above the S.W.L. reducing the overall resistance of the structure itself. Moreover, the
with respect to the vertical wall case (see F* column 6 of Table 3) ranges increased height of the application point of this impulsive force induces
between 1.14 and 2.86, for increasing values of H and T. As far as the higher overturning moments on the structure.
forces evaluated on the entire vertical structure are considered (see F* The impulsive nature of the forces may have a lower influence on the
column 11 of Table 3), the increase ranges from 1.01 to 1.67. entire structure of the vertical breakwater, especially when the latter is
It is worth noticing that for wave heights and periods larger than made of a massive precast caisson, characterized by a large inertia.
those of W4 condition, the ratio F* remains almost constant, i.e. F* it is Nevertheless, the increase of the ratio F  ¼ 1:67 (see Table 3 column 11)

10
M. Castellino et al. Coastal Engineering 136 (2018) 1–15

Fig. 14. Instant t4 ¼ 21:50s. Diagrams of the pressure and


velocity fields for the W5 wave condition for vertical and
recurved wall. a) horizontal velocity field - vertical wall; b)
pressure field - vertical wall; c) horizontal velocity field -
recurved wall; d) pressure field - recurved wall.

may have some influence on the global stability of the structure and on same wave used to evaluate the maximum force for each wave condition.
the geotechnical response of the foundation (e.g. fluidization). The volumes have been obtained by considering a vertical section
localized on the seaward side of the top of the wall. Along this section, the
5. Trade off between overtopping volumes and impulsive forces vertical profile of the horizontal fluid velocity component and the sea
surface level have been measured for each time step, so to obtain the
In order to understand the influence on the wave overtopping and instantaneous overtopping discharge. Its time integration provides the
total forces of the opening angle characterizing the recurved crown wall, overtopping volume.
two further geometrical configurations have been analyzed: α ¼ 70∘ and In order to describe with more details the dependence of the forces
α ¼ 45∘ (upper panel of Fig. 19). Both the additional curved parapets and the overtopping volumes on Rc =H, two intermediate wave conditions
have a radius of 1.0 m, as in the previous case. Therefore, with reference between W4 and W5 have been eventually considered, namely: W4:33
to the horizontal extent of the recurved parapet, defined for this specific and W4:65 (see the upper part of Fig. 19). The incoming wave height is
case as Br ¼ 1  cosðαÞ, independent of the radius (see Fig. 1), we have H ¼ 4:33 m and the wave period is T ¼ 8:0 s for the first condition, while
analyzed three different configurations, Br ¼[1; 0.66; 0.29]. H ¼ 4:65 m and T ¼ 8:0 s for the second one. A further wave condition
The results in terms of forces, for the two new geometrical configu- (W8: H ¼ 8:0 m, T ¼ 13 s, see the upper part of Fig. 19) has been
rations and for the same regular wave conditions (W4, W5, W6 and W7) considered in order to analyze if (and when) the effect of the curved wall
are shown in Table 4 (α ¼ 70∘ ) and 5 (α ¼ 45∘ ). on the overtopping reduction tends to disappear.
Fig. 19 compares the results in terms of normalized forces F  (on the In Fig. 19 circles, squares and diamonds refer to 90∘ , 70∘ and 45∘
left axis) and normalized wave overtopping volumes V  (on the right respectively. Blue lines and empty markers refer to the dimensionless
axis) of the three geometries as a function of the dimensionless freeboard. forces, while red lines and filled markers refer to the dimensionless
The forces F  of Fig. 19 are calculated above the S:W:L: overtopping volumes.
The normalized wave volumes are defined as V  ¼ VR =VV , where VR The strong reduction of the overtopping volumes induced by the 90∘
and VV are the overtopping volumes as obtained for the recurved wall recurved wall with respect to the vertical parapet is clearly revealed in
and the vertical wall respectively. Fig. 19. This significant decrease of the overtopping volume (filled red
The single overtopping volume has been calculated considering the circle) is accompanied by an increase of the maximum force (empty blue

11
M. Castellino et al. Coastal Engineering 136 (2018) 1–15

Fig. 15. Time evolution of the free surface elevation


η for the W5 wave condition. Values of the free sur-
face elevation are reported on the z-axis. Non-
breaking wave condition is assured. t2 is the time of
the peak force.

Fig. 16. Time series of the pressure on four probes


along the wall for the W4 wave condition.

circles). The reduction to 70∘ of the circumference sector still maintains a Fig. 19 shows that the efficiency of the curved parapets in reducing
good performance in terms of the wave overtopping reduction, but with a wave overtopping volumes occurred in the range 0:81  Rc =H  1:62
lower force ratio (see b - filled and empty squares). With the 45∘ sector of and that for Rc =H < 0:81 the overtopping reduction becomes negligible.
circumference the overtopping reduction is quite limited, while almost The force ratios, shown in Fig. 19 for Rc =H  1:3 confirm a weak
no force increase is produced by the recurved wall (see c - filled and dependence of the results on the incoming non-breaking wave height and
empty diamonds). period.

12
M. Castellino et al. Coastal Engineering 136 (2018) 1–15

Fig. 17. Diagrams of the pressure and velocity fields at various time intervals for W4 the wave condition. The times considered are: t1 ¼ 20:9s, t2 ¼ 21:0s
and t3 ¼ 21:37s.

The observed force decrease, for decreasing angle of the recurved


wall is similar to the results obtained by Stagonas et al. (2014) for
breaking wave conditions.

6. Concluding remarks

Deep water breakwaters with a recurved parapet wall are often used
to reduce the wave overtopping discharge and therefore to protect and
safely use the port side. In this paper, we compared the pressures and
forces acting on a typical recurved wall versus those acting on a tradi-
tional pure vertical wall of the same height.
Aiming to understand the importance of the geometrical character-
istics on the hydrodynamics, numerical simulations have been carried
out using the IHFOAM model, based on the OpenFOAM®. To examine the
essence of the hydrodynamic processes only regular non-breaking waves,
and flat bottom without any berm at the toe of the structure, have been
considered. Although regular waves have been used, it is interesting to
note that the simulated regular wave conditions could represent (ac-
cording to the Rayleigh distribution) the highest waves for irregular
wave conditions with a wave steepness of roughly sop  0:5s, where sop is
the wave steepness calculated as the ratio of the significant wave height
and the peak wave length in deep water and s represents the wave
steepness reported in Table 1. Therefore the tests represent conditions
with a steepness up to sop  0:03:
The numerical simulations have shown that a fully recurved wall (90∘
angle) can induce a significantly high impulsive pressure peak. Fig. 18. Recurved parapet wall: comparison between numerical pressures
This pressure peak is due to the wave evolution at the wall. We diagrams along the recurved (red line) and vertical (black line) wall. The
identified the phenomenological genesis of the impulsive pressures and envelopes of the pressure peaks are reported on the entire structure for the W5
forces on the recurved wall. wave condition. (For interpretation of the references to colour in this figure
legend, the reader is referred to the Web version of this article.)
When the wave crest reaches the top of the recurved part, a high-

13
M. Castellino et al. Coastal Engineering 136 (2018) 1–15

speed flow runs upward to exit seaward of the recurved part. The
impulsive pressure is generated by the flow confinement induced by the
surging wave crest. This impulsive phenomenon, which we refer to as
“confined-crest impact”, has nothing to do with breaking wave condi-
tions, neither can be fully explained by the flow deflection only. Rather, it
is only when the surging wave crest “blocks” its own running flow,
creating a sudden stop of the horizontal momentum of a significant water
mass, that an instantaneous pressure shock wave occurs, and high
impulsive pressures and forces are obtained. A large part of the vertical
wall is affected by the impulsive pressure increase. The maximum values
of the impulsive pressure are localized under the recurved parapet. The
force increase on the entire structure may be significant with respect to
the pure vertical wall case.
When the confined-crest impact occurs the normalized force ratio for
non-breaking waves is found to be dependent only on the recurved wall
angle α (i.e. it is not a function of wave height and period) and is not
linearly dependent of the horizontal extension of the recurved Br . As
previously mentioned, this is because the force increase is, for a given
bathymetry, mainly a function of the recurved wall geometry (angle and
radius).
Numerical simulations have shown that the recurved shape of the
wall excludes the possibility that trapped air occur in the phenomenon.
Therefore, water may be assumed incompressible. In real sea states it is
not excluded that entrained air may be present in the fluid, inducing
water compressibility not treated by the present numerical application.
However, when non-breaking wave conditions are considered, the water
compressibility is unlikely to play a significant role. A significant amount
of entrained air may be induced in deep water conditions in the presence
of perforated caissons made to absorb the incoming waves (as shown in
Fig. 19. Upper panel: geometrical configurations of the analyzed parapets:
Fig. 3).
α ¼ 90∘ , α ¼ 70∘ and α ¼ 45∘ . Lower panel: normalized overtopping volumes Further research is necessary to provide useful comprehensive design
V  (right axis) and normalized forces F  (left axis) for z 2 [ 0.0; þ6.5 ] m guidelines both in terms of wave forces and overtopping. Ongoing
above the S:W:L as a function of the dimensionless freeboard Rc =H. Circles, research is focussed on: (i) validation of the numerical code for impulsive
squares and diamonds refer to 90∘ , 70∘ and 45∘ respectively. impact; (ii) the influence of different recurved wall geometries not
limited to a sector of circumference; (iii) the presence of a toe berm

Table 4
α ¼ 70∘ . Characteristics and main results of numerical simulations for z2[0.0;þ6.5]m and z2[20.0;þ6.5]m. First column: name of the simulation. Second to sixth columns: maximum total
forces FRmax , maximum horizontal forces FRhmax , maximum vertical forces FRv max calculated for the recurved wall, maximum force FVmax calculated at the vertical parapet wall, forces ratio
F  ¼ FRmax =FVmax for z2[0.0;þ6.5]m. Seventh to eleventh columns: maximum total forces FRmax , maximum horizontal forces FRhmax , maximum vertical forces FRv max calculated for the recurved
wall, maximum force FVmax calculated for the vertical wall, forces ratio F  ¼ FRmax =FVmax for z2[20.0;þ6.5]m.

Wave z2[0.0;þ6.5]m z2[20.0;þ6.5]m


name
FRmax FRhmax FRv max FVmax F FRmax FRhmax FRv max FVmax F

(KN/m) (KN/m) (KN/m) (KN/m) () (KN/m) (KN/m) (KN/m) (KN/m) ()

W4 109.30 109.22 4.28 98.92 1.10 417.50 417.48 4.28 415.10 1.01
W5 225.80 223.75 30.39 127.40 1.77 813.50 812.93 30.39 546.60 1.49
W6 295.60 292.11 45.31 169.20 1.75 922.60 921.49 45.31 705.70 1.31
W7 430.10 425.74 61.12 252.20 1.71 1497.20 1495.95 61.12 1228.00 1.22

Table 5
α ¼ 45∘ . Characteristics and main results of numerical simulations for z2[0.0;þ6.5]m and z2[20.0;þ6.5]m. First column: name of the simulation. Second to sixth columns: maximum total
forces FRmax , maximum horizontal forces FRhmax , maximum vertical forces FRv max calculated for the recurved wall, maximum force FVmax calculated at the vertical parapet wall, forces ratio
F  ¼ FRmax =FVmax for z2[0.0;þ6.5]m. Seventh to eleventh columns: maximum total forces FRmax , maximum horizontal forces FRhmax , maximum vertical forces FRv max calculated for the recurved
wall, maximum force FVmax calculated for the vertical wall, forces ratio F  ¼ FRmax =FVmax for z2[20.0;þ6.5]m.

Wave z2[0.0;þ6.5]m z2[20.0;þ6.5]m


name
FRmax FRhmax FRv max FVmax F FRmax FRhmax FRv max FVmax F

(KN/m) (KN/m) (KN/m) (KN/m) () (KN/m) (KN/m) (KN/m) (KN/m) ()

W4 100.30 100.29 1.01 98.90 1.01 416.22 416.21 1.01 415.10 1.00
W5 143.40 143.27 6.03 127.42 1.13 595.80 595.77 6.03 546.60 1.09
W6 182.20 181.89 10.67 169.20 1.08 755.30 755.22 10.67 705.70 1.07
W7 273.30 272.89 14.91 252.20 1.08 1289.41 1289.32 14.91 1228.00 1.05

14
M. Castellino et al. Coastal Engineering 136 (2018) 1–15

(composite breakwaters); (iv) the influence of irregular waves; (v) Hattori, M., Arami, A., Yui, T., 1994. Wave impact pressure on vertical walls under
breaking waves of various types. Coast. Eng. 22 (1–2), 79–114.
comparison with dedicated hydraulic model.
Higuera, P., Lara, J.L., Losada, I.J., 2013a. Realistic wave generation and active wave
absorption for Navier–Stokes models: application to openfoam®. Coast. Eng. 71,
Acknowledgements 102–118.
Higuera, P., Lara, J.L., Losada, I.J., 2013b. Simulating coastal engineering processes with
openfoam®. Coast. Eng. 71, 119–134.
The authors wish to thank Paolo Contini (MODIMAR s.r.l) for the Higuera, P., Lara, J.L., Losada, I.J., 2014a. Three-dimensional interaction of waves and
fruitful discussions on the engineering problem and the reviewers who porous coastal structures using openfoam®. part i: formulation and validation. Coast.
provided a great contribution to improve the paper. Eng. 83, 243–258.
Higuera, P., Lara, J.L., Losada, I.J., 2014b. Three-dimensional interaction of waves and
porous coastal structures using openfoam®. part ii: Application. Coast. Eng. 83,
References 259–270.
Isaacson, M., 1991. Measurement of regular wave reflection. J. Waterw. Port Coast. Ocean
Anand, K., Sundar, V., Sannasiraj, S., 2010. Dynamic pressures on curved front seawall Eng. 117 (6), 553–569.
models under random waves. J. Hydrodyn. Ser. B 22 (5), 538–544. Kisacik, D., Troch, P., Bogaert, P.V., 2012a. Description of loading conditions due to
Bagnold, R., 1939. Interim report on wave-pressure research. Excerpt from J. Inst. Civil violent wave impacts on a vertical structure with an overhanging horizontal
Eng. 12 (1939), 202–226. cantilever slab. Coast. Eng. 60, 201–226.
Berberovic, E., van Hinsberg, N.P., Jakirlic, S., Roisman, I.V., Tropea, C., 2009. Drop Kisacik, D., Troch, P., Bogaert, P.V., Caspeele, R., 2014. Investigation of uplift impact
impact onto a liquid layer of finite thickness: dynamics of the cavity evolution. Phys. forces on a vertical wall with an overhanging horizontal cantilever slab. Coast. Eng.
Rev. 79 (3), 036306. 90, 12–22.
Blenkinsopp, C., Chaplin, J., 2011. Void fraction measurements and scale effects in Kisacik, D., Troch, P., Van Bogaert, P., 2012b. Experimental study of violent wave impact
breaking waves in freshwater and seawater. Coast Eng. 58 (5), 417–428. on a vertical structure with an overhanging horizontal cantilever slab. Ocean. Eng.
Bredmose, H., Bullock, G., Hogg, A., 2015. Violent breaking wave impacts. part 3. effects 49, 1–15.
of scale and aeration. J. Fluid Mech. 765, 82–113. Korobkin, A., Pukhnachov, V., 1988. Initial stage of water impact. Annu. Rev. Fluid Mech.
Bredmose, H., Peregrine, D., Bullock, G., 2009. Violent breaking wave impacts. part 2: 20 (1), 159–185.
modelling the effect of air. J. Fluid Mech. 641, 389–430. Kortenhaus, A., Haupt, R., Oumeraci, H., 2002. Design aspects of vertical walls with steep
Bullock, G., Obhrai, C., Peregrine, D., Bredmose, H., 2007. Violent breaking wave foreland slopes. In: Breakwaters, Coastal Structures and Coastlines: Proceedings of
impacts. part 1: results from large-scale regular wave tests on vertical and sloping the International Conference Organized by the Institution of Civil Engineers and Held
walls. Coast. Eng. 54 (8), 602–617. in London, UK on 26-28 September 2001. Thomas Telford, p. 221.
Chen, X., 2011. Hydrodynamic Loads on Buildings Caused by Overtopping Waves. Ph.D. Kortenhaus, A., Pearson, J., Bruce, T., Allsop, N., Van der Meer, J., 2003. Influence of
thesis. Delft University of Technology, TU Delft. parapets and recurves on wave overtopping and wave loading of complex vertical
Chen, X., Hofland, B., Altomare, C., Suzuki, T., Uijttewaal, W., 2015. Forces on a vertical walls. Proc. Coast. Struct. 2003, 369–381.
wall on a dike crest due to overtopping flow. Coast. Eng. 95, 94–104. Martinelli, L., Lamberti, A., 2011. Dynamic response of caisson breakwaters: suggestions
Cooker, M., Peregrine, D., 1990. Computations of violent motion due to waves breaking for the equivalent static analysis of a single caisson in the array. Coast Eng. J. 53 (01),
against a wall. In: Proceedings of 22nd International Conference on Coastal 1–20.
Engineering. ASCE, pp. 164–176. Mei, C.C., Stiassnie, M., Yue, D.K.-P., 2005. Theory and Applications of Ocean Surface
Cooker, M., Peregrine, D., 1992. Violent motion as near breaking waves meet a vertical Waves: Nonlinear Aspects, vol. 23. World scientific.
wall. In: Breaking Waves. Springer, pp. 291–297. Murakami, K., Irie, I., Kamikubo, Y., 1997. Experiments on a non-wave overtopping type
Cornett, A., Li, Y., Budvietas, A., 1999. Wave overtopping at chamfered and overhanging seawall. In: Coastal Engineering 1996, pp. 1840–1851.
vertical structures. In: Proc. International Workshop on Natural Disasters by Storm Oumeraci, H., Kortenhaus, A., Allsop, W., de Groot, M., Crouch, R., Vrijling, H.,
Waves and Their Reproduction in Experimental Basins, Kyoto, Japan. Voortman, H., 2001. Probabilistic Design Tools for Vertical Breakwaters. CRC Press.
Cuomo, G., Allsop, W., Bruce, T., Pearson, J., 2010a. Breaking wave loads at vertical Owen, M., Steele, A., 1993. Effectiveness of Recurved Wave Return Walls. Technical
seawalls and breakwaters. Coast. Eng. 57 (4), 424–439. Report. HR Wallingford.
Cuomo, G., Allsop, W., Takahashi, S., 2010b. Scaling wave impact pressures on vertical Pearson, J., Bruce, T., Allsop, W., Kortenhaus, A., van der Meer, J., 2004. Effectiveness of
walls. Coast. Eng. 57 (6), 604–609. recurve walls in reducing wave overtopping on seawalls and breakwaters. In: Coastal
De Gerloni, M., Franco, L., Noli, A., Rossi, U., 1989. Industrial port of porto torres: model Engineering Conference, vol. 29. ASCE American Society of Civil Engineers, p. 4404.
test on the new west caisson breakwater (in italian). In: II AIOM Congress, Napoli, Peregrine, D., 2003. Water-wave impact on walls. Annu. Rev. Fluid Mech. 35 (1), 23–43.
Italy. Plumerault, L.-R., Astruc, D., Maron, P., 2012. The influence of air on the impact of a
De Girolamo, P., Noli, A., Spina, D., 1996. Field measurements of loads acting on smooth plunging breaking wave on a vertical wall using a multifluid model. Coast. Eng. 62,
and perforated vertical walls. In: Advances in Coastal Structures and Breakwaters: 62–74.
Proceedings of the International Conference Organized by the Institution of Civil Ramkema, C., 1978. A model law for wave impacts on coastal structures. In: Coastal
Engineers and Held in London on 27-29 April 1995, vol. 7. Thomas Telford, p. 64. Engineering 1978, pp. 2308–2327.
Di Risio, M., Beltrami, G., De Girolamo, P., 2007. Laboratory investigation on wave Ravindar, R., Schimmels, S., Sriram, V., Stagonas, D., Murthy, M.R., 2016. Spatial
overtopping of composite breakwaters: the port of civitavecchia case. In: distribution of impact pressure on a parapet using tactile sensors. In: PIANC
International Conference on Coastal Engineering, pp. 4616–4627. COPEDEC 2016, At Rio de Jeneiro, Brazil.
Sainflou, G., 1928. Essai sur les digues maritimes verticales. Tech. rep.. Ecole nationale
Di Risio, M., Beltrami, G.-M., De Girolamo, P., Noli, A., Ievolella, G., 2009. Investigation
on overtopping and reflection performance of perforated caisson breakwaters: the des Ponts et Chaussees.
ponza harbour case. In: Coastal Structures 2007: (In 2 Volumes). World Scientific, Scott, J.C., 1975a. The preparation of water for surface-clean fluid mechanics. J. Fluid
pp. 1562–1573. Mech. 69 (02), 339–351.
Elsafti, H., Oumeraci, H., 2017. Analysis and classification of stepwise failure of Scott, J.C., 1975b. The role of salt in whitecap persistence. In: Deep Sea Research and
monolithic breakwaters. Coast. Eng. 121, 221–239. Oceanographic Abstracts, vol. 22. Elsevier, pp. 653–657.
EurOtop, M., 2016. Manual on Wave Overtopping of Sea Defences and Related Structures. Stagonas, D., Lara, J.L., Losada, I.J., Higuera, P., Jaime, F.F., Muller, G., 2014. Large scale
van der Meer, JW and Allsop, NWH and Bruce, T and De Rouck J, J and Kortenhaus, measurements of wave loads and mapping of impact pressure distribution at the
A and Pullen, T and Schüttrumpf, H and Troch, P and Zanuttigh, B. www. underside of wave recurves. In: Proceedings of the HYDRALAB IV Joint User Meeting.
overtopping-manual.com. Takahashi, S., 1996. Design of vertical breakwaters. breakwater design. In: International
Franco, L., 1996. History of coastal engineering in Italy. In: Kraus, N.C. (Ed.), History and Conference on Coastal Engineering, Orlando. Short Course.
Heritage of Coastal Engineering. ASCE, New York, pp. 275–335. Van Doorslaer, K., Romano, A., De Rouck, J., Kortenhaus, A., 2017. Impacts on a storm
Franco, L., Arena, F., Mazzola, O., Petrosemolo, L., 2013. The new deepwater marina di wall caused by non-breaking waves overtopping a smooth dike slope. Coast. Eng.
cicerone at formia (i). In: ICE Conference Coasts, Marine Structures and Breakwaters 120, 93–111.
2013. Thomas Telford. Von Karman, T., 1929. The Impact on Seaplane Floats during Landing.
Franco, L., De Gerloni, M., Van der Meer, J., 1995. Wave overtopping on vertical and Wagner, H., 1932. Über stoß-und gleitvorg€ange an der oberfl€ache von flüssigkeiten.
composite breakwaters. In: Coastal Engineering 1994, pp. 1030–1045. ZAMM-J. Appl. Math. Mech./Z. Angew. Math. Mech. 12 (4), 193–215.
Gibson, F.W., 1970. Measurement of the effect of air bubbles on the speed of sound in Wood, D.J., Peregrine, D.H., 1999. Two and three-dimensional pressure-impulse models
water. J. Acoust. Soc. Am. 48 (5B), 1195–1197. of wave impact on structures. In: Coastal Engineering 1998, pp. 1502–1515.
Goda, Y., 2010. Random Seas and Design of Maritime Structures. World scientific. Wood, D.J., Peregrine, D.H., Bruce, T., 2000. Wave impact on a wall using pressure-
impulse theory. i: trapped air. J. Waterw. Port Coast. Ocean Eng. 126 (4), 182–190.

15

You might also like