Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Marine Geology 427 (2020) 106225

Contents lists available at ScienceDirect

Marine Geology
journal homepage: www.elsevier.com/locate/margo

Numerical simulation of sedimentation and erosion caused by the 2011 T


Tohoku Tsunami in Oarai Port, Japan
Yoshiaki Kuriyamaa, , Yu Chidaa, Yoshiyuki Unob, Kazuhiko Hondac

a
Port and Airport Research Institute, Nagase 3-1-1, Yokosuka, Kanagawa 239-0826, Japan
b
ECOH Co., Ltd., Kita-Ueno 2-6-4, Taito-ku, Tokyo 110-0014, Japan
c
National Institute for Land and Infrastructure Management, Nagase 3-1-1, Yokosuka, Kanagawa 239-0826, Japan

ARTICLE INFO ABSTRACT

Keywords: The 2011 Tohoku Tsunami induced sediment deposition of 334,000 m3 mainly in the basin and 9 m erosion at
Bathymetry change the tips of a breakwater and a groin in Oarai Port, Japan. The tsunami-induced bathymetry change was simu-
Vortex lated using a numerical model and the simulated bathymetry change was compared with the measured one. The
Suspended sediment transport comparison showed that the simulated bathymetry change quantitatively agreed with the measured one in-
Basin
cluding the sediment deposition and the erosion mentioned above although the erosion measured along the
Scour
Inundation height
breakwater was not well reproduced by the model. The time series of simulated bathymetry change, water level,
current and suspended sediment concentration showed that the sediment deposition in the center of the basin
was mostly induced by the sediments that moved into the basin through its entrance between the breakwater and
the groin and were redistributed by counterclockwise vortices. The severe erosion at the tip of the breakwater
was mainly induced by the strong southward currents.

1. Introduction where the tsunami heights were 2 to 2.5 m (Wilson et al., 2012).
The erosion of sandy beaches deteriorates their ability to prevent or
Over the last couple of decades, the topography and bathymetry of mitigate coastal disasters by dissipating wave energy. Sediment de-
some sandy beaches have been greatly altered by large tsunamis such as positions on land can be obstacles to rescue and recovery efforts. The
the 1998 Papua New Guinea Tsunami (e.g., Gelfenboum and Jaffe, bathymetry changes inside ports and harbors are also problematic.
2003), the 2004 Indian Ocean Tsunami (e.g., Borrero, 2005; Umitsu Erosion and scour can be threats to the stability of breakwaters.
et al., 2007; Paris et al., 2009; Meilianda et al., 2010) and the 2011 off Sedimentation may interfere with vessel traffic and prevent the delivery
the Pacific coast of Tohoku Tsunami (e.g., Tanaka et al., 2012; Tappin of relief supplies through ports and harbors. To investigate counter-
et al., 2012; Udo et al., 2012; Adityawan et al., 2014). Following these measures against such changes, being able to accurately predict tsu-
events, sandy beaches were heavily eroded and huge amounts of sedi- nami-induced morphological changes and understand their processes is
ment were transported landward and seaward and deposited. required. Numerical simulation models for sediment transport and
Ports and harbors constructed on sandy beaches have also been morphological change due to tsunami were developed and applied to
affected by tsunamis in recent years. The 2004 Indian Ocean Tsunami those caused by the 1960 Chilean Tsunami (e.g., Takahashi et al.,
induced an erosion of 3 m inside Kirinda Harbor in Sri Lanka, where the 2000), the 2004 Indian Ocean Tsunami (e.g., Apotsos et al., 2011; Goto
tsunami height reached approximately 6 m (Goto et al., 2011). During et al., 2011; Gusman et al., 2012; Li et al., 2012; Ontowirjo et al., 2013)
the 2011 Tohoku Tsunami, the tsunami heights reached 4.5 to 6 m in and the 2011 Tohoku Tsunami (e.g., Sugawara and Takahashi, 2013;
Hachinohe, Hitachi and Oarai Ports, leading scours of up to 7 to 10 m Kuriyama et al., 2014; Sugawara et al., 2014a, 2014b; Yamashita et al.,
developing at the gaps between breakwaters and at their tips 2016; Gusman et al., 2018; Sugawara, 2018).
(Kuriyama, 2014). Further, around the centers of the Hitachi and Oarai Tsunami-induced bathymetry changes in and around ports and
Port basins, sediment depositions of about 1 to 2 m were observed. The harbors have also been simulated using numerical models. Kihara et al.
2011 Tohoku Tsunami also induced scours of 2 to 4 m and depositions (2012) calculated the bathymetry change in and around a harbor with
of 1 to 4 m in Crescent City and Santa Cruz Harbors in California, USA, two breakwaters at the laboratory scale using a three dimensional

Corresponding author.

E-mail addresses: kuriyama@p.mpat.go.jp (Y. Kuriyama), chida-y@p.mpat.go.jp (Y. Chida), yoshiyuki_uno@ecoh.co.jp (Y. Uno),
honda-k852a@mlit.go.jp (K. Honda).

https://doi.org/10.1016/j.margeo.2020.106225
Received 1 February 2020; Received in revised form 6 May 2020; Accepted 8 May 2020
Available online 15 May 2020
0025-3227/ © 2020 Elsevier B.V. All rights reserved.
Y. Kuriyama, et al. Marine Geology 427 (2020) 106225

hydrodynamic model and compared the calculated result with that


measured by Fujii et al. (2009). The model well reproduced the scour at
the harbor entrance and the deposition in the center of the basin.
Nishihata et al. (2006), Kihara and Matsuyama (2010) and Ranasinghe
et al. (2013) applied numerical models to the bathymetry change in
Kirinda Harbor mentioned above. Their models were slightly different,
but they successfully reproduced the relatively large erosion near the
central breakwater and the sediment deposition near the shoreline be-
tween the central and south breakwaters.
To investigate the accuracy of a numerical model for tsunami-in-
duced bathymetry change in a port with channels and basins at the
prototype scale, Kuriyama et al. (2015) conducted a numerical simu-
lation of the bathymetry change in Oarai Port induced by the 2011
Tohoku Tsunami mentioned above by using the model developed by
Kuriyama et al. (2014). The model underpredicted the bathymetry
change partially because the calculation period of 180 min was too
short. A second cause of the underestimation would be that the initial
sea level displacement generated by the tsunami source employed in
Kuriyama et al. (2015) might have been a little bit smaller than the
actual one. The calculated inundation heights were smaller than the
measured ones. Another cause of the discrepancy might be that the
model was one-way, not two-way. In a two-way model, the bathymetry
calculated at a time step is used in the tsunami propagation calculation
at the next step. Although the model developed by Kuriyama et al.
(2014) was two-way, the one used in Kuriyama et al. (2015) was one-
way.
The main purpose of this study is to investigate the mechanisms of
erosion and accretion in Oarai Port induced by the 2011 Tohoku
Tsunami using simulation results. First, we improve the simulation used Fig. 1. Location of study site.
in Kuriyama et al. (2015) by extending the calculation period, adjusting
the initial sea level displacement and using a two-way model. Then,
after confirming the accuracy of the model by comparing the simulated
and measured bathymetry changes, we discuss the processes of the
bathymetry changes including scours and sediment depositions in the
channel and the basin.

2. Study site

Oarai Port is located in eastern Japan facing the Pacific Ocean and
285 km south of the epicenter of the 2011 Tohoku Earthquake (Fig. 1).
The tsunami height at Oarai was approximately 5 m. The bathymetry in
Oarai Port was measured in August 2010, seven months before the 2011
Tohoku Tsunami, and between 20 and 22 March 2011, approximately
ten days after the tsunami. Inside the port, scours were observed at the
tips of and along the South Breakwater and the West Groin (Fig. 2). The
maximum scour depth was 9 m. Sediment deposition took place inside
the port and at the area between the South Breakwater and the West
Groin. The erosion and the sedimentation volumes in the investigation
area were 1,037,000 m3 and 334,000 m3, respectively. The volume
change induced by the land subsidence due to the 2011 earthquake was
excluded from the values mentioned above. The amount of land sub-
sidence in the study site was assumed to be 0.25 m as it was 0.25 m at a
nearby city, Hitachinaka (Geospatial Information Authority of Japan,
2011).

3. Numerical simulation for tsunami propagation and bathymetry


change
Fig. 2. Measured bathymetry change. Letters A to H indicate eight areas set for
further analysis.
3.1. Outline of the model

The numerical simulation model for the tsunami propagation and submodel of sediment transport and bathymetry change.
bathymetry change used in this study is based on those developed by The submodel predicts elevation changes from the difference be-
Kuriyama et al. (2014, 2015). It is the two-way, one-layer hydrostatic tween the amount of sediment settling from suspension and the amount
version of the Storm surge and Tsunami simulator in Oceans and picked up, and the difference between incoming and outgoing bed se-
Coastal areas (STOC) (Tomita et al., 2006; Tomita and Honda, 2007), diment transport rates.
which is a Reynolds Averaged Navier-Stokes model, coupled with a

2
Y. Kuriyama, et al. Marine Geology 427 (2020) 106225

Zb 1 (Qb, v + Qb, a )
= ws Cb P+
t 1 x (1)

where zb is elevation, t is time, λ is porosity (= 0.4), ws is the set-


tling velocity of sediment, Cb is the suspended sediment concentration
at the bottom, P is the pick-up rate, which determines the amount of
sediments picked up from the bottom, and Qb,v and Qb,a are the bed
sediment transport rates due to velocity and acceleration, respectively.
The pick-up rate P was estimated by using Eq. (2), which is based on
Nielsen's formula (Nielsen, 1992):

P = c1 ws ( cr )
1.5 , = u 2/ sgd (2)

where c1 is a free parameter, θ is the Shields parameter, θcr is the cri-


tical Shields parameter, u⁎ is the friction velocity, s is the specific
gravity of sediment in water, g is the gravitational acceleration, and d is
the sediment particle diameter.
The rest of the details of the submodel of sediment transport and
bathymetry change including the bed sediment transport rates, the
suspended sediment concentration at the bottom and the friction ve-
locity is described in Appendix A. Fig. 3. Initial topography and bathymetry and locations of stations for mon-
The governing equations were solved by the finite difference itoring water level and current (solid circles) and inundation height (open
method; a staggered mesh in space and a leapfrog method in time were circles).
applied. The discretization of conservative terms was carried out with
the control volume method. The diffusion terms were discretized with
the second-order central scheme, while the advection terms were ex-
panded with the first-order upwind scheme. The elevations calculated
using Eq. (1) at a time step were used in the calculation of tsunami
propagation at the next step.
The simulations were conducted using a seven-stage nested grid
system, in which the grid size was changed in succession from 2592 m
to 864 m, 288 m, 96 m, 48 m, 12 m, and 4 m. The values of water
surface elevation, current velocity and suspended sediment concentra-
tion were exchanged on the boundaries between two nested grid re-
gions. At the open boundaries in the largest grid region, free trans-
mission of tsunami waves was assumed. The initial water level was
assumed to be the sum of the tidal level when the tsunami reached the
study site and the water level displacement induced by the earthquake.
The current velocities and suspended sediment concentrations were
initially assumed to be zero.

3.2. Setup of the model

The calculation period and the time interval were set at 720 min
(12h) and 0.1 s, respectively. The sediment size was determined to be
0.2 mm based on Nemoto et al. (2009), Yazawa et al. (2009) and
Hanzawa (2017). The critical Shields parameter was assumed to be
0.05. The parameter c1 was set at 0.00035, which is 1/10 of the value Fig. 4. Comparison between measured and simulated inundation heights.
used by Kuriyama et al. (2014). The difference will be discussed in 5. Triangles and solid and open circles indicate the values simulated using the
initial sea level displacement 1.0, 1.2 and 1.4 times that of Takagawa and
Discussion. The morphological data used in the calculation (Fig. 3)
Tomita (2012), respectively.
were provided by the Central Disaster Prevention Council of the Japa-
nese Government. Inside the port, the bathymetry measured in August
2010 was used. The topographical data in the finest and the second 4. Results
finest grid regions were based on the survey data obtained by the
Geospatial Information Authority of Japan. 4.1. Water level and velocity
The tsunami propagation calculation used the initial sea level dis-
placement multiplied by a ratio of 1.2 obtained by Takagawa and At around 18 min after the earthquake occurred, the water level at
Tomita (2012), who applied a waveform inversion to offshore tsunami Station 1 (Fig. 3) started to change due to the resulting tsunami
records. This is because the comparisons between the inundation (Fig. 5(a)). The water level reached 3.27 m at 54 min and then dropped
heights measured at the study site (Mori et al., 2012) (see the locations to −3.47 m at 74 min. During the first 600 min, the port experienced
in Fig. 3) and those simulated using the initial sea level displacement approximately ten subsequent waves. Inside the port, the water levels at
1.0, 1.2 and 1.4 times that of Takagawa and Tomita (2012) show that Stations 3 to 6 changed similarly to that at Station 1. However, the
the inundation heights calculated using Takagawa and Tomita's value water level changes at Stations 3 to 6 lagged 5 to 10 min behind that at
were smaller than the measured values and that those calculated using Station 1 and its amplitudes at Stations 3 to 6 were slightly larger than
the initial sea level displacement 1.2 times that of Takagawa and To- that at Station 1.
mita agreed well with the measured values (Fig. 4). Near the tip of the South Breakwater, at Station 2, strong eastward

3
Y. Kuriyama, et al. Marine Geology 427 (2020) 106225

velocities, respectively. The subscript numbers indicate the station


numbers shown in Fig. 3.

4.2. Bathymetry change

The comparison between the measured (Fig. 9a) and simulated


(Fig. 9b) bathymetry changes shows that the model well reproduced the
tsunami-caused bathymetry change including the sediment deposition
in the center of the basin and the erosion at the tips of the South
Breakwater and the West Groin. However, the erosion along the South
Breakwater was not well reproduced by the model.
For more detailed analyses, the area where the bathymetry data
were obtained was divided into eight areas according to the properties
of the measured elevation changes (Fig. 2). Areas D, E and G experi-
enced severe erosion and Area F had a relatively large sediment de-
position. The time series of the volume changes in the eight areas show
that the tsunami-induced bathymetry changes were almost complete by
500 min (Fig. 10).
The simulated volume changes in the eight areas quantitatively
agreed well with the measured ones (Fig. 11). The correlation coeffi-
cient r was 0.93. The linear regression coefficient a1 of the simulated
volume changes on the measured ones was 1.1. These results indicate
that the model predicted the tsunami-induced bathymetry change in
and around the port with high accuracy.
The mechanisms of the erosion in Area D and the sediment de-
position in Area F are now discussed. The time series of the volume
change rate in Area D showed that the erosion rate had seven peaks at
54, 62, 84, 118, 182, 208 and 306 min (Fig. 12) and that these peaks
seem to have been correlated with the current velocity averaged along
Line aa’. The relationship between the volume change rate in Area D
and the current velocity indicates that the erosion at the tip of the South
Breakwater was caused by strong currents (Fig. 13). The strong currents
moved mostly southward in particular when the current velocities were
larger than 4 m/s (Fig. 14).
To further investigate the mechanism of the erosion, we focused on
the seven peaks of the erosion in Area D and examined the volume
change rate in Area D, the water level and the current velocity and
direction during 14 min before and after the peaks. The water level and
Fig. 5. (a) Water level at Stations 1 (thin purple line) and 3 (thick red line). (b)
the current velocity and direction were averaged along Line aa’. The
East-west (thick black line) and north-south (thin black line) velocities at severe erosion mostly took place when the strong southward current
Station 2. East-west (c) and north-south (d) velocities at Stations 3 (red line) developed as mentioned above and the water level started to increase
and 5 (green line). East-west (e) and north-south (f) velocities at Stations 4 from a trough (Fig. 15).
(orange line) and 6 (blue line). (For interpretation of the references to colour in As an example of the spatial distributions of the current and the
this figure legend, the reader is referred to the web version of this article.) depth-averaged suspended sediment concentration in and around Area
D, those at 84 min, when the erosion rate was maximum, are displayed
and southward currents were developed (Fig. 5(b)) and the velocity in Fig. 16. The current entered the port area over the northern tip of the
reached 8.43 m/s at 176 min. Inside the port, the current directions Offshore Breakwater. It then strongly developed southward over the tip
showed frequent turnover (Fig. 5(c) and (d)). The currents tended to of the South Breakwater and large amounts of sediment were sus-
develop northward at Station 3, but southward at Station 5. A similar pended. Then, the current moved south-westward and subsequently
velocity pattern was also observed at Stations 4 and 6. Westward and changed direction to move north-east and into the basin. This pattern
eastward currents developed at Stations 4 and 6, respectively (Fig. 5(e) was also observed more or less at 54, 118, 182, 208 and 306 min. The
and (f)). These results indicate the generation of vortices, which is only exception is the erosion at 62 min, when the current developed
partially supported by Fig. 6. The photo shown in Fig. 6, taken at eastward as shown in Fig. 17.
96 min, demonstrates that a counterclockwise vortex was generated in The volume change rate in Area F had five peaks of deposition at 94,
the basin. This vortex was also reproduced by the simulation as shown 124, 182, 218 and 310 min (Fig. 18). As in the analysis of the erosion in
in Fig. 7. This result suggests the high accuracy of the model in pre- Area D mentioned above, we investigated the volume change rate in
dicting tsunami-induced currents. The vorticity ω estimated by Eq. (3) Area F, the water level, the current velocity, the suspended sediment
shows that the counterclockwise vortices as shown in Figs. 6 and 7 as transport rate toward the basin and the vorticity ω defined by Eq. (3)
well as in Kim et al. (2013) dominantly developed due to the 2011 during 30 min before and after the peaks. The water level, the current
tsunami (Fig. 8). velocity and the suspended sediment transport rate were averaged
along Line bb’ located at the entrance of the basin. The results show that
=
v3 v5 u4 u6 sediments were deposited in Area F just after they moved into the basin
x3 x5 y4 y6 (3) through the entrance when the water level was increasing from a
trough to a crest (Fig. 19). The timings of the sediment depositions
where x and y are the distances in the east-west and north-south di- coincided with the development of vortices.
rections, respectively, and u and v are the east-west and north-south The correlations among the sediment deposition, the current and

4
Y. Kuriyama, et al. Marine Geology 427 (2020) 106225

Fig. 6. Aerial photo taken at 96 min.

the suspended sediment transport rate toward the basin were also seen
in the spatial distributions of the current and the depth-averaged sus-
pended sediment concentration at 178 and 182 min. At 178 min, when
the suspended sediment transport rate toward the basin at the entrance
had a peak, the current moved into the basin after coming over the tip
of the South Breakwater and turning direction over the tip of the West
Groin (Fig. 20). Large amounts of sediment were suspended and
transported into the basin through the entrance. Then, 4 min later at
182 min, when the deposition rate was at its maximum, the area of high
suspended sediment concentration moved northward in the basin with
the development of a counterclockwise vortex (Fig. 21).
Although most of the sediment depositions in Area F took place after
the transport of a large amount of sediment through the entrance as
mentioned above, even before such sediment influx to the basin, sedi-
ments were deposited in Area F from −22 min before the first peak
(Fig. 19). This sediment deposition was caused by sediments trans-
ported from a small vessel basin located north-west of Area F owing to a
clockwise vortex (Fig. 22), which developed when the water level al-
most reached the local minimum (Fig. 19). The development of this
clockwise vortex is also confirmed in Fig. 8.
The comparison between the sediment volumes deposited in Area F
Fig. 7. Simulated velocity at 96 min. The arrow indicates the direction in which and transported into the basin from the entrance through suspended
Fig. 6 was taken. and bed loads during 30 min before and after the five peaks shows that,
except for the first peak, 56% to 89% of the transported sediments were
deposited in Area F (Table 1).
In the simulations, both the suspended and bed sediment transport
rates were calculated. However, the suspended load was vastly

Table 1
Volumes of sediments deposited in Area F during 30 min before and after the
five peaks, those transported into the basin through the entrance and the ratios
of the volumes deposited to those transported.
Peak 1st 2nd 3rd 4th 5th

3
Volumes deposited (m ) 47,759 29,915 46,617 21,365 34,780
Fig. 8. Vorticity in the basin.
Volumes transported (m3) 25,520 41,744 52,510 30,863 61,831
Ratio (%) 187 72 89 69 56

5
Y. Kuriyama, et al. Marine Geology 427 (2020) 106225

Fig. 9. Bathymetry changes measured (a) and simulated (b). Letters A to H and symbols aa’ and bb’ indicate eight areas and two lines, respectively.

Fig. 10. Time-series of simulated volume changes in eight areas.

Fig. 12. (a) Volume change rate in Area D (see the location in Fig. 9). Water
level (b) and east-west (c, black line) and north-south (c, red line) velocities
averaged along Line aa’ (see the location in Fig. 9). The vertical broken lines
indicate the timings of the seven erosion rate peaks. (For interpretation of the
references to colour in this figure legend, the reader is referred to the web
version of this article.)

predominant in our calculations. The ratio of the suspended sediment


transport rate to the total sediment transport rate, which is the sum of
the suspended and bed transport rates, was over 95% in most of the
calculation domain.
Fig. 11. Comparison between volume changes measured and simulated in eight
5. Discussion
areas. The solid line indicates that the two values are equal.

Sediment depositions in Area F coincided with the development of

6
Y. Kuriyama, et al. Marine Geology 427 (2020) 106225

Fig. 13. Relationship between current velocity averaged along Line aa’ and
volume change rate in Area D.

Fig. 15. Volume change rate in Area D (a), water level (b), current velocity (c)
and current direction (d) during 14 min before and after peaks of volume
change rate. Water level and current velocity and direction were averaged
along Line aa’. Black, red, blue, green, orange, purple and pink lines show the
values for the first, second, third, fourth, fifth, sixth and seventh peaks, re-
spectively. (For interpretation of the references to colour in this figure legend,
the reader is referred to the web version of this article.)
Fig. 14. Relationship between current velocity and direction averaged along
Line aa’. The current direction was defined relative to the eastward direction
and positive counterclockwise.

vortices as shown in Figs. 19 and 21. Kihara and Matsuyama (2010) and this study, to 0.000035, the linear regression coefficient a1 decreased
Kihara et al. (2012) conducted numerical simulations of bathymetry from 1.1 to 0.27 (Fig. 23). The a1 value is of the simulated volume
changes in Kirinda Harbor and a harbor in a laboratory experiment changes in the eight areas on the measured ones as mentioned in 4.2.
using a three-dimensional hydrodynamic model and showed that se- When the c1 value increased from 0.00035 to 0.0017 and 0.0035, the
diment depositions inside the ports were induced by the near-bottom coefficient a1 increased from 1.1 to 2.2 and 2.1, respectively. This result
secondary flows of vortices, which moved toward the centers of the indicates that although the amount of the bathymetry change simulated
vortices. However, the influences of the vortices on the sediment de- by the model increased as the c1 value increased, the rate of increase
positions in this study and in Kihara and Matsuyama (2010) and Kihara decreased. At around a c1 value of 0.0017, the amount of bathymetry
et al. (2012) are different. Our simulation was depth-averaged, not change seems to be saturated. As shown in Fig. 24, in Area D, the
three-dimensional, and thus, it is not able to reproduce the secondary erosion rate was larger with c1 = 0.0035 than with c1 = 0.0017 from
flow associated with a vortex. The role of the vortices simulated in this the start of the tsunami-induced bathymetry change to approximately
study was to redistribute sediments circularly around the centers of the 100 min. However, from 100 min onwards, the volume change with
vortices as shown in Fig. 21. When a three-dimensional simulation is c1 = 0.0035 almost reached an equilibrium state at approximately
carried out for the bathymetry change in the study site, the sediment 200 min. The volume change with c1 = 0.0017 also reached an equi-
depositions would be more concentrated in the centers of the vortices as librium state but at approximately 500 min and both of the volume
indicated by Kihara et al. (2012). changes with c1 = 0.0035 and with c1 = 0.0017 were almost the same
The parameter c1 in Eq. (2) for sediment suspension is one of the key at the end of the simulation period. Similar bathymetry changes also
parameters for bathymetry changes as indicated by Sugawara et al. took place in Area F as shown in Fig. 24. As a result, the linear re-
(2014b) and Jaffe et al. (2016). The sensitivity test showed that when gression coefficient a1 did not change much even as the c1 value in-
the c1 value decreased from 0.00035, which was the value employed in creased from 0.0017 to 0.0035.

7
Y. Kuriyama, et al. Marine Geology 427 (2020) 106225

Fig. 16. Velocity and suspended sediment concentration at 84 min.

Fig. 18. Volume change rate in Area F (see the location in Fig. 9). Water level
(b), east-west (c, black line) and north-south (c, red line) velocities and sus-
pended sediment transport rate toward the basin (d) averaged along Line bb’
(see the location in Fig. 9). The vertical broken lines indicate the timings of the
five deposition rate peaks. (For interpretation of the references to colour in this
figure legend, the reader is referred to the web version of this article.)

were −9.1 m and −15.3 m, respectively, and those in Area E


were −8.8 m and −6.2 m, respectively. When the threat of an erosion
Fig. 17. Velocity and suspended sediment concentration at 62 min. to the stability of a port facility such as a breakwater or wharf is ex-
amined, a more accurate prediction may be required.
In this study, we used the two-way model instead of the one-way
The c1 value used in this study for the bathymetry change in a port, one. The comparison between the volume changes in Areas A to H si-
0.00035, is 1/10 the value used by Kuriyama et al. (2014) for the beach mulated by the one- and two-way models shows that the amounts of
profile change on a natural sandy beach. Both this study and Kuriyama erosion and deposition are larger in the one-way model's simulation
et al. (2014) obtained quantitative agreements between the measured (Fig. 25). The result is understandable because the sediment transport
and simulated morphological changes. When a c1 of 0.0035, as em- rate becomes smaller as the morphology approaches to the equilibrium
ployed in Kuriyama et al. (2014), was applied in the study site, the state. However, in the one-way model, such process cannot be re-
simulated volume changes in the eight areas were 2.1 times the mea- produced.
sured values mentioned above (Fig. 23). In this study, the current ve-
locity reached 8.4 m/s, whereas it was 5.3 m/s at maximum in 6. Conclusions
Kuriyama et al. (2014). The difference in the velocity may influence the
c1 value appropriate for the bathymetry change. Further investigation We simulated the bathymetry change in Oarai Port induced by the
on the applicability of the sediment suspension formula under a wide 2011 Tohoku Tsunami using an improved two-way model and model
range of conditions is required as suggested by Apotsos et al. (2011). setup. The simulated bathymetry change quantitatively agreed with the
The comparisons of the volume changes measured and simulated in measured one. The sediment deposition in the basin and the erosion at
Areas D and E indicate that the current model well reproduced the the tips of the breakwater and the groin were well reproduced, whereas
erosion at the tips of the breakwater and the groin. However, the the erosion along the breakwater was not. The simulated volume
maximum amounts of erosion measured and simulated in Area D changes in the eight areas, which were set according to the properties of

8
Y. Kuriyama, et al. Marine Geology 427 (2020) 106225

Fig. 20. Velocity and suspended sediment concentration at 178 min.

Fig. 21. Velocity and suspended sediment concentration at 182 min.

Fig. 19. Volume change rate in Area F (a), water level (b), current velocity (c),
current direction (c), suspended sediment transport rate toward the basin (d)
and vorticity in the basin (f) during 30 min before and after peaks of volume
change rate. Water level, current velocity and direction and suspended sedi-
ment transport rate were averaged along Line bb’. Black, red, blue, green, and
orange lines show the values for the first, second, third, fourth and fifth peaks,
respectively. (For interpretation of the references to colour in this figure legend,
the reader is referred to the web version of this article.)

the measured elevation changes, were 1.1 times the measured ones, and
the correlation coefficient was 0.93.
The simulation results showed that the sediment deposition in the
center of the basin took place mostly when the counterclockwise vor-
tices redistributed sediments circularly around their centers; the sedi-
ments moved into the basin through the entrance between the break-
water and the groin before the developments of the vortices. The severe
erosion at the tip of the breakwater was mainly induced by strong
southward currents.
Once a large tsunami takes place, ports and harbors may be heavily
damaged by tsunami-induced morphological changes such as sediment Fig. 22. Velocity and suspended sediment concentration at 82 min.
depositions in basins. It is recommended that sediment depositions in
ports and harbors induced by expected tsunamis such as the Nankai

9
Y. Kuriyama, et al. Marine Geology 427 (2020) 106225

Fig. 23. Relationship between parameter c1 and linear regression coefficient


a1 .
Fig. 25. Comparison between the volume changes in Areas A to H simulated by
the one- and two-way models.

with the depositions including the arrangement of dredging vessels are


discussed.

Declaration of Competing Interest

The authors declare that they have no known competing financial


interests or personal relationships that could have appeared to influ-
ence the work reported in this paper.

Acknowledgements

Fig. 24. Time series of volume changes in Area D (green liens) and F (orange The authors would like to thank Professor Edward Anthony, the
lines) simulated with c1 = 0.00035 (thin lines), c1 = 0.0017 (thick broken Editor-in-Chief, and two anonymous reviewers for their useful, in-
lines) and c1 = 0.0035 (thick solid lines). (For interpretation of the references formative and constructive comments to improve the original manu-
to colour in this figure legend, the reader is referred to the web version of this script. We also appreciate the advice and assistance of Kenya Takahashi
article.) in conducting numerical simulations.

Trough tsunami are predicted in advance and response plan to deal

Appendix A. Appendix

A.1. Submodel of sediment transport and beach profile change

The bed sediment transport rates Qb,v and Qb,a are as a function of the velocity u (Eq. (A.1)) and acceleration a (Eq. (A.2)), respectively:
Qb, v = c2 u3 (A.1)

Qb, a = c3 (a acr ) (A.2)


where c2 and c3 are free parameters, which were set at 2.78 × 10−4 (s2/m) and 9.84 × 10−4 (m s), respectively, as in Kuriyama et al. (2014). The
parameter of acr is a threshold value, which was set at 0.2 m/s2 as Hoefel and Elgar (2003) and Kuriyama (2012).
The friction velocity was estimated using Eq. (A.3), which assumes a logarithmic vertical distribution of velocity:
h Zs
u = Us
h (ln(h/ Zs ) 1) + Zs (A.3)
where h is water depth, zs is the roughness height, κ is the van Karman constant (= 0.4) and U is the depth-averaged velocity. The zs value was
assumed to be 0.0033 m, which was estimated using the relationship zs = ks/30 (ks is the bed roughness) and substituting the apparent bed
roughness of 0.10 m obtained based on the longshore current velocities at Hasaki, Japan and Duck, North Carolina, USA (Kuriyama, 2010) for ks.
The suspended sediment concentration at the bottom was estimated by using Eq. (A.4) under the assumption that the vertical distribution of
suspended sediment concentration is exponential:

10
Y. Kuriyama, et al. Marine Geology 427 (2020) 106225

ws hC
Cb = , = c0 u h
( ))
s
s 1 exp ( ws
s
h
(A.4)
where C is the depth-averaged suspended sediment concentration, εs is the sediment diffusivity coefficient and c0 is a constant (assumed to be 0.2 as
in Fujii et al. (2009)).
The depth-averaged suspended sediment concentration was obtained from the advection equation (Eq. (A.5)). The term for horizontal diffusion
was omitted from the equation as in Kuriyama et al. (2014) because horizontal diffusion was assumed to have little influence on the suspended
sediment concentration.
C (Cuh)
= +P ws Cb
t x (A.5)
The saturated suspended sediment concentration Cb,max was determined by the following equation proposed by Engelund and Fredsoe (1976) and
Soulsby (1997).
Cb, max = Cs/(1 + 1/ 3
b, max )

( cr ) µ p
6 d
b, max = 1
0.013s (A.6)

4 1/4
µ
6 d
p= 1+
cr

where cs is the maximum concentration (= 0.65) and α1 is a constant value, which was set at 2.0 as in Soulsby (1997). The variable μd represents the
dynamic friction and was assumed to be 0.5 as in Soulsby (1997). When the suspended sediment concentration at the bottom became larger than
Cb,max, no sediment suspension was assumed.

References 10.1142/S0578563410002130.
Kuriyama, Y., 2012. Process-based one-dimensional model for cyclic longshore bar evo-
lution. Coast. Eng. 62, 48–61. https://doi.org/10.1016/j.coastaleng.2011.12.001.
Gusman, A.R., Goto, T., Satake, K., Takahashi, T., Ishibe, T., 2018. Sediment transport Kuriyama, Y., 2014. Bathymetry change in channels and basins. In: PIANC (Ed.), Tsunami
modeling of multiple grain sizes for the 2011 Tohoku tsunami on a steep coastal disasters in ports due to the Great East Japan Earthquake, Report No. 122. 2014.
valley of Numanohama, northeast Japan. Mar. Geol. 405, 77–91. PIANC, pp. 78–80.
Hanzawa, H., 2017. Annual changes and relationship between distribution of hard clam Kuriyama, Y., Takahashi, K., Yanagishima, S., Tomita, T., 2014. Beach profile change at
juveniles and sand grain size in the shoreline area of Kashimanada, Ibaraki, Japan. In: Hasaki, Japan, caused by 5-m-high tsunami due to the 2011 off the Pacific coast of
Report of the Ibaraki Prefectural Fisheries Experimental Station. 46. Ibaraki Tohoku Earthquake. Mar. Geol. 355, 234–243. https://doi.org/10.1016/j.margeo.
Prefectural Government, pp. 1–7 (in Japanese). 2014.06.003.
Adityawan, M.B., Dao, N.X., Tanaka, H., Mano, A., Udo, K., 2014. Morphological changes Kuriyama, Y., Uno, Y., Honda, K., 2015. Hindcast of bathymetry change in Oarai Port,
along the Ishinomaki Coast induced by the 2011 Great East Japan Tsunami and the Japan, caused by the 2011 tsunami. In: Proceedings of the Coastal Sediments 2015.
relationship with coastal structures. Coast. Eng. J. 56 (3), 1450016. (21 pages). ASCE (14 pages).
https://doi.org/10.1142/S0578563414500168. Li, L., Qiu, Q., Huang, Z., 2012. Numerical modeling of the morphological change in Lhok
Apotsos, A., Gelfenbaum, G., Jaffe, B., 2011. Process-based modeling of tsunami in- Nga, West Banda Aceh, during the 2004 Indian Ocean tsunami: understanding tsu-
undation and sediment transport. J. Geophys. Res. 116, F01006. https://doi.org/10. nami deposits using a forward modeling method. Nat. Hazards 64 (2), 1549–1574.
1029/2010JF001797. https://doi.org/10.1007/s11069-012-0325-z.
Borrero, J.C., 2005. Field survey northern Sumatora and Banda Aceh, Indonesia after the Meilianda, E., Dohmen-Janssen, C.M., Maathuis, B.H.P., Hulscher, S.J.M.H., Mulder,
tsunami and earthquake of 26 December 2006. Seismol. Res. Lett. 76 (3), 312–320. J.P.M., 2010. Short-term morphological responses and developments of Banda Aceh
Engelund, F., Fredsoe, J., 1976. A sediment transport model for straight alluvial channels. coast, Sumatra Island, Indonesia after the tsunami on 26 December 2004. Mar. Geol.
Nord. Hydrol. 7, 293–306. 275, 96–109. https://doi.org/10.1016/j.margeo.2010.04.012.
Fujii, N., Ikeno, M., Sakakiyama, T., Matsuyama, M., Takao, M., Mukohara, T., 2009. Mori, N., Takahashi, M., The 2011 Tohoku Earthquake Tsunami Joint Survey Group,
Hydraulic experiment on flow and topography change in harbor due to tsunami and 2012. Nationwide post event survey of the 2011 Tohoku Earthquake Tsunami. Coast.
its numerical simulation. J. Coast. Eng. 56, 291–295 (in Japanese with English ab- Eng. J. 54 (1), 1–27. https://doi.org/10.1142/S0578563412500015.
stract). Nemoto, T., Matsuura, T., Nihira, A., Okayasu, A., 2009. Changes of shoreline position
Gelfenboum, G., Jaffe, B., 2003. Erosion and sedimentation from the 17 July, 1998 Papua and sand grain size at Kashimanada Coast and their effects to distribution of hard
New Guinea Tsunami. Pure Appl. Geophys. 160, 1969–1999. clam juveniles, Meretrix lamarckii. J. Fish. Eng. 46 (1), 51–64 (in Japanese with
Geospatial Information Authority of Japan, 2011. https://www.gsi.go.jp/common/ English abstract).
000062924.pdf (in Japanese). Nielsen, P., 1992. Coastal Bottom Boundary Layers and Sediment Transport. World
Goto, K., Takahashi, J., Oie, T., Imamura, F., 2011. Remarkable bathymetric change in Scientific (324p).
the nearshore zone by the 2004 Indian Ocean tsunami: Kirinda Harbor, Sri Lanka. Nishihata, T., Tajima, Y., Moriya, Y., Sekimoto, T., 2006. Topography change due to the
Geomorphology 127, 107–116. Dec 2004 Indian Ocean Tsunami –Filed and numerical study at Kirinda Port, Sri
Gusman, A.R., Tanioka, Y., Takahashi, T., 2012. Numerical experiment and a case study Lanka. In: Proceedings of 30th International Conference on Coastal Engineering.
of sediment transport simulation of the 2004 Indian Ocean tsunami in Lhok Nga, ASCE, pp. 1456–1468.
Banda Aceh, Indonesia. Earth Planets Space 64 (3), 817–827. https://doi.org/10. Ontowirjo, B., Paris, R., Mano, A., 2013. Modeling of coastal erosion and sediment de-
5047/eps.2011.10.009. position during the 2004 Indian Ocean tsunami in Lhok Nga, Sumatra, Indonesia.
Hoefel, F., Elgar, S., 2003. Wave-induced sediment transport and sandbar migration. Nat. Hazards 65 (3), 1967–1979. https://doi.org/10.1007/s11069-012-0455-3.
Science 299, 1885–1887. Paris, R., Wassmer, P., Sartohadi, J., Lavigne, F., Barthomeuf, B., Desgages, E., Grancher,
Jaffe, B., Goto, K., Sugawara, D., Gelfenbaum, G., La Selle, S., 2016. Uncertainty in tsu- D., Baumert, P., Vautier, F., Brunstein, D., Gomez, C., 2009. Tsunamis as geomorphic
nami sediment transport modeling. J. Disaster Res. 11 (4), 647–661. crisis: lessons from the December 26, 2004 tsunami in Lhok Nga, West Banda Aceh
Kihara, N., Matsuyama, M., 2010. Numerical simulations of sediment transport induced (Sumatra, Indonesia). Geomorphology 104, 59–72. https://doi.org/10.1016/j.
by the 2004 Indian Ocean Tsunami near Kirinda Port in Sri Lanka. In: Proceedings of geomorph.2008.05.040.
32nd International Conference on Coastal Engineering. ASCE (6 pages). Ranasinghe, D.P., Goto, K., Takahashi, T., Takahashi, J., Wijetunge, J.J., Nishihata, T.,
Kihara, N., Fujii, N., Matsuyama, M., 2012. Three-dimensional sediment transport pro- Imamura, F., 2013. Numerical assessment of bathymetric changes caused by the 2004
cesses on tsunami-induced topography change in a harbor. Earth Planets Space 64, Indian Ocean tsunami at Kirinda fishery harbor, Sri Lanka. Coast. Eng. 81, 67–81.
787–797. Soulsby, R.L., 1997. Dynamics of Marine Sands: A Manual for Practical Applications.
Kim, K.O., Choi, B.H., Pelinovsky, E., Jung, K.T., 2013. Three-dimensional simulation of Thomas Telford Publications, New York, USA.
2011 East Japan-off Pacific coast earthquake tsunami induced vortex flows in the Sugawara, D., 2018. Evolution of numerical modeling as a tool for predicting tsunami-
Oarai port. J. Coast. Res. 65, 284–289. https://doi.org/10.2112/SI65-049. induced morphological changes in coastal areas: a review since the 2011 Tohoku
Kuriyama, Y., 2010. A one-dimensional parametric model for undertow and longshore Earthquake. In: Santiago-Fandino, V. (Ed.), The 2011 Japan Earthquake and
current velocities on barred beaches. Coast. Eng. J. 51 (2), 133–155. https://doi.org/ Tsunami: Reconstruction and Restoration. Advances in Natural and Technological

11
Y. Kuriyama, et al. Marine Geology 427 (2020) 106225

Hazards Research 47. 2018. Springer International Publishing AG, pp. 451–467. interpretations of time series satellite images and helicopter-borne video footage.
https://doi.org/10.1007/978-3-319-58691-5_26. Sediment. Geol. 282, 151–174.
Sugawara, D., Takahashi, T., 2013. Numerical simulation of coastal sediment transport by Tomita, T., Honda, K., 2007. Tsunami estimation including effect of coastal structures and
the 2011 Tohoku-oki earthquake tsunami. In: Kontar, Y.A., Santiago-Fandino, V., buildings by 3d model. In: Proceedings of Coastal Structures 2007. ASCE, pp.
Takahashi, T. (Eds.), Tsunami Events and Lessons Learned, Advances in Natural and 681–982.
Technological Hazards Research. 2013. Springer, pp. 99–112. Tomita, T., Honda, K., Kakinuma, T., 2006. Application of three-dimensional tsunami
Sugawara, D., Goto, K., Jaffe, B., 2014a. Numerical models of tsunami sediment transport simulator to estimation of tsunami behavior around structures. In: Proceedings of
- current understanding and future directions. Mar. Geol. 352, 295–320. 30th International Conference on Coastal Engineering. ASCE, pp. 1677–1688.
Sugawara, D., Takahashi, T., Imamura, F., 2014b. Sediment transport due to the 2011 Udo, K., Sugawara, D., Tanaka, H., Imai, K., Mano, A., 2012. Impact of the 2011 Tohoku
Tohoku-oki tsunami at Sendai: results from numerical modeling. Mar. Geol. 358, Earthquake and Tsunami on beach morphology along the northern Sendai coast.
18–37. Coast. Eng. J. 54 (1), 1250009. https://doi.org/10.1142/S057856341250009X.
Takagawa, T., Tomita, T., 2012. Effects of rupture processes in an inverse analysis on the Umitsu, M., Tanavud, C., Patanakanog, B., 2007. Effects of landforms on tsunami flow in
tsunami source of the 2011 off the Pacific Coast of Tohoku Earthquake. In: the plains of Banda Aceh, Indonesia, and Nam Khem, Thailand. Mar. Geol. 242,
Proceedings of 22nd International Offshore and Polar Engineering Conference, 141–153. https://doi.org/10.1016/j.margeo.2006.10.030.
International Society of Offshore and Polar Engineering, pp. 14–19. Wilson, R., Davenport, C., Jaffe, B., 2012. Sediment scour and deposition within harbors
Takahashi, T., Shuto, N., Imamura, F., Asai, D., 2000. Modeling sediment transport due to in California (USA), caused by the March 11, 2011 Tohoku-oki tsunami. Sediment.
tsunamis with exchange rate between bed load layer and suspended load layer. In: Geol. 282, 228–240. https://doi.org/10.1016/j.sedgeo.2012.06.001.
Proceedings of 27th International Conference on Coastal Engineering. ASCE, pp. Yamashita, K., Sugawara, D., Takahashi, T., Imamura, F., Saito, Y., Imato, Y., Kai, T.,
1508–1519. Uehara, H., 2016. Numerical simulations of large-scale sediment transport caused by
Tanaka, H., Tinh, N.X., Umeda, M., Hirao, R., Pradjoro, E., Mano, A., Udo, K., 2012. the 2011 Tohoku Earthquake Tsunami in Hirota Bay, Southern Sanriku Coast. Coast.
Coastal and estuarine morphology changes induced by the 2011 great East Japan Eng. J. 58 (4), 1640015. (28 pages). https://doi.org/10.1142/S0578563416400155.
Earthquake Tsunami. Coast. Eng. J. 54 (1), 1250010. https://doi.org/10.1142/ Yazawa, H., Uda, T., Matsuura, T., Kikuchi, I., Fukumoto, T., Kumada, T., 2009. Trapping
S0578563412500106. of sediment supplied from Naka River by Oharai Port breakwater and sedimentation
Tappin, D.R., Evans, H.M., Jordan, C.J., Richmond, B., Sugawara, D., Goto, K., 2012. inside its wave-shelter zone. J. Jpn Soc. Civ. Eng. B2-65 (1), 566–570 (in Japanese
Coastal changes in the Sendai area from the impact of the 2011 Tohoku-oki tsunami: with English abstract).

12

You might also like