Sie sind auf Seite 1von 13

Ocean Engineering 199 (2020) 107028

Contents lists available at ScienceDirect

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

Tsunami intrusion through port breakwaters enclosed with


self-elevating seawalls
Hiroshi Takagi a, *, Ryoichi Tomiyasu b, Tomoyuki Oyake c, Taketo Araki b, Kyosuke Mori b,
Yasuhiro Matsubara b, Yohei Ninomiya c, Yoshifumi Takata a
a
School of Environment and Society, Tokyo Institute of Technology, Tokyo, Japan
b
Kyodo Engineering Corporation Ltd., Fukuoka, Japan
c
Oriental Shiraishi Corporation Ltd., Tokyo, Japan

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

Keywords: An innovative self-elevating seawall is attracting attention as an effective countermeasure to protect ports and
Self-elevating seawall their hinterland from tsunamis. However, seawater may intrude into the port basin through a narrow gap be­
Tsunami intrusion tween its gate units, resulting in a significant water level change when a tsunami strikes. Herein, a 2D–3D hybrid
Gate gap
hydrodynamic model is used to investigate water inflow through the gap in a self-elevating seawall, induced by a
Practical formula
tsunami that approaches a port. A practical hydraulic head-to-discharge formula is then derived. This formula is
Case study
Breakwater rubble mound validated by a case analysis that numerically simulates the intrusion with a hypothetical time series of water
Level 1 tsunami levels originating from a tsunami and astronomical tide. A self-elevating seawall acts as an effective low-pass
Tide filter system, though it may not completely prevent inflow. A case study for assessing a medium sized fishery
Storm surge port also demonstrates that tsunami-induced impact can be significantly reduced by sheltering with the self-
Gate-opening-induced tsunami elevating seawall system. However, a longer wave component, such as astronomical tide, will inevitably
intrude into the port through the gate gap as well as the void of breakwater’s rubble mound as seepage flow. The
risk of a secondary tsunami, gate-opening-induced tsunami, is also addressed for appropriate gate operations after a
tsunami has receded. The proposed model can calculate rise/fall in water level over a port basin enclosed by the
self-elevating seawall and port breakwaters using a simple spreadsheet without expensive computational fluid
dynamics.

1. Introduction destroyed during the tsunami in 2011, allowing significant tsunami


intrusion into Kamaishi city. The hydrodynamic scour due to the over­
The Great East Japan Earthquake and resulting tsunami that topping jet plunging down the landward side of the caisson and the
occurred in 2011 revealed weaknesses in the flood-prevention structures punching failure of the rubble mound foundation were considered two
that are currently in place. Tsunami breakwaters have been constructed fundamental causes of breakwater failure (Arikawa et al., 2012; Bricker
in several bay mouths in northern Japan over the last few decades. These et al., 2013; Esteban et al., 2017). Immediately after the disaster, a
breakwaters were exclusively designed to protect bays, ports, and governmental research institute publicly estimated that the tsunami
nearby communities from tsunamis. When designing these structures, height was probably reduced from 13.7 to 8.0 m, and that the tsunami
seismic and high-wave forces, as well as tsunami forces, must be breakwater allowed residents an extra 6 min to evacuate. However, this
considered, resulting in breakwaters with massive dimensions and estimation may not be completely reliable as it neglected the influence
expensive costs. For example, the Ofunato bay-mouth breakwater, the of several breakwater sections that were toppled during the tsunami
oldest tsunami breakwater in Japan, was constructed immediately after (Cyranoski, 2012). Therefore, much more research is needed to validate
the Chilean tsunami in 1960. However, this breakwater was almost this (Shibayama et al., 2013). In addition to uncertainties related to the
completely destroyed by the tsunami in 2011. The Kamaishi tsunami effectiveness and durability of the structure, the enormous construction
breakwater was the largest in the world and was built in a maximum cost (e.g., 124 billion JPY (¼ approx. 1.1 billion $)) of the Kamaishi
water depth of 63 m (Raby et al., 2015). However, this was also largely tsunami breakwater will make it difficult to disseminate such

* Corresponding author.
E-mail address: takagi@ide.titech.ac.jp (H. Takagi).

https://doi.org/10.1016/j.oceaneng.2020.107028
Received 9 September 2019; Received in revised form 28 January 2020; Accepted 28 January 2020
Available online 9 February 2020
0029-8018/© 2020 Elsevier Ltd. All rights reserved.
H. Takagi et al. Ocean Engineering 199 (2020) 107028

technologies across the country and the world. which consists of four separate walls, is 32 m wide and 12 m high and is
On the other hand, general breakwaters are attracting attention as placed at a water depth of 4.5 m. The estimated construction cost may
effective countermeasures to protect ports and their hinterland from reach 2.5 billion JPY (22 million $). More recently, the Macau SAR
tsunamis as these appeared to reduce fatality ratios in most coastal areas Government announced a plan for protecting the Inner Harbor Area­
during the 2011 tsunami (Latcharote et al., 2016). Extensive investiga­ —one of the areas most impacted—from storm surge. This plan involves
tion of the 1983 Japan Sea Tsunami also indicated that upright re­ a 650-m long huge movable floodgate system at the mouth of the Wan
vetments and breakwaters in many ports probably reduced tsunami Chai waterway (Moura, 2018; Takagi et al., 2019a).
run-up heights (Tanimoto et al., 1983). However, the effectiveness of Although technologies are still being investigated, many researchers
general breakwaters, which were designed to reflect wind waves, in have studied other types of self-elevating tsunami gates over the last
reducing tsunami impact should not be overestimated (Takagi and decade. For example, Kimura et al. (2010) proposed the flap-gate
Bricker, 2014; Takagi, 2015). When comparing the region protected by breakwater, which normally lies down on the seabed and rises as a sea
breakwaters with an unsheltered area in Ishinomaki, a tsunami simu­ wall driven by buoyancy when a tsunami strikes. This type of structure
lation model showed no noticeable differences in inundation due to the appears similar to the gate of the MOSE project but employs a different
2011 tsunami, because the tsunami intrudes through the port’s wide supporting method. Nakashima et al. (2011) investigated the perfor­
openings (a few hundred meters) and rapidly fills the basin with mance of a vertical telescopic breakwater, which consists of movable
seawater. However, a hypothetical simulation revealed that water levels upper and fixed lower steel pipes. Hofland et al. (2015) proposed a novel
could be greatly reduced if the port were fully enclosed by a tsunami barrier consisting of a membrane, floater, and cables, which are
self-elevating seawall before a tsunami struck (Takagi and Bricker, normally stored underground and are automatically floated by buoy­
2014). ancy. A Japanese Research Group for Self-elevating Seawall (RGSS)
Storm-surge barriers have been studied, developed, and constructed (2016) proposed a new type of structure, which is stored under a
over the last few decades in several countries, such as the United States, pneumatic caisson system and is elevated when a tsunami strikes
UK, Netherlands, Italy, France, and Japan (PIANC, 2006). Dircke et al. (Fig. 2). The pneumatic caisson method is utilized in this technology,
(2012) present a systematic overview, comparison, and selection of six which allows the construction of an underground structure and simul­
navigable storm surge barriers: miter gate, vertical lifting gate, flap gate, taneous excavation of the soil, based on compressed-air geotechnical
horizontal rotating gate, vertical rotating gate, and inflatable rubber engineering (Abe et al., 2017).
dam. The flap gate appears to be one of the most common self-elevating A common floodgate, such as that shown in Fig. 1(c), must be sup­
structures. This consists of a straight or curved retaining surface, which ported by massive columns to resist tsunami as well as seismic forces,
is pivoted on a fixed axis at the sill (Erbisti, 2004). Among several others, resulting in massive dimensions and costs. On the other hand, the self-
the gates being constructed in the MOSE project (https://www.moseve elevating seawall, which is normally stored under the sea, has many
nezia.eu/), which aim to protect the city of Venice, may be the most advantages over traditional floodgates. Particularly, it is highly benefi­
famous structure of this type in the world (Fig. 1 (a)). Iwate Prefecture cial that ships can enter the port without height restrictions under
was the most damaged by the tsunamis in 2011 as these had heights normal operation. Also, these structures will not mar the scenery, unlike
exceeding 20 m (Mori et al., 2012). With much support from the local vertical lifting gates supported by huge columns and beams or an
inhabitants, a port flap gate has been approved and its construction has extension of the existing breakwater.
started at a fishing port in Ofunato Bay in Iwate (Fig. 1 (b)). Although As shown in Fig. 3, a 2-m tsunami can be perfectly hold back by
this has not been well recognized internationally, it will become the first implementing the self-elevating seawall if it can completely enclose the
fully operational flap gate against tsunamis in the world. The gate, opening of a medium sized fishery port (RGSS, 2016). However, there

Fig. 1. Coastal flood gates (Photos were all taken by the authors).

2
H. Takagi et al. Ocean Engineering 199 (2020) 107028

Fig. 2. Schematic drawings of a self-elevating seawall with a pneumatic caisson. The steel shell wall is stored inside the caisson under normal sea conditions. When
tsunamis or storm surges strike, the wall is elevated to close the port by remotely operating a wire rope and pulley system operating the inner wall (Adapted from
RGSS, 2016).

Fig. 3. Effectiveness of the self-elevating seawall against a 2-m tsunami, which is expected to arrive after an hour of a scenario earthquake. However, this simulation
was performed under an ideal condition without considering the gate gaps (Adapted from RGSS, 2016).

appear to be weaknesses in the self-elevating seawall structure. Most resulting flow pattern is not readily subject to any analytical solution
importantly, a narrow gap between gate units would allow intrusion of (Chow, 1959).
seawater into the port basin, which would probably cause significant The mechanism of tsunami intrusion through the gaps cannot be
water level changes when a tsunami struck. While being sufficiently easily evaluated using a commonly-used tsunami numerical model
small, a gap must exist between adjacent gate units because of me­ because the gap is too narrow to be captured in the simulation. A
chanical systems that ensure smooth elevation of the gates even after a tsunami simulation protocol with a shallow-water wave model suggests
major earthquake (Fig. 2). that the grid size should be reduced at a ratio of one third (UNESCO,
The gate shown in Fig. 1(c), classified as a vertical lifting gate, will 1997). For example, if the smallest grid is connected to an 810 m grid,
not allow seawater intrusion when closed as the edges of the gate can be which is a typical grid size over a large computational domain covering
almost perfectly sealed with the concrete columns. On the other hand, an ocean basin containing a seismic rupture area, the nesting procedure
self-elevating seawalls, such as those proposed by RGSS (2016) and Fujii must be repeated up to 10 times to obtain a 1 cm grid that is sufficient to
et al. (2018) (Figs. 2 and 4), flap gates in Venice (Fig. 1(a)), and the capture the narrow gap between individual gates. While theoretically
rotating gate in the Netherlands (Fig. 1 (d)) inevitably have slight gate possible, such multiple connections appear to be computationally too
gaps of at least a few tens of centimeters. Nakashima et al. (2011) sug­ expensive.
gested that the gap width in the floating breakwater system should be A tsunami may also intrude through the void of rubble mound
less than 3% of the total stretch of the port breakwater. However, no beneath an existing breakwater (Fig. 5). A hybrid structure known as
technical procedure for estimating the tsunami inflow was presented. caisson-type breakwater is composed of caisson concrete and rubble
Flow through such a constriction is usually complicated that the mound (Tanimoto and Goda, 1991). The caisson structure is

3
H. Takagi et al. Ocean Engineering 199 (2020) 107028

Fig. 4. Construction procedure of a self-elevating seawall proposed by Fujii et al. (2018).

Fig. 5. Schematic diagram of the seepage flow induced by a seawater intruding through the rubble mound.

impermeable, whereas the rubble mound inevitably undergoes seepage head difference between the sea side and the port side of the breakwater.
flow when difference in water level appears before and after the The value of β is recommended to use 0.0354 to best estimate the rate of
breakwater. The rubble mound is submerged under the sea and gener­ seepage flow through the rubble mound composed of stones with the
ally as thin as 20–40% of the height of the caisson (Takagi and Esteban, mean diameter of 50 cm.
2013). A tsunami inflow due to the overtopping of a breakwater has Rapid varied flows through the contractions in hydraulic structures
been extensively investigated over the last decade (Esteban et al., 2015). have been extensively studied by hydraulic engineers. For example, the
However, the seepage flow during the rising phase of a tsunami appears discharge coefficients for a vertical embankment with an opening can be
to be invisible, and thus this phenomenon tends to be overlooked. The readily derived from the diagrams provided by Chow (1959). As these
assessment of tsunami intrusion through the rubble mound is particu­ studies mostly considered steady flow in an open channel, the continuity
larly important when a vertical lifting gate is installed in the port equation is valid between the upstream and downstream boundaries,
opening surrounded by long-stretched breakwaters. Even if the lifting thereby simplifying the required calculations. However, to the best of
gate perfectly prevents from a tsunami entering, seawater through the the authors’ knowledge, no reliable and practical methods have been
mound could raise the water level inside the port and might damage proposed for estimating the intrusion of seawater, caused by tsunamis or
ships, facilities, and buildings. other types of long waves (e.g., storm surges, seiches), through gaps in
In light of the seepage flow, Takata and Takagi (2019) recently seawalls. Unlike in a river stream, the downstream discharge is not an a
proposed the following practical formula to estimate the inflow priori condition when a tsunami approaches, representing an unsteady
discharge rate through a breakwater mound by conducting a hydraulic flow. Therefore, the conventional diagrams for open-channel flows may
experiment and numerical analysis using OpenFOAM. Their formula not be applied when assessing the intrusion of seawater through the gate
considered the nonstationary flow (Forchheimer flow) in porous me­ gaps due to a tsunami.
dium using the formula of Ergun (1952), which incorporates size of On the other hand, wind waves passing through cylinders have been
medium and porosity of the layer (usually about 40% for rubble mound). extensively studied by coastal engineers aiming to investigate the per­
pffiffiffiffiffiffiffiffiffiffi formance of special types of breakwaters, such as wave screens and slit-
v ¼ β 2gΔh (1) type breakwaters (e.g., Kakuno and Liu, 1993; Mei et al., 2005). The
most advanced solutions of these systems are based on potential flow
where v is the mean velocity within rubble mound, β is an experimental
methods, in which the velocity potentials are solved for both the linear
coefficient, and g is the gravitational acceleration. Δh is the hydraulic
incident and reflected waves by applying both far-field boundary

4
H. Takagi et al. Ocean Engineering 199 (2020) 107028

Fig. 6. A total of eight tsunami scenarios, with four different heights (1, 2, 3, and 4 m) and two temporal patterns (flat and Gaussian).

conditions and matching conditions at the wall boundary. Theoretical (2011):


and experimental studies have both demonstrated that wave trans­ –Level 1 Tsunamis: These are events with a return period of several
mission decreases as the wave steepness increases and the porosity de­ decades to 150 years. They generate low to medium inundation depths,
creases (Kriebel, 1992). The solutions derived by those potential typically less than several meters. Both human lives and properties
approaches will satisfy the equation K2r þ K2t ¼ 1, where Kr and Kt are should be essentially saved by hard countermeasures.
the reflection and transmission coefficients, respectively. This implies –Level 2 Tsunamis: These are much rarer events, typically taking
that the solution lacks information pertaining to the time lag between place at intervals of a few hundred to a few thousand years. The tsunami
the reflected and transmitted waves. As wind waves are repetitive and inundation depths would be much larger, typically over 10 m, but would
short-period processes, this lag may not be very important. However, encompass inundations reaching 20–30 m. As hard countermeasures
tsunamis cause water-level rises/falls over a much longer duration. cannot work against this level of tsunami, soft countermeasures need to
Hence, the time lag of the water level between the two sides of a gate is be implemented to save people.
more significant. For example, a shorter time lags indicates a prompt Extremely high tsunami waves, such as those categorized as a Level 2
response of the water level inside of the port in response to the rise of the tsunami, may break offshore and propagate in shallow coastal areas in
water level due to tsunamis outside of the port, and consequently the the form of a violent hydraulic bore, which consists of hydrostatic, hy­
seawater will quickly rise in the port. However, this in turn reduces the drodynamic, and surge forces (Nistor et al., 2009). In particular, the
gradient of the sea surface between the outer sea and the port basin, surge force, which is generated by the impingement of the advancing
resulting in a reduction of the inflow discharge. This expected feedback water front of a tsunami bore on a structure, causes huge forces.
mechanism between water levels inside and outside of the port cannot Although accurately estimating the impact force remains a challenging
be readily accounted for by the potential theory used for wind waves. task, the City and County of Honolulu Building Code (CCH) has rec­
Given the lack of practical methods, the present paper attempts to ommended the following formula to estimate this force based on a report
derive a method for calculating the inflow discharge caused by a by Dames and Moore (1980):
tsunami whereby engineers can quickly estimate the water levels in an
Fs ¼ 4:5ρgh2 (2)
enclosed port. One of the greatest motivations of this study is to provide
a methodology that can be performed using a spreadsheet and involves where Fs is the surge force per unit width of the wall and h is the surge
no expensive numerical modeling processes. By simply using the derived height. This force is equivalent to 9 times the hydrostatic force. Typical
formula in a spreadsheet, a case study is then conducted to examine coastal structures, buildings, and self-elevating seawalls may not with­
whether or not a self-elevating wall is effective even under unfavorable stand such enormous forces.
conditions considering inflows through the gate gap and the rubble Given these emerging consensuses among Japanese engineers, the
mound of a breakwater. present study considers only moderately high tsunami waves, such as
those categorized as a Level 1 tsunami. Against this level of tsunami,
2. Methodology protection measures must protect human lives and properties and
effectively facilitate the evacuation of local inhabitants to safer places
This section describes tsunami waveforms assumed in this study, (Esteban et al., 2017). In this study, it was hypothesized that the
followed by explanation on numerical models for the estimation of self-elevating seawall would not be breached by Level 1 tsunamis,
tsunami intrusions through the gap of a gate. As a result, a hydraulic though water levels would rise in front of the gate. The range of tsunami
head-to-discharge formula is derived to provide an assessment tool for heights was set at 1–4 m because an ordinary flood protection system
tsunami intrusion. can protect against this level of tsunami. On the other hand, a tsunami
exceeding this range may overtop the existing port breakwater even if
2.1. Model tsunamis the port opening is closed by a gate, nullifying the entire system.
The present study assumes that the long waves of a tsunami are
Since the Great East Japan earthquake and tsunami in 2011, the idea represented by the flat pattern of water elevations shown. The Gaussian-
that hard measures can completely protect the lives of inhabitants has type tsunami pattern was also examined to investigate whether flow
been abandoned (Shibayama et al., 2013). In order to classify according characteristics are influenced by a rapid increase in water surface
to the frequency/magnitude of tsunami, two levels were introduced by elevation in front of the gate. A time series of water levels could be
the Central Disaster Management Council of the Cabinet Office of Japan

5
H. Takagi et al. Ocean Engineering 199 (2020) 107028

Fig. 7. Water levels observed at the three locations (Sendai, Mera, and Chiba) on March 11, 2011, when the Great East Japan Earthquake occurred. Sendai and Mera
both face the Pacific Ocean, while Chiba is located in the inner-most part of Tokyo Bay (data obtained from the NOWPHAS wave observation station at the Sendai
Port and the JMA tidal stations at Mera and Chiba).

represented by the following Gaussian distribution, which can repro­ behave either as leading depression waves or leading elevation waves,
duce a rapid rise and subsequent fall in water level of η. rather than as steep solitary waves (Madsen et al., 2008). Although
� � tsunami wave forms may appear to be solitary waves, the actual wave
η ¼ ​ η0 exp ðt rÞ2 r (3) periods are generally much longer than those derived using the KdV
equation (typically shorter than 1 min).
where η0 denotes the tsunami height, t is time (minutes), and r is the rise
For example, Fig. 7 shows the water levels observed at three tide
time, which is the time taken to reach the peak water level. A total of
stations before and after the occurrence of the tsunami generated by the
eight water levels were examined in this study (Fig. 6).
Great East Japan Earthquake. Sendai Port is located close to the
The numerical offshore boundary condition will depend on how
epicenter, while Mera and Chiba are both relatively distant from the
quickly the seawater rises as a tsunami approaches. If the tsunami
epicenter. Mera faces the Pacific Ocean, while Chiba is inside of Tokyo
strongly impacts the gate as an impulsive tsunami, a solitary wave or
Bay. The identifiable major signals of the tsunami in these figures are not
dam-break-type wave boundary may be used. However, it is currently
very short, but it appears that each wave continued for at least 20 min. It
generally acknowledged that incident tsunamis approaching coasts
is also interesting to note that higher-frequency waves disappeared in

6
H. Takagi et al. Ocean Engineering 199 (2020) 107028

the inner bay at Chiba, wherein a wave period of approximately 60 min step was set as 0.03 s, which appears to be sufficiently small to stably
became predominant. This is likely because of the shallow bathymetry perform the computation, including the gate gap with the ADI method.
that smoothed the waves propagating into the bay. Incident water levels, as shown in Fig. 6, are imposed from the
Using satellite data, which was obtained during the 2004 Indian offshore boundary of domain #1 with no incident angle. On the other
Ocean Tsunami, Gower (2005) estimated a wavelength at 430 km with a hand, an impermeable and indefinite high wall is assigned as the land­
wave period of 37 min as the tsunami traveled over the deep ocean. ward boundary in domain #3 to inhibit overflow of the port’s apron. The
Given these observations, the long waves shown in Fig. 6, which possess width of the tsunami gate was assumed to be 2 m by referencing to a
periods of 6 min, may be used to represent Level 1 tsunamis at a port site. standard dimension of the self-elevating seawall proposed by RGSS
(2016).
2.2. Model for tsunami intrusion via gate gaps
2.3. Water levels and flow velocities inside and outside the gap
The authors used Delft3D-FLOW (Deltares, 2011), constructed on the
supercomputer TOKYO-TECH TSUBAME (Ver2.5; 17.1 PFLOPS of single The profiles of simulated water levels along the cross section of the
precision performance), to simulate the flow through the narrow gap of gate gap clearly demonstrate that water surface elevation substantially
a self-elevating seawall. The Delft3D-FLOW solves the unsteady dropped behind the gate (Fig. 9). A total of eight incident wave patterns,
shallow-water equations in two or three dimensions. The system of as shown in Fig. 6, were generated in front of the gate. For the flat water-
equations consists of the horizontal momentum equations, the conti­ level pattern, the profiles are those output 3 min after the start of the
nuity equation, the transport equation, and a turbulence model. The simulation, which coincides with the peak time of the Gaussian scenario.
vertical momentum equation is reduced to the hydrostatic pressure All the water levels are normalized by dividing by the incident heights
relationship because the vertical accelerations are assumed to be small (1, 2, 3, or 4 m) measured at 5 m in front of the dyke (shown as “FRONT”
compared to the gravitational acceleration. This feature makes the in Fig. 8) to show the transmission ratio (Tr) during the passage of waves
model particularly suitable for simulating flows that occur in shallow through the gate gap.
seas, coastal areas, and estuaries (Lesser et al., 2004; Deltares, 2011; The water level starts to drop immediately in front of the gate, and
Sasaki et al., 2012). the Tr decreases by about 40% for all the scenarios. The water level
This hydrodynamic model was the most commonly used open-source further drops within the gate gap, but the extent of the drop is highly
software in the world and selected from other equivalent models for two dependent on the initial tsunami height. All the lines re-converge behind
main reasons. the gate and drop rapidly to almost zero. The profiles of flat and
Firstly, it appears that a very small computational grid spacing, and Gaussian patterns are not remarkably different.
time step must be adopted to simulate a tsunami passing through such The Tr decreases with increasing incident height inside the gate gap.
narrow gate gaps, which connect with both the sea and port basin. If the The larger the tsunami height, the faster the flow velocity, as shown in
simulation is performed using an explicit scheme, such as Euler’s Fig. 10. For example, the calculated velocity is about 1.5 m/s for a height
method, a preliminary study revealed that a time step of an order of of 1 m, while it increases to 6 m/s for a height of 4 m. The velocity shown
0.001 s would be required to stably simulate tsunami propagation, as a in Fig. 10 is the depth-averaged velocity, which is calculated by dividing
longer time step would cause numerical instability. This would not be the flow discharge through the gap by the water depth at a given time
economical. On the other hand, the Delft model adopts the ADI (Alter­ and location. It was useful that the Delft3D-FLOW model contains a
nating Direction Implicit) method for time integration, which splits one- module for estimating the depth-averaged velocity. Although the ve­
time step into two stages. Since water velocities and water levels can be locity is relatively small before and after the gate, it increases consid­
implicitly solved along grid lines by this method, the Courant number erably when water passes through the gate gap.
does not need to be below unity. In practical situations, the Courant Although the water level drops in a staircase pattern, there is no
number could be relaxed to a value of 10 (Deltares, 2011), contributing strong undulation and turbulence for a gate gap of 30 cm. This is also
to a realistic computation. Secondly, the model adopts domain decom­ confirmed by Fig. 11, which shows the Froude Number (Fr) in addition
position (DD), which divides the computational domain into several to the Tr. Fr increases as the gap width decreases. Fr exceeds unity when
smaller domains. Communication between the domains occurs along the gate width is 10 cm and 15 cm, demonstrating that a hydraulic jump
internal boundaries, so-called DD-boundaries, when using a parallel occurs near the gate exit where water intrudes into the port basin. Ac­
computing technique. It should be noted that using a fully unstructured cording to Chow (1959), Fr values between 1.7 and 2.5 and between 1
grid would allow as much or even more flexibility than using the DD and 1.7 are classified as weak jumps and undular jumps, respectively. In
method. However, connecting an unstructured grid with a spacing of these states, the velocity is fairly uniform, and energy loss is low.
several meters to those with spacings of a few centimeters by a short Although a series of small rollers develop on the surface of the jump, the
distance appears to be complex, and the resultant model may be both downstream water surface is expected to remain fairly smooth. On the
inaccurate and inefficient. other hand, Fr is below unity for gap widths of 20 and 30 cm; therefore,
Three computational domains, (1) outside the port, (2) the tsunami no remarkable jumps occur. One may consider making the gap as narrow
gate, and (3) inside the port, are sequentially connected as shown in as possible to prevent water inflow. However, the gap width then
Fig. 8 with the DD boundary. The middle domain was constructed with significantly influences the flow regime, implying a trade-off whereby a
the finest grids to reproduce the narrow gap width, assuming a range narrower gap may reduce the stability of hydrodynamic conditions in­
from 10 to 30 cm. It is noted that a 3D grid vertically dividing 20 layers side and behind the gate.
was applied to the middle domain to determine whether vertical flows
are significant in addition to horizontal flows. The k-ε model was 2.4. Hydraulic head-to-discharge formula
adopted for simulating turbulence, while the bottom friction was eval­
uated with a Manning’s n value of 0.025, which is a typical value for In this study, the middle point of the gate, indicated as “MID” in
seabed (Bricker et al., 2015). The other two domains used a 2D hori­ Fig. 8, was selected for estimating the inflow discharge. This is because
zontal grid; thus, the code is equivalent to a shallow-water wave model the flow is relatively stable at this point, even in those cases with a
(Takagi et al., 2016). The grid size in the 2D–3D hybrid model varies in hydraulic jump, as demonstrated by Figs. 9, 10, and 11, enabling precise
the flow direction from a maximum of 50 m inside and outside the port measurement of the water discharge. Interestingly, Figs. 9 and 10
to a minimum of 4 cm within the narrow gap. The total area of the port demonstrate that the influence of the tsunami rise time on the inflow
basin considered in this model is about 75,000 m2 (7.5 ha), which is discharge is unremarkable, and both the velocity and water elevation
typical of medium-sized fishing ports in Japan. The computational time are similar for flat and Gaussian patterns. Nevertheless, Fig. 10 shows

7
H. Takagi et al. Ocean Engineering 199 (2020) 107028

Fig. 8. Model for tsunami intrusion through gate


gaps. The entire computational domain is
composed of three sub-domains (#1: offshore
domain; #2: middle domain; #3: landward
domain). The self-elevating gate and its gap are
modelled within domain #2. The gap widths are
assumed to be W ¼ 10, 15, 20, and 30 cm, while
the cross section is kept constant at 2 m. The points
denoted as FRONT, IN, MID, and OUT indicate the
locations at which the computational results, such
as the water level and velocity, are output.

Fig. 9. Tsunami transmission through the gate gap, reproduced by inputting


Fig. 10. Depth-averaged flow velocities through the gate gap, reproduced by
eight tsunami scenarios. The gray rectangle indicates the horizontal width of
inputting eight tsunami scenarios.
the gate. The water rises in front of the gate, as shown on the left-hand side. The
transmission ratio is defined by the water level η at a given location along the
cross-shore line divided by the incident water level η 5m at 5 m in front of case study, the gate width is as narrow as 2 m). Given that the quanti­
the gate. fication of the energy loss is highly uncertain, this factor that reduces the
calculated inflow discharge should be neglected during design to ensure
that the velocity in the Gaussian case is about 6% smaller than in the flat that a safe design is achieved. The discharge coefficient C, which rep­
one for ηinput values of 2 and 3 m. However, for the safety of the design as resents the total head loss, can be estimated by rearranging Equation (4),
pffiffiffiffiffiffiffiffiffiffi
well as its practical simplicity, the flat pattern is used to investigate i.e., C ¼ Q= 2gΔh ​ w ​ h, using the numerical result for Q, which is
water inflow via the gap in a self-elevating seawall from here onwards. presented in the previous section. The discharge coefficient represents
The flow discharge via the gap in a gate was calculated using the the balance between the momentum of a tsunami and energy loss that
following formula, which is like Torricelli’s theorem: leads to a reduction in discharge through friction, contraction, expan­
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi pffiffiffiffiffiffiffiffiffiffi sion, and eddy. The larger the coefficient, the more predominant the
Q ¼ γC 2gðΔh hl Þ ​ w ​ h � γC 2gΔh ​ w ​ h (4) momentum, which facilitates water flow via the gap. The regression
between the discharge coefficient and hydraulic head was derived for w
where Q is the flow discharge; C is the discharge coefficient (<1); g is the
¼ 10, 15, 20, and 30 cm, as follows:
gravitational acceleration; Δh and hf are the hydraulic head between the
w ¼ 10 cm:
sea side and port side of the gate and the friction loss within the gate gap,
respectively; w is the gap width; and h is the water depth in the gate gap. Δh � 2 : ​ C ¼ 0:24 ​ Δh þ 0:45 (5)
γ is a correction factor, as will be explained later in this paper. The
friction loss, hl , can be calculated using the Darcy-Welsbach formula, Δh > 2 : C ¼ 0:93 (6)
and this value should be small when the gate width is narrow (in this w ¼ 15 cm:

8
H. Takagi et al. Ocean Engineering 199 (2020) 107028

discharge coefficients with Equations (5)-(9) were purely derived from


the numerical analysis. Therefore, the correction factor, γ, aims to
modify the formula by incorporating some of the other energy dissi­
pating mechanisms that cannot be fully reproduced by the present hy­
drodynamic model. For example, the Delft3D FLOW model can
reproduce moderate turbulence using the k-ε model. However, this
model cannot simulate highly turbulent flow, such as breaking waves
and dam-break waves, because the hydrostatic pressure assumption may
limit the representation of fluid dynamics within/around the gate gap.
Nevertheless, as explained earlier, impulsive tsunami impacts are not
considered in the present analysis because the self-elevating seawall is
expected to be designed considering the gradual rise/fall in water levels.
In such an environment, the hydrostatic assumption may be reasonable
as the forces resulting from the tsunami should predominantly act in the
horizontal direction.
The correction factor, γ, may be modified via experiment using a
water flow or wave flume. However, the physical hydraulic test cannot
also be readily carried out because the gate gap (10–30 cm in this study)
Fig. 11. Transmission ratios, Tr (left vertical axis) and Froude numbers, Fr
(right axis) for three different gap widths, W ¼ 10, 15, 20, and 30 cm. A con­
must be the same between the test-scale model and the actual-scale gate
stant water level of η ¼ 4 m is imposed. (i.e. the components in the experiment cannot be scaled down using the
Froude similitude). This condition cannot be reproduced within a
normally-sized flume, which typically possesses a length of 10–20 m and
C¼ 0:04Δh2 þ 0:34Δh þ 0:18 ð � 0:93Þ (7)
a width of 1 m or less. Even if a large flume is available, a very large
w ¼ 20 cm: pump with a high discharge capacity (over 1 m3/s) is required to
maintain the water level difference in front of/behind the gate. Obvi­
C ¼ 0:15 ​ Δh þ 0:29 ð � 0:93Þ (8)
ously, this is not a readily available option. In addition, the pump will
w ¼ 30 cm: also generate unnecessary and strong turbulence. The authors wish to
carry out large-sized flume tests to validate the calculated discharge
C ¼ 0:15 ​ Δh þ 0:25 ð � 0:93Þ (9)
coefficients in the near future, but it is recommended that the correction
Fig. 12 demonstrates some general characteristics of these relation­ factor, γ, is simply assumed to be unity for now.
ships. For example, the discharge coefficient increases with the hy­
draulic head, implying that seawater flows more easily into the port 3. Results
basin with relatively less momentum reduction as the tsunami height
increases. The discharge coefficient for a 10 cm gap is composed of two This section describes the reliability of the proposed formula by
regression lines, while the coefficients with the other gap widths are comparing it with the numerical model. A spectrum analysis is also
drawn in a simple linear manner. This is because a hydraulic jump ap­ performed to investigate if any difference in performance of the self-
pears for a width of 10 cm, and it begins when the hydraulic head ex­ elevating seawall occurs as wave period changes. Finally, a case study
ceeds 2 m. If the flow becomes supercritical (Fr > 1), the fluid motion is is conducted to examine the performance of the self-elevating seawall
determined by the water levels in the downstream part of the port, and under a realistic condition. Some of extra considerations are also
subsequently the flat regression line is derived. Interestingly, the addressed for achieving appropriate gate operations.
regression line for a 15-cm gap can be expressed with a polynomial
curve, exhibiting a transition between the linear regression groups. 3.1. Simplified formula vs numerical model
Equation (4) is based on Torricelli’s theorem. However, the
Water discharge through a gate gap can be easily calculated by
applying the discharge coefficients to the formula of Equation (4). It is
straightforward to investigate changes in water surface variation inside
the port over time by assuming parameters such as the number of gate
gaps, gap width, depth, tsunami scenario, and area of the port basin.
This level of calculation can be achieved even using an ordinary
spreadsheet.
The proposed model was examined based on its ability to a reason­
able prediction equivalent to the numerical simulation. The water level
in the port basin was output using Delft3D to compare the proposed
method with the numerical simulation. Although the same domain as
that shown in Fig. 8 was used for the simulation, a total of five gaps, each
with a 30 cm width, were allocated at intervals of 15 m. Thus, the sum of
the gate gaps accounts for 0.7% (¼1.5/215 m) of the total extent of the
self-elevating seawall. For this trial, a hypothetical test tsunami up to
2.5 m was generated by combining a 2 m high tsunami (r in Equation (3)
was set as 60 min) and a semi-diurnal tidal oscillation with a 1 m
amplitude, as shown in Fig. 13. The output point for the numerical result
was placed at the center of a port basin with an area of 350 � 215 m.
After 60 min, the tsunami reached its highest peak with a resultant water
level reaching 2.5 m. Water-level variations inside the port showed
Fig. 12. Hydraulic head-to-discharge formula, which varies with the gate-gap excellent agreement with each other, validating the proposed formula.
width, derived for widths of 10, 15, 20, and 30 cm. In both results, the sharp tsunami signal almost disappeared inside the

9
H. Takagi et al. Ocean Engineering 199 (2020) 107028

Fig. 13. Comparison between numerical and simplified models of a tsunami


intrusion. The Delft3D model is used for the numerical simulation, while the
Fig. 14. A white noise tsunami generated by combining 60 different periodic
simplified discharge model is formulated on an Excel spreadsheet.
waves with equal amplitudes of 20 cm. The signal response inside the port was
calculated using the spreadsheet.
port, clearly showing the effectiveness of the self-elevating seawall. On
the other hand, the semi-diurnal tide penetrated the port shows only
slight attenuation. As discussed further in the following section, this
implies that the performance of the gate depends on wave frequency, as
it tends to admit longer-period waves.

3.2. Frequency analysis

To further determine to what extent wave transmission through a


self-elevating seawall depends on its frequency, a random tsunami was
used as a test case, generated by composing multiple component waves.
The range of long-period waves is typically 5 min to 12 h, covering
tsunamis to astronomical tides (Munk, 1950). In response to such a wide
range of periods, a total of 60 multiple component waves with equal
amplitudes (20 cm was assumed for this study) and different periods
were superimposed, resulting in a pseudorandom signal, as shown in
Fig. 14. The phase lag of each wave was determined by generating a
uniformly distributed random number. Here the test wave is referred as
a white noise tsunami because the signal is equivalent to random white
noise in signal processing. This test intends to examine the inflow
characteristics in terms of the wave periods. Tsunamis naturally consist Fig. 15. Spectrum of a white noise tsunami, represented by a transmission
of a number of transient, non-periodic, shorter, and longer waves. ratio, which was calculated by dividing the output signal by the input signal.
During the propagation of these waves from the ocean to a nearshore
area, these waves are gradually modified with respect to their ampli­ reduced because typically it lasts over a few hours, though shorter wave
tudes, wave lengths, and wave periods (Madsen et al., 2008). Therefore, components during a storm (e.g. wind waves, infragrabity waves) will be
this test that involves the use of a white noise tsunami may be mean­ much reduced. On the other hand, a tsunami component, which is
ingful to investigate the performance of the self-elevating seawall typically shorter than an hour, will be efficiently attenuated on passing
against tsunamis composed of many different wave components, such as the gate gap. For this case, the tsunami amplitude was reduced by 80%
those shown in Fig. 7. The thick red line in Fig. 14 shows the output and 90% with ten and five gaps, respectively. The self-elevating seawall
signal, which corresponds to the water level inside a port protected by a effectively functions as a low-pass filter system, though it may not
self-elevating seawall with five gaps (each with a 30 cm width). completely hold back seawater.
Short-period waves found in the input signal appear to mostly disappear,
whereas a long-period fluctuation remains visible. 3.3. Case study
A more detailed response is shown in Fig. 15, which is the result of a
fast Fourier transform (FFT). Rather than showing an ordinary ampli­ All the observed tsunamis shown in Fig. 7 include a high degree of
tude spectrum, the vertical axis indicates the transmission ratio of each uncertain fluctuations. Therefore, a detailed design must be conducted
wave component, which was calculated by dividing the output ampli­ by incorporating a more realistic time series of tsunami fluctuation
tude (inside the port) by the input amplitude (outside the port, 20 cm). A rather than a distinctive solitary wave, such as the tsunami model shown
longer wave component, which is associated with the astronomical tide, in Fig. 13. For example, the Cabinet Office of Japan has publicized
can be sufficiently reduced if there is only one gap. However, astro­ various tsunami scenarios that may occur along the coast of western
nomical tidal components cannot be eliminated as the number of gaps Japan during the imminent megathrust earthquake, known as the
increases. Likewise a typhoon storm surge may not be substantially Nankai Trough Earthquake, which is anticipated to happen within the

10
H. Takagi et al. Ocean Engineering 199 (2020) 107028

next 30 years at a likelihood of 80% (http://www.bousai.go.jp/jishi be the most important factor to make the planned project to be success
n/nankai/). In the case that the authority responsible for overseeing by finding a suitable port for installation of self-elevating seawall.
tsunami disasters has already established such future tsunami scenarios,
engineers may directly use the official scenarios without performing any
advanced simulations themselves. 3.4. Discussion - extra considerations required
Fig. 3 shows that a 2-m level tsunami, which is one of the scenarios of
the Cabinet Office of Japan, could be completely hold back by con­ The fewer the gate gaps, the lower the wave transmission. To mini­
structing a self-elevating seawall in the opening of a port (RGSS, 2016). mize water intrusion induced by tsunamis and tides, it is important to
The study by Takagi and Bricker (2014) also supports these expectations reduce the number of gate gaps as much as possible. However, a port
through their hypothetical analysis applied to Ishinomaki Fishery Port, opening of an order of 100 m will be protected by at least several
where the 2011 Tohoku tsunami devastated the town behind the port. separate gates of the self-elevating seawall system (RGSS, 2016).
However, their analysis essentially neglected possibility of tsunami Therefore, planners and engineers in charge must firstly recognize that
intrusion through various voids such as gate gap and breakwater rubble tsunami intrusion is unavoidable to some extent even if the port is
mound. protected by a well-designed self-elevating seawall. A tsunami mitiga­
Therefore, the authors re-evaluated effectiveness of the self-elevating tion plan should be planned to minimize the impact of a tsunami inside
seawall system by considering the combined influence of tsunami the port area, rather than aim for total elimination of tsunamis.
intrusion through the gate gap and the mound. The port geometry is Once water is allowed to enter to some extent, engineers must
similar with the one studied by RGSS (2016), as shown in Fig. 16. consider additional design issues. For example, a strong water jet
Equations (1) and (4) were used to evaluate inflow discharge through through a narrow gate gap might cause scouring/erosion behind the self-
the mound and the gap, respectively, and these results were linearly elevating seawall. Such strong flows might even destabilize ships on the
combined to calculate increase/decrease in water level inside the port port basin, and thus an appropriate design for ensuring that ships are
with its basin of 7.5 ha. The breakwater is composed of caisson and safely moored will also be required. Although this is outside the scope of
rubble mound. As the caisson is an impermeable structure, only the the present paper, the calculated discharge with the proposed formula
submerged mound layer with a 2-m thickness is considered as a can be used as an input boundary for computing velocity fields within
permeable part. This port is assumed to be enclosed by a total of three the port basin. Such analysis will be highly useful for investigating the
self-elevating gate units. As a result, 4 gate gaps (each with a 30 cm influence of tsunami intrusion on moored ships or estimating scouring
width) are created. The discharge coefficient can be calculated by behind the gate.
applying Equation (9). Given that it takes 15–20 min to mechanically uplift the embedded
Fig. 17 shows how the water level inside the port changes with water gate, it is required for the disaster management authority to immedi­
level outside, which is composed of a Level 1 tsunami for this port and ately start the operation after the earthquake. However, it should be
astronomical tide (a semi-diurnal tide). The time-history of this potential recognized that any modern system, including self-elevating seawall, is
tsunami was originally provided by a prefectural office, which was based unlikely to work against a very short-warning time tsunami, such as that
on the fundamental scenarios presented by the Cabinet Office of Japan triggered by submarine and/or subaerial landslide (Takagi et al., 2019b;
(http://www.bousai.go.jp/jishin/nankai/taisaku/pdf/1_1.pdf). While Heidarzadeh et al., 2020).
the longest wave component induced by the tide is not significantly The risk of a gate-opening-induced tsunami should also be addressed.
reduced, the rapid fluctuation due to the tsunami is mostly vanished This secondary tsunami could potentially occur if the gate is uncon­
within the port basin. Unlike the assessment by RGSS (2016) under the sciously reopened. For example, if the water level outside the port is
ideal assumption, tsunami substantially intrudes into the port. However, about to drop to zero at 360 min as shown in Fig. 13, the port authorities
the peak of the tsunami can be reduced by about 90 cm, which would may decide to reopen the gate at this moment because rapid resumption
mitigate the most powerful impact caused by the tsunami. Although this of port operations is critical. However, the water level inside the port
is highly dependent on the layout of port, Fig. 16 demonstrates that the still exceeds 50 cm. If the gate is suddenly opened, a gate-opening-
predominant influence is not caused by the inflow through the gate gap, induced tsunami could be generated, which could result in many unfa­
but that through the rubble mound. Hence, the preliminary survey will vorable effects such as a strong drag on moored vessels and a returning
tsunami engulfing adjacent coastal areas. To avoid such unexpected

Fig. 16. Case study of installation of self-elevating seawall in a medium sized fishery port.

11
H. Takagi et al. Ocean Engineering 199 (2020) 107028

Fig. 17. Estimated water levels inside the port, when tsunami intrusion through the gate gaps and rubble mound take place.

impacts, the port authorities must wait for the water levels outside and sea side and the port side of the seawall, h is the water depth, and w is the
inside the port to coincide (e.g., at 600 min in Fig. 13). gap width. The correction factor γ may be modified via experiment. The
A self-elevating seawall can work only when adjacent port facilities discharge coefficient of C can be estimated by a regression formula,
also work against earthquakes as well as tsunamis. Out of 67 port which varies with the gap width between the relevant floodgates. For
breakwaters in the Tohoku region, 29 suffered damage due to the 2011 example, the relationship C ¼ 0:15 ​ Δh þ 0:25 was derived for the case
tsunami (Takagi, 2015). Many floodgates even withstood the 2011 Great where w ¼ 0.3m. The proposed model was verified through case studies
East Japan earthquake and tsunami (e.g., those in Minami Sanriku, in which a time series of hypothetical water levels composed of a Level 1
Japan, Fig. 18), but their adjacent coastal structures were destroyed, tsunami and semi-diurnal tide was simulated for a medium-sized port.
nullifying the function of the gate itself. Such disastrous incidents may With this proposed method and given the water level outside the port,
occur unless the adjacent structure is also carefully examined and the water level rising over a port basin enclosed by floodgates and
improved if necessary, as a part of the self-elevating seawall system. breakwaters can be calculated using a spreadsheet that do not require
expensive computational fluid dynamics. A frequency analysis is also
4. Conclusions performed to investigate to what extent the performance of the system is
influenced by differences in wave period. A self-elevating seawall system
Self-elevating seawalls are attracting attention as effective counter­ acts as an effective low-pass filter system against tsunamis, though it
measures against tsunamis, particularly after the Great East Japan may not completely hold back much longer wave components such as
earthquake and tsunami in 2011. However, a narrow gap between tides and storm surges. Last, but not least, the tsunami waveforms
adjacent gate units would inevitably cause an intrusion of seawater into assumed in this study may not represent an impulsive solitary wave, and
the port basin, which cannot be simply simulated using common therefore the derived formula should not be applied to such a violent
tsunami models because the gap is too small to be reproduced in the tsunami.
model. The present study examines the water flow passing through a
gate gap by applying a 2D-3D hybrid hydrodynamic model to derive the Declaration of competing interest
hydraulic head-to-inflow discharge conversion formula, which is
pffiffiffiffiffiffiffiffiffiffi
defined as Q ¼ γC 2gΔh ​ w ​ h ​ , where Q is the discharge, g is gravi­ We have no conflict of interest to declare.
tational acceleration, Δh is the hydraulic head difference between the

Fig. 18. Floodgates in Minami Sanriku City that withstood the 2011 Great East Japan tsunami, though their adjacent protection was totally destroyed (photos taken
by the author in April 2011).

12
H. Takagi et al. Ocean Engineering 199 (2020) 107028

CRediT authorship contribution statement Kriebel, D.L., 1992. Vertical wave barriers: wave transmission and wave forces. In:
Proceedings of 23th Conference on Coastal Engineering, ASCE. ASCE, New York,
pp. 1313–1326.
Ryoichi Tomiyasu: Conceptualization, Writing - review & editing. Latcharote, P., Suppasri, A., Hasekawa, N., Takagi, H., Imamura, F., 2016. Effect of
Tomoyuki Oyake: Conceptualization, Writing - review & editing. breakwaters on loss reduction of fatality ratio during the 2011 Great East Japan
Taketo Araki: Conceptualization, Writing - review & editing. Kyosuke earthquake and tsunami. J. Jpn. Soc. Civil. Eng., Ser. B2 (Coast. Eng.) 72 (2),
1591–1596.
Mori: Conceptualization, Writing - review & editing. Yasuhiro Mat­ Lesser, G.R., Roelvink, J.A., van Kester, J.A.T.M., Stelling, G.S., 2004. Development and
subara: Conceptualization, Writing - review & editing. Yohei Nino­ validation of a three-dimensional morphological model. Coast. Eng. 51, 883–915.
miya: Conceptualization, Writing - review & editing. Yoshifumi Madsen, P.A., Fuhrman, D.R., Schaffer, H.A., 2008. On the solitary wave paradigm for
tsunamis. J. Geophys. Res. 113 (22p), C12012.
Takata: Conceptualization, Writing - review & editing. Mei, C.C., Stiassnie, M., Yue, D.K., 2005. Theory and Applications of Ocean Surface
Waves (Advanced Series on Ocean Engineering). World Scientific.
Acknowledgments Mori, N., Takahashi, T., 2012. The 2011 Tohoku earthquake tsunami joint survey group.
Nationwide post event survey and analysis of the 2011 Tohoku earthquake tsunami.
Coast Eng. J. 54 (No. 1), 27.
Funding for this research was supported by the grants to Tokyo Moura, N., 2018. Flood Prevention Tidal Gates Near Inner Harbour to Be Able to Sustain
Institute of Technology from the JSPS KAKENHI (No.16KK0121 and 5.8m High Flood. Macau Business accessed on January 20, 2020. https://www.maca
ubusiness.com/macau-flood-prevention-tidal-gates-near-inner-harbour-able-sustain-
19K04964) and the Maeda Engineering Foundation. The spreadsheet 5-8mtr-high-flood/.
can be shared upon request (takagi@ide.titech.ac.jp). Munk, W.H., 1950. Origin and generation of waves. In: Proceedings of First Conference
on Coastal Engineering, pp. 1–4. California, US.
Nakashima, S., Takayama, T., Obara, K., Kawasaki, T., Kurokawa, F., Onodera, T., 2011.
References
Performance based design of vertically telescopic breakwater for protection from
tsunami. J. Jpn. Soc. Civil. Eng., Ser. B2 (Coast. Eng.) 67 (2), 786–790.
Abe, S., Oyake, T., Suzuki, M., Otsuki, N., 2017. Applicability of pneumatic caisson Nistor, I., Palermo, D., Nouri, Y., Murty, T., Saatcioglu, M., 2009. Tsunami-induced
method. In: 42nd Conference on Our World in Concrete & Structures, vol 10. forces on structures. In: Book: Handbook of Coastal and Ocean Engineering. World
Singapore. Scientific, pp. 261–286.
Arikawa, T., Sato, M., Shimosako, K., Tomita, T., Tatsumi, D., Yeom, G., Takahashi, K., PIANC, 2006. Design of movable weirs and storm surge barriers, p. 123. Report of
2012. Investigation of the Failure Mechanism of Kamaishi Breakwaters Due to Working Group 26 of the Inland Navigation Commission.
Tsunami: Initial Report Focusing on Hydraulic Characteristics, Technical Note No. Raby, A., Macabuag, J., Pomonis, A., Wilkinson, S., Rossetto, T., 2015. Implications of
1251 of the Port and Airport Research Institute. Yokosuka, Japan. the 2011 Great East Japan tsunami on sea defence design. Int. J. Disaster Risk
Bricker, J.D., Takagi, H., Mitsui, J., 2013. Turbulence model effects on VOF analysis of Reduct. 14, 332–346.
breakwater overtopping during the 2011 Great East Japan Tsunami. In: Proceedings Research Group for Self-elevating Seawall (RGSS), 2016. Research and Development of
of the 2013 IAHR World Congress, p. 10p. Sichuan, China. Steel Self-Elevating Seawall, the 1st Technical Meeting on the Steel Self-Elevating
Bricker, J.D., Gibson, S., Takagi, H., Imamura, F., 2015. On the need for larger Manning’s Seawall, vol. 57. Ohita, Japan in Japanese).
roughness coefficients in depth-integrated tsunami inundation models. Coast Eng. J. Sasaki, J., Ito, K., Suzuki, T., Wiyono, U.A., Oda, Y., Takayama, Y., Yokota, K., Furuta, A.,
57 (13p). Takagi, H., 2012. Behavior of the 2011 Tohoku earthquake tsunami and resultant
Chow, V.T., 1959. Open-channel Hydraulics. McGRAW-HILL, p. 680p. damage in Tokyo bay. Coast Eng. J. 54 (26p).
Cyranoski, D., 2012. After the deluge: Japan is rebuilding its coastal cities to protect Shibayama, T., Esteban, M., Nistor, I., Takagi, H., Nguyen, D.T., Matsumaru, R.,
people from the biggest tsunamis. Nature 483, 141–143. Mikami, M., Aranguiz, R., Jayaratne, R., Ohira, K., 2013. Classification of tsunami
Dames & Moore, 1980. Design and Construction Standards for Residential Construction and evacuation areas. Nat. Hazards 67 (2), 365–386.
in Tsunami-Prone Areas in Hawaii, Prepared by Dames and Moore Inc. for the Takagi, H., Esteban, M., 2013. Practical methods of estimating tilting failure of caisson
Federal Emergency ManagementAgency. breakwaters using a monte-carlo simulation. Coast Eng. J. 55 (3), 22p.
Deltares, 2011. Delft3D-FLOW – Simulation of Multi-Dimensional Hydrodynamic Flows Takagi, H., Bricker, J., 2014. Assessment of the effectiveness of general breakwaters in
and Transport Phenomena, Including Sediments, vol 690p. User Manual Delft3D- reducing tsunami inundation in Ishinomaki. Coast Eng. J. 56 (No. 4), 21p.
FLOW. Takagi, H., 2015. Method for quick assessment of caisson breakwater failure due to
Dircke, P.T.M., Jongeling, T.H.G., Jansen, P.L.M., 2012. An Overview and Comparison of tsunamis: retrospective analysis of data from the 2011 Great East Japan earthquake
Navigable Storm Surge Barriers. The Innovative Dam and Levee Design and and tsunami. Coast Eng. J. 15p.
Construction for Sustainable Water Management, pp. 65–87. New Orleans, LA. Takagi, H., Esteban, M., Mikami, T., Fujii, D., 2016. Projection of coastal floods in 2050
Erbisti, P.C.F., 2004. Design of Hydraulics Gates. A.A. Balkema Publishers, p. 351. Jakarta. Urban Clim. 17, 135–145.
Ergun, S., 1952. Fluid flow through packed columns. Chem. Eng. Prog. 48 (2). Takagi, H., Yi, X., Fan, J., 2019a. Public Perception of Typhoon Signals and Response in
Esteban, M., Takagi, H., Shibayama, T., 2015. Handbook of Coastal Disaster Mitigation Macau: Did Disaster Response Improve between the 2017 Hato and 2018 Mangkhut
for Engineers and Planners. Elsevier, p. 780. Typhoons? Georisk.
Esteban, M., Glasbergen, T., Takabatake, T., Hofland, B., Nishizaki, S., Nishida, Y., Takagi, H., Pratama, M.B., Kurobe, S., Esteban, M., Ar� anguiz, R., Ke, B., 2019b. Analysis
Stolle, J., Nistor, I., Bricker, J., Takagi, H., Shibayama, T., 2017. Overtopping of of generation and arrival time of landslide tsunami to Palu City due to the 2018
coastal structures by tsunami waves. Geosciences 7 (121). Sulawesi earthquake. Landslides 16 (5), 983–991.
Fujii, N., Araki, T., Takagi, H., Tomiyasu, R., Oyake, T., Suzuki, M., Otsuki, N., 2018. Takata, Y., Takagi, H., 2019. Hydraulic experiment and numerical analysis on tsunami/
Self-elevating Sea Wall Constructed with Pneumatic Caisson Method, 43rd storm surge entering through breakwater mound. J. Jpn Soc. Civ. Eng. Ser. B3
Conference on Our World in Concrete & Structures. Singapore. (Ocean Eng.) 75 (2) in Japanese.
Gower, J., 2005. Jason 1 detects the 26 december 2004 tsunami. Eos 86, 37–38. Tanimoto, K., Takayama, T., Murakami, K., Murata, S., Tsuruya, H., Takahashi, S.,
Heidarzadeh, M., Ishibe, T., Sandanbata, O., Muhari, A., Wijanarto, A.B., 2020. Morikawa, M., Yoshimoto, Y., Nakano, S., Hiraishi, T., 1983. Field Laboratory
Numerical modeling of the subaerial landslide source of the 22 December 2018 Anak Investigations of the Tsunami Caused by 1983 Nihonkai Chubu Earthquake, vol 470.
Krakatoa volcanic tsunami, Indonesia. Ocean Eng. 195 (1). Technical Note of the Port and Harbour Research Institute, Japan, No, p. 299 in
Hofland, B., Marissen, R., Bergsma, O., 2015. Dynamic behaviour of a flexible membrane Japanese).
tsunami Barrier with Dyneema®. In: Proceedings of Coastal Structures & Solutions to Tanimoto, K., Goda, Y., 1991. Historical Development of Breakwater Structures in the
Coastal Disasters Joint Conference. 12p., Boston. World, Coastal Structures and Breakwaters. Thomas Thelford, London, pp. 193–220.
Kakuno, S., Liu, P.-F., 1993. Scattering of water waves by vertical cylinders. J. Waterw. The Cabinet Office of Japan, 2011. Report of the Committee for Technical Investigation
Port, Coast. Ocean Eng. 119 (3), 302–322. on Countermeasures for Earthquakes and Tsunamis Based on the Lessons Learned
Kimura, Y., Niizato, H., Nakayasu, K., Yasuda, T., Mori, N., Mase, H., 2010. Numerical from the “2011 off the Pacific Coast of Tohoku Earthquake”.
Model for Flap-Gate Response to Tsunami and its Verification by Hydraulic UNESCO, 1997. Numerical Method of Tsunami Simulation with the Leap-Frog Scheme,
Experiments. In: 32nd International Conference on Coastal Engineering. Shanghai. IUGG/IOC Time Project, vol 35. IOC Manuals and Guides.

13

Das könnte Ihnen auch gefallen