You are on page 1of 23

Coastal Engineering Journal

ISSN: 2166-4250 (Print) 1793-6292 (Online) Journal homepage: http://www.tandfonline.com/loi/tcej20

Tsunami mitigation by combination of coastal


vegetation and a backward-facing step

Ghufran Ahmed Pasha, Norio Tanaka, Junji Yagisawa & Fuadi Noor Achmad

To cite this article: Ghufran Ahmed Pasha, Norio Tanaka, Junji Yagisawa & Fuadi Noor Achmad
(2018): Tsunami mitigation by combination of coastal vegetation and a backward-facing step,
Coastal Engineering Journal, DOI: 10.1080/21664250.2018.1437014

To link to this article: https://doi.org/10.1080/21664250.2018.1437014

Published online: 27 Feb 2018.

Submit your article to this journal

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=tcej20
COASTAL ENGINEERING JOURNAL, 2018
https://doi.org/10.1080/21664250.2018.1437014

ORIGINAL RESEARCH PAPER

Tsunami mitigation by combination of coastal vegetation and a


backward-facing step
Ghufran Ahmed Pashaa,b, Norio Tanakaa,c, Junji Yagisawaa,c and Fuadi Noor Achmada
a
Graduate School of Science and Engineering, Saitama University, Saitama, Japan; bDepartment of Civil Engineering, University of
Engineering and Technology, Taxila, Pakistan; cInternational Institute for Resilient Society, Saitama University, Saitama, Japan

ABSTRACT ARTICLE HISTORY


Since the 2011 Great East Japan tsunami, many improvements have been made in both hard Received 13 December 2017
and soft solutions for tsunami mitigation. After the 2011 Great East Japan tsunami, a post- Accepted 15 January 2018
tsunami survey was conducted along the coast in Miyagi Prefecture, which was one of the KEYWORDS
most tsunami-affected sites because rapid acceleration of the tsunami currents broke and Tsunami disaster mitigation;
washed away the trees, resulting in extensive damage to inland houses. In contrast, some of coastal vegetation; dropping
the houses located inland and away from vegetation with a dropping step survived. This step; backwater rise; energy
shows a possibility that a step combined with the vegetation offers greater tsunami energy reduction
reduction by providing additional resistance. Laboratory experiments were conducted to
investigate the energy reduction through a compound defense system (vegetation and a
backward-facing step). Vegetation with a step (VS) showed greater energy reduction com-
pared to that of only vegetation without a step (OV) due to additional loss by collision with
the bed surface. However, the relative energy reduction in OV remained almost constant with
the increase in the initial Froude number (F0, where the Froude number is obtained from a
model without vegetation in a flume), whereas the relative energy reduction in VS showed a
decreasing trend with increasing F0 because the energy reduction due to collision decreases
with increase in water depth or F0.

1. Introduction capacity of the artificial structure. Presently, from both


economic and environmental points of view, natural
The 2011 Great East Japan tsunami opened new chal-
methods are considered effective to mitigate tsunami
lenges for researchers to find better ways to prevent
damage. Based on previous research (Shuto, 1987;
tsunami disasters across the globe. Despite the best-
Tanaka et al. 2007), the roles of coastal vegetation in
known tsunami mitigation measures, the 2011 tsunami
tsunami mitigation include trapping, and providing a
had a devastating nationwide impact in Japan (Mori
soft-landing place, an escape route and energy dissipa-
and Takahashi 2012; Ishigaki et al. 2013). Post-tsunami
tion. Many previous studies documented the effective-
surveys showed that the major damage occurred in the
ness of coastal vegetation in mitigation of tsunamis in
Tohoku and Kanto districts of Japan (Suppasri et al.
post-disaster surveys after the 1998 Papua New Guinea
2012; Mikami et al. 2012), where the tsunami destroyed
tsunami (Dengler and Preuss 2003), 2004 Indian Ocean
parts of sea walls (tsunami gates, large embankments)
tsunami (Danielsen et al. 2005; Mascarenhas and
(Tappin et al. 2012; Suppasri et al. 2012) and coastal
Jayakumar, 2008; Tanaka et al. 2007), and 2011 Great
forests (Tanaka, Yagisawa, and Yasuda 2013) and
East Japan tsunami (Nandasena, Sasaki, and Tanaka,
caused catastrophic damage to buildings and people
2012; Tanaka et al. 2014). In fact, field plantation has
(Fraser et al. 2013; Ishigaki et al. 2013).
also been done along the coast of southern Asian
For tsunami mitigation, both artificial methods (hard
countries as a bioshield against a tsunami (Tanaka,
solutions: construction of sea walls and embankments,
2009; Tanaka et al. 2011).
installment of tsunami gates, and erection of break-
Attempts to evaluate the effectiveness of a coastal
water structures) and natural methods (soft solutions:
forest were also conducted in laboratory experiments
coastal vegetation, sand dunes, and coral reefs) can be
(Harada and Imamura, 2001; Iimura and Tanaka, 2012;
implemented (Tanaka 2009). However, artificial meth-
Pasha and Tanaka, 2016c, 2017), and numerical simu-
ods could prove to be costly for developing countries
lations were made by changing the tsunami and for-
because they require enormous capital investment
est characteristics (Hiraishi and Harada, 2003; Harada
(Tanaka 2009) and have greater chances of failure if a
and Imamura, 2005; Nandasena, Tanaka, and
large tsunami arrives that exceeds the designed

CONTACT Norio Tanaka tanaka01@mail.saitama-u.ac.jp

© 2018 Japan Society of Civil Engineers.


2 G. A. PASHA ET AL.

Tanimoto, 2008; Iimura and Tanaka 2013). The effec- design an effective combination to mitigate the
tiveness of coastal forests against tsunamis has also damage to residential areas by the next large tsunami
been investigated by Kathiresan and Rajendran that overtops the sea embankment.
(2005), Irtem et al. (2009), Yanagisawa et al. (2009), Recently, Pasha and Tanaka (2016c, 2017) con-
and Nateghi et al. (2016). On the other hand, limita- ducted laboratory experiments to clarify the energy
tions of the tsunami mitigation capacity of a coastal reduction through vegetation mounted on a uniform
forest were also discussed in terms of destruction of bed and confirmed an increase in energy reduction
the coastal forest itself (Tanaka et al. 2007), produc- with increasing vegetation density. The objective of
tion of driftwood (Dengler and Preuss, 2003; Cochard the current study was to clarify the effectiveness of a
et al. 2008), deficiencies in coastal forests such as compound defense system composed of vegetation
open gaps (Mascarenhas and Jayakumar, 2008; Thuy and a backward-facing step. To fulfill these objectives,
et al. 2009; Nandasena, Sasaki, and Tanaka 2012), the a post-tsunami field investigation survey was con-
breaking moment of the trees (Tanaka, Yagisawa, and ducted along the coastal zone of Miyagi Prefecture,
Yasuda 2013), and the magnitude of the tsunami Japan. Flume experiments were also performed to
(Tanaka 2009). In the 2011 Japan tsunami, although investigate the flow structure and the amount of
the coastal forest was destroyed, its role as a bioshield energy lost through emergent vegetation of variable
was not negligible compared to that of a sea embank- thickness and changing density, mounted over a step
ment in the mitigation of fluid force (Tanaka et al. of constant height against different flow conditions.
2014) and in protecting houses by trapping floating On the basis of field survey and experiment results, a
debris (Tanaka 2012; Pasha and Tanaka, 2016a, 2016b; tsunami mitigation plan for areas near port is also
Tanaka and Onai 2017; Tanaka and Ogino 2017). proposed. This study will help to design multiple
According to the Ministry of Land, Infrastructure, protective measures for tsunami mitigation in the
Transport and Tourism of Japan, the mitigation future.
method of a “line defense” has been changed to
“compound defenses” for a level 2 tsunami (the return
period of which is within hundreds to a thousand 2. Materials and methods
years). It was also observed after the 2011 Great East 2.1. Post-tsunami field survey
Japan tsunami that the coastal forests without
embankments suffered greater damage because they 2.1.1. Information on investigated sites
had received the tsunami fluid force directly (Tanaka, Large structures (houses and buildings) and sudden
Yagisawa, and Yasuda 2013), and forests in front of drops in elevation dissipate considerable amounts of
steep slopes received a large fluid force due to reflec- tsunami energy. This was confirmed by the post-tsu-
tion from steep hills in the 2010 Mentawai tsunami nami field survey carried out in April and May 2011
(Huang et al. 2013). In contrast, a compound defense along the tsunami-affected coast in Miyagi Prefecture
arrangement with a coastal forest and an embank- after the 2011 Great East Japan tsunami. Figure 1(a)
ment helped to reduce energy more than a forest- shows representative site locations of the field inves-
only arrangement (Igarashi and Tanaka 2016) and was tigation covering an area of 22 sq. km marked by 15
ecologically beneficial as well. On the other hand, a lines, as shown in Figure 1(b). Both conditions, i.e.
coastal forest behind an embankment also had the before and after the tsunami, are shown in Figure 1
disadvantage of producing a large amount of drift- (b). The representative vegetation comprised mainly
wood due to erosion of the substrate, which Japanese red pine (Pinus densiflora) and Japanese
destroyed much vegetation (Tappin et al. 2012; black pine (Pinus thunbergii). These species have
Tanaka and Onai, 2017). Moreover, a coastal forest been used in the development of coastal forests in
behind an embankment cannot be expected to Japan to suppress sand movement and mitigate
decrease the overtopping flow from the embank- strong winds for many decades. Table 1 shows details
ment. Thus, there is scope for further research on of the investigation sites including characteristics of
optimizing a compound defense system. Kanai and the vegetation and houses. The locations of damaged
Tanaka (2016) revealed that the damage to houses houses relative to the coastline, embankment, coastal
behind a coastal forest that was followed by a sudden vegetation, canal, and a step are shown in Figure 2(a).
drop or step was less than that behind a forest with-
out a step. Niimi et al. (2013) studied the effects of the 2.1.2. Measurement method
Teizan Canal on tsunami mitigation numerically and To analyze the effects of a coastal forest and a back-
revealed time delays in the tsunami run-up and a ward-facing step, the tsunami water depth and
decrease of inundation depth. In addition, canals can damage to the sea embankment or sea wall, trees,
help to reduce the tsunami wave velocity and energy and houses were investigated during a field survey.
(Rahman et al. 2016, 2017). Thus, it is very important The vegetation characteristics were measured near
to combine a coastal forest with a dropping step and the coast (sea side) and also away from the coast
COASTAL ENGINEERING JOURNAL 3

Figure 1. Details of field investigation sites: (a) location of investigation sites and (b) plan view of investigated sites before and
after the 2011 Great East Japan tsunami (images taken from Google Earth): (i) Ishinomaki City (lines 1–2), (ii, iii) Sendai City (lines
3–8), (iv) Iwanuma City (lines 9–10), (v) Watari Town (line 11), (vi) Yamanoto Town (lines 12–14), and (vii) Souma City (line 15).

(land side). The tsunami water depth at each site was determined. Table 1 includes all the information mea-
determined by the evidence of collisions, e.g. the sured at the site.
height of scars made by debris on the tree trunks or Figure 2(b) shows the conditions in the area after the
by broken branches and water marks on the walls of 2011 Great East Japan tsunami. High acceleration of
damaged houses, marks on broken roofs, or debris tsunami currents destroyed various parts of the sea
located on roofs. At each site, the tree-trunk diameter embankment, trees, and houses at many locations.
at breast height (m), vegetation width on both sea During the field investigation, it was observed that
side and land side, vegetation density (trees/m2), some trees remained standing at various sites of survey
number of houses in each line (remaining and washed lines that included a step down, e.g. Lines 1, 2, 3, 4, 5, 7, 8,
out), average width of houses, and step height were and 15. However, just at the landward side of the step,
4 G. A. PASHA ET AL.

Table 1. Vegetation and house characteristics of investigated sites.


Tree
diameter at
breast Vegetation Number of
height dBH width W Density nv houses
(m) (m) (Nos./m2) nH (Nos./line)
Breaking situation of Average
Investigated sea embankment after Sea Land Sea Land Sea Land Before After width of Step height
line Location tsunami side side side side side side tsunami tsunami houses dH (m) (m)
Line 1 Ishinomaki City N – 0.25 – 5 – 0.10 98 63 9.5 1.8
Line 2 Ishinomaki City N – 0.25 – 20 – 0.10 93 51 11.1 1.8
Line 3 Sendai City W 0.12 0.29 140 65 0.24 0.24 1 1 11.3 1.8
Line 4 Sendai City W 0.12 0.29 300 100 0.29 0.24 41 27 11.5 1.5
Line 5 Sendai City W 0.15 0.26 270 335 0.17 0.2 27 21 12.3 1.5
Line 6 Sendai City NW 0.15 – 20 – 0.2 – 107 11 11 None
Line 7 Sendai City NW 0.15 0.36 220 165 0.2 0.2 57 19 9.8 2.8
Line 8 Sendai City NW 0.15 0.36 270 400 0.2 0.2 2 2 11.6 1.8
Line 9 Iwanuma City NW 0.2 – 50 – 0.21 – 2 2 10.4 None
Line 10 Iwanuma City NW 0.2 – 50 – 0.27 – 24 13 11.6 None
Line 11 Watari Town W 0.16 – 200 – 0.32 – 41 18 11.7 None
Line 12 Yamamoto NW 0.22 – 70 – 0.2 – 8 5 11 None
Town
Line 13 Yamamoto W 0.24 – 20 – 0.21 – 26 5 10.9 None
Town
Line 14 Yamamoto W 0.26 – 120 – 0.23 – 18 5 8.6 None
Town
Line 15 Souma City N 0.26 – 120 – 0.23 – 34 17 9.8 1.5
N: No embankment; NW: embankment was not broken; W: embankment was washed out.

enormous scouring was also observed, which resulted in V2


E ¼ hw þ α (1)
the washing out of houses and formation of scour holes 2g
(Figure 2(b)). The arrow represents the direction of flow in
where E is specific energy, hw is height above the
Figure 2(b). The depth of scour was not included in
datum, V is velocity, and α is a coefficient to account
determining the step height, and the step height was
for variations in velocity, which was considered to be
calculated with reference to ground level. However, the
1 in this study.
depth of scour holes also contributed to the energy loss
due to generation of eddies and mixing of flow because
increasing the energy head also increases the scour 2.2.1.1. Field investigation to estimate sea side
depth and length (Tanaka and Sato 2015). velocity head. Figure 3(b) shows different para-
The area near Souma Port before the earthquake meters measured during the field investigation. The
(left) and the damage after the earthquake and tsunami location of the calculated energy head is shown in
(right) are shown in Figure 3(a). Figure 3(a) shows clear Figure 2(a). The moment acting on a tree trunk at
damage to houses located inland, which was due to the ground height as proposed by Tanaka, Yagisawa,
combined effect of tsunami currents and floating debris and Yasuda (2013) (Equation (2)) and the critical bend-
(driftwood produced from coastal vegetation, trucks lost ing moment of trees as proposed by Gardiner, Peltola,
from the port area, rubble of destroyed structures). and Kellomäki (2000) (Equation (3)) were used to
However, a vegetation belt (20–50 m wide) that fol- judge tree-trunk bending. These are shown below.
lowed a sudden backward-facing step/drop of 1.5 m h1 1 h1
just landward of vegetation (Figure 3(b)) remained M¼F ¼ Cdv ρw V1 2 h1 dBH 
2 2 2
intact and trapped tsunami-borne debris. A few of the 1 3 dBH
¼ Cdv ρw Fr gh1
2
(2)
houses located inland and away from vegetation with a 2 2
σMax I
step upward survived, which is also confirmed by Mbcri ¼ ffi kD3BH (3)
Figure 3(c). Thus, there is a possibility of a greater energy R
reduction due to the added resistance to the tsunami where F is the drag force, Cdv is the drag coefficient
flow offered by a step in combination with vegetation of a tree, ρw is the density of sea water (1030 kg/m3),
itself, resulting in survival of houses to some extent. V1 is the velocity over the step (m/s), h1 is the water
depth at the coast line (sea side) on the step of
height hs (m) (Figure 3(b)), dBH and DBH (=100 dBH)
are tree-trunk diameters at breast height in meters
2.2. Parameters used in this study
and centimeters, respectively, k is a dimensional con-
2.2.1. Energy head stant [=2 for hard trunks and 3 for elastic trees
The energy head is defined as the energy per volume (Tanaka and Yagisawa 2009)], Fr (=V1/(gh1)0.5) is the
of water at any section measured with respect to the Froude number of the step, g is the gravitational
datum line (Chow 1959): constant (m/s2), R (m) is the radius of the breaking
COASTAL ENGINEERING JOURNAL 5

Figure 2. (a) Location of damaged houses relative to the coastline, embankment, coastal vegetation, canal, and a step, and (b)
representative sites of lines having a backward-facing step showing standing trees over the step and scouring conditions just
landward of the step.

section of a tree trunk, σMax is the critical bending When M was larger than Mbcri, the tree was judged
stress of the outer fiber of trees (N/m2), and I is the to be bent, dBH became dBH-cri, and the drag coeffi-
second moment of area (m4), which is a function of cient (Cdv) was changed from 1.0 to 0.2 considering
R4. If σMax is assumed to be constant, the breaking the drag coefficient of a cylinder inclined in the flow
conditions become a function of R3, thus the func- direction (Poulin and Larsen 2007; Thuy, Tanaka, and
tion of d3. Tanimoto 2012). Combining Equations 2 and 3,
6 G. A. PASHA ET AL.

Figure 3. Study site condition: (a) area near Suma Port before and after the 2011 Great East Japan tsunami (images taken from
Google Earth), (b) arrangement of vegetation with a backward facing step, and (c) survival of some houses located away from
vegetation with a step.

1 dBHcri the sea side in Equation (1) in a unit width was


cdv pw Fr2 gh31 ¼ ¼ 100kdBHcri
3
(4)
2 2 estimated from Equation (6).
ρ g 12 3
dBHcri ¼ w
Cdv Fr2 h1 2 (5) V1c 2 Fr2
400k ¼ h1 (6)
2g 2
where dBH-cri is defined as the tree diameter at breast where V1c is the critical velocity at tree-trunk bending
height when the tree was judged to be bent, which over the step (m/s).
was determined by a field investigation, and h1, which
was also measured during a field survey. From
Equation (5), CdvFr2 was found as discussed by 2.2.1.2. Field investigation to estimate land-side
Tanaka, Yagisawa, and Yasuda (2013) by using dBH-cri velocity head. According to FEMA (2008), a structural
and h1. Finally, the Fr was determined against Cd = 1 failure by overturning is mostly due to hydrodynamic
when there was no attached debris (Tanaka, forces of the incoming or returning flow. However, in
Yagisawa, and Yasuda 2013). The velocity head at the 2011 Japan tsunami, failure of wooden houses due
COASTAL ENGINEERING JOURNAL 7

to displacement of the entire house was also wide- at ground height (Equation (7)) and the critical bend-
spread. The uplift loading is mainly applied to the ing moment of washout of houses due to fluid force
underside of floor systems and is blamed for the col- (Equation (8)) as calculated by Tanaka et al. (2014) and
lapse of elevated floor levels in structures like parking Tanaka and Onai (in press) are
garages, although most failures of this type did not
result in collapse of the entire structure (FEMA 2008). h2 1 h2
M¼F ¼ ρw CdH V2 2 h2 dH
Many previous studies had developed fragility 2 2 2
1
curves for washout of buildings as a function of ¼ ρw CdH V2 2 h22 dH (7)
4
water depth and had concluded that the damage
varies according to the tsunami conditions and the 1 h2 2
Mc ¼ ρw CdH V2 2c dH (8)
strength of buildings (Reese et al. 2011; Leone et al. 2 2
2011; Suppasri et al. 2012). However, Takahashi,
where F is the drag force, h2 is the water depth in
Nakagawa, and Kanou (1985) estimated the critical
front of surviving houses (m), ρw is the density of sea
value of the moment index (u2h2, where u is depth-
water (1030 kg/m3), CdH is the average drag coeffi-
averaged velocity and h is water depth) for washing
cient of houses [2.13 for a house (Takahashi,
out houses to be around 76 m4/s2 according to a real-
Nakagawa, and Kanou 1985)], V2 is the velocity in
scale experiment. Hatori (1984) demonstrated the cri-
front of the house (m/s), dH is the average width of
tical value of the fluid force index (u2h) to be between
a house (m), and V2c is the critical velocity at wash out
15 and 100 m3/s2 in order to cause one-third to
of a house (Figure 3(b)). From Equation (8), the critical
complete destruction of houses in three previous
velocity of houses washed out by fluid force can be
Japanese tsunamis. To determine the necessary con-
written as
ditions for breakage or washing out of houses, Tanaka
et al. (2014) utilized the experience of Hatori (1984)  1=2
4Mc
and Takahashi, Nakagawa, and Kanou (1985) of fluid V2c ¼ (9)
force index (u2h) and moment index (u2h2), respec- CdH ρw dH h2 2
tively. Nandasena, Paris, and Tanaka (2011) assumed a On the land side of the field investigation, the velocity
boulder to be a rectangular prism and used the head in Equation (1) per unit width was estimated by
moment principle to calculate the minimum flow the following equation:
velocity necessary to initiate boulder transport by a
 
tsunami. The fluid force index and the moment index V2c 2 1 4Mc
depend on the Froude value. When a tsunami travels ¼ (10)
2g 2g CdH ρw dH h2 2
inland, the range of Froude number is between 0.7
and 1 (Spiske et al. 2010). For the Great East Japan where Mc is the critical moment for a house to be
tsunami, Tanaka, Yagisawa, and Yasuda (2013) esti- washed out. Takahashi, Nakagawa, and Kanou (1985)
mated the Froude number to be 0.9–0.6 at reported the critical value of the moment for washing
500–550 m distance from the shoreline, and a simula- out a house (=418,730 N m) according to a real scale
tion by Tanaka et al. (2014) also estimated the Froude experiment. This value was 11% greater than the Mc
number to decrease from 1.6 near the shoreline to 0.6 value obtained by the fragility curve with a trapping
around 550 m from the shoreline in the Sendai Plain. effect developed by Tanaka and Onai (2016). The
Tanaka et al. (2014) showed the relationship between washout rate of houses observed by Tanaka et al.
Froude number and critical water depth for washing (2014) was around half of that of Takahashi,
out houses to be based on fluid force index and Nakagawa, and Kanou (1985) because Tanaka et al.
moment index. The study expressed that for flows (2014) did not consider the impact force and debris
with a Froude number less than 0.8, the moment collision force which are assumed to occur in an
index is more suitable than the fluid force index to actual tsunami. Thus, the critical value of the moment
evaluate the washout condition of houses in an inland for washing out a house as reported by Takahashi,
region. Beyond Froude number = 0.8, the difference Nakagawa, and Kanou (1985) was used in this study.
between critical depths calculated by fluid force index Using the critical velocity V2c obtained from
and moment index was very much less (Tanaka et al. Equation (10) and water depth h2, the values of fluid
2014). In the present study, as shown in Figure 2(a), force and moment indexes were calculated. The prob-
the calculated location of the land-side energy head ability of damage of houses based on the moment
for most of the lines ranged between 600 and 2000 m index was around 80% of the fragility curve devel-
from the coastline; therefore, the range of Froude oped by Tanaka, Onai, and Kondo (2015) and Tanaka
number is assumed to be less than 1. Thus, the and Onai (2016), while it was around 60–70% based
moment index is more suitable than the fluid force on fluid force index (Tanaka, Onai, and Kondo 2015).
index to evaluate the washout condition of houses. Thus, in this study, the moment index to calculate the
The moment acting on a house due to the drag force failure mode of houses was used.
8 G. A. PASHA ET AL.

2.2.1.3. Experimental velocity head. To find the where Cdv is the drag coefficient of a tree [0.2 for a
experimental loss of the energy head, the velocity bending tree (Thuy, Tanaka, and Tanimoto 2012)], dBH
head in Equation (1) in a unit width was estimated is the diameter of a tree at breast height in meters
by Equation (11). (m), nv is the density of trees (number/m2), CdH is the
 2  drag coefficient of house [2.13 for a house (Takahashi,
V3 2 1 Q Nakagawa, and Kanou 1985)], dH is the average width
¼ (11)
2g 2g b2 y2 of houses in a single line (m), and nH is the number of
houses in one line.
where V3 (m/s) is the mean velocity calculated from
the experimental data against the known discharge
2.2.3. Relative backwater rise
(Q), measured water depth (y), and channel width (b).
Backwater rise is the rise of water in front of vegeta-
The water level was measured at the center of the
tion due to reflection and resistance by the vegetation
channel and throughout the length of the channel
(Figure 4(a)). Backwater rise was calculated using the
with an increment of 2–5 cm depending upon fluc-
following equation:
tuations of the water surface (Figure 4).
Δh ¼ h1  h0 (13)

2.2.2. Index of resistance where h1 is the water depth on the seaward side with
The importance of large structures (trees and houses) vegetation, i.e. on the upstream side of the vegetation
and a backward-facing step in energy reduction was model, and h0 is the water depth on the seaward side
also determined by developing a relationship without vegetation, i.e. the water depth in a model
between the index of resistance and total water channel without vegetation (Figure 4(a)).
head difference by tree/house. Both the trees and
houses offered sufficient resistance to the tsunami 2.2.4. Moment index
flow, which was calculated in terms of index of resis- To determine the possible damage to trees with the
tance (Cd dn), and defined as a “product of the drag provision of a backward facing step, the moment (M)
coefficient and thickness of trees/houses.” The index acting on a tree trunk at ground height is given by
of resistance per unit strip was calculated using the the following equation:
following equation: h 1 h 1
M¼F ¼ CD ρu2 h  ¼ CD ρ  u2 h2
2 2 2 2
Index of resistance ðCd dnÞ 1
¼ Cdv dBH nv þ CdH dH nH ðNos:mÞ (12) ¼ CD ρ  IM (14)
2

Figure 4. Flow structure scheme and definition of various parameters: (a) without step and (b) with step, where h0 is the initial
water depth without vegetation model, Δh is the backwater rise, hs is the step height, F0 is the Froude number at h0 (initial
Froude number), W is the width of the vegetation model, y1 the is minimum water depth during the jump (water depth at the
toe of a hydraulic jump), y2 is the mean water depth after the hydraulic jump, HGL is the hydraulic grade line, EGL is the energy
grade line, E1 the is specific energy at vegetation front, and E2 is the mean specific energy after formation of a hydraulic jump
where the flow is subcritical.
COASTAL ENGINEERING JOURNAL 9

where F is the fluid force in unit widths, h is the water considered to be the cylinder diameter (m), v is the
depth at the step location, CD is the drag coefficient, ρ kinematic viscosity (m2/s) of fluid based on the dia-
is the density of the fluid, u is the depth-averaged meter of a single cylinder, and the mean velocity of
velocity, and IM is the moment index (=u2h2). approach flow without a vegetation model ranged
from 1300 to 2500 against different flow conditions.
The drag coefficient of a cylinder remains almost
2.3. Experimental setup and conditions constant for the range of Reynolds numbers between
2.3.1. Experimental procedure and flume 103 and 105 (Anderson 1991). Also, the viscosity and
characteristics density of water were same for every case; thus,
Laboratory experiments with nine different cases Reynolds number was ignored.
(Table 2) were conducted at Saitama University in a The Weber number W is defined as U2ρL/σ, where
glass-sided water flume that was 5 m in length, 0.7 m U is the reference velocity (m/s), ρ is the density of sea
in width, and 0.5 m in height. Figure 5(a) shows the water (1030 kg/m3), L is the reference length (m), and
detailed setup of the experimental channel from sea- σ is the surface tension (0.0728 N/m for water–air
ward to landward sides. The bed slope was set to a interface at 20°C). By considering water depth without
constant value of 1/500 before and after the step. In a vegetation model as a reference length, the Weber
the experimental channel, the ground conditions and number ranged from 47 to 359 against different flow
tsunami characteristics implicated were not specific to conditions. In the current experimental study, the
any location but were considered in general. mean velocity of approaching flow without a vegeta-
Vegetation models covering the full width of the chan- tion model against nine water depths (0.03–0.07 m)
nel, as shown in Figure 5(a), were mounted in the ranged from 0.33 to 0.61 m/s, which is greater than
water flume bed about 2.5 m from the upstream the conditions to reduce surface tension effects
inlet, followed by a downward step of 2.5 cm height. defined by Novak and Cabelka (1981). For models of
The water level was measured throughout the center intake (shallow flows), the Weber number must be
of the channel by using a rail-mounted point gauge. greater than 11 to eliminate the effects of surface
The discharge was measured using a flowmeter (Signet tension and viscosity (Novak et al. 1990). Thus, it is
8150 Flow Totalizer). The depth-averaged velocity was assumed that the Weber number does not affect the
computed using the measured water depth and dis- scaling of the measured forces. Moreover, Froude
charge. The water velocity calculated from the flow- scaling is commonly used to measure free surface
meter discharge reading was compared with the gravity flows; therefore, to set the flow conditions in
depth-averaged velocity calculated from particle the current study, Froude similarity was applied to set
image velocimetry (Laser Light Sheet: G200, high the model scale (1/100) of the physical experiment.
speed digital CCD camera: K-II, fps: 600, flow analyzing
software: FlowExpert2D2C, Katokoken Co., Ltd.). The 2.3.3. Setting initial Froude number to a steady-
difference was less than 5%. subcritical state
All experiments were conducted in steady state flow
2.3.2. Similarity conditions conditions. In case of an unsteady flow, reflection by
The Reynolds number Re = UL/v, where U is the the forest affects the flow discharge toward the inland
reference velocity (m/s), L is a reference length (downstream vegetation), and these reflection effects

Table 2. Experimental conditions.


Initial Froude number Vegetation density Distance between cylinders Vegetation width Vegetation thickness
Case no. F0 G/d D (cm) W (cm) dn (No. cm) Vegetation type
1 0.57, 0.62, 0.65, 0.66, 0.68, 0.25 1 3.88 179.21 Dense
0.69, 0.70, 0.71, 0.73
2 0.57, 0.62, 0.65, 0.66, 0.68, 1.09 1.67 10.55 174.72 Intermediate
0.69, 0.70, 0.71, 0.73
3 0.57, 0.62, 0.65, 0.66, 0.68, 2.13 2.5 24.27 179.36 Sparse
0.69, 0.70, 0.71, 0.73
4 0.57, 0.62, 0.65, 0.66, 0.68, 0.25 1 8.23 380.13 Dense
0.69, 0.70, 0.71, 0.73
5 0.57, 0.62, 0.65, 0.66, 0.68, 1.09 1.67 23.60 390.85 Intermediate
0.69, 0.70, 0.71, 0.73
6 0.57, 0.62, 0.65, 0.66, 0.68, 2.13 2.5 52.48 387.83 Sparse
0.69, 0.70, 0.71, 0.73
7 0.57, 0.62, 0.65, 0.66, 0.68, 0.25 1 12.58 581.05 Dense
0.69, 0.70, 0.71, 0.73
8 0.57, 0.62, 0.65, 0.66, 0.68, 1.09 1.67 35.2 582.96 Intermediate
0.69, 0.70, 0.71, 0.73
9 0.57, 0.62, 0.65, 0.66, 0.68, 2.13 2.5 78.52 580.27 Sparse
0.69, 0.70, 0.71, 0.73
10 G. A. PASHA ET AL.

Figure 5. Experimental setup: (a) schematic figure of channel with vegetation model, (b) detailed arrangement of vegetation
model, (c, d) vegetation model with 2.5 cm step, (e) dense vegetation model with dn = 380.13 No. cm, (f) intermediate
vegetation model with dn = 390.85 No. cm, and (g) sparse vegetation model with dn = 387.83 No. cm.

can be assumed to be larger than those of the steady Yasuda, 2013; Pasha and Tanaka, 2016a, 2016b), it is
flow case. At the point of maximum water depth, the better to set a coastal forest inland where it is
change in water depth w.r.t. time is small, and the expected to trap tsunami-borne floating debris.
reflection is limited near the forest and is also Moreover, in Souma Port (line 15), the CdvFr2 (Cd = 1,
assumed to be small. However, in the case of a steady with no attached debris; Tanaka, Yagisawa, and
flow, the water is deeper and the relative reflection is Yasuda 2013) was found to be less than 1 by the
less, so the difference between steady and unsteady field investigations; as a result, Fr was less than 1.
flows can be neglected. In addition, to understand the According to the field investigation, a larger energy
fundamental effect of flow on the energy loss due to a head occurred against a water depth-to-step height
step, it is important to study it first with a steady flow. ratio (hz/hs) of 1.0–3.5 (Figure 6(a)). Beyond hz/hs of
However, in the future, more study of unsteady flow 3.5, the increased water depth over the step reduced
conditions is needed. the impact of collision with the ground, and as a
Near the coast, the tsunami flow was supercritical result, loss of the energy head decreased.
and had a Froude number greater than 1 (Nandasena, For the above-mentioned reasons, a slightly lower
Sasaki, and Tanaka, 2012; Tanaka et al. 2014); how- range of initial Froude numbers (F0) was selected to
ever, the flow changed to a subcritical state when it keep the range of water depth-to-step height ratios
traveled inland, with the range of Froude numbers without a vegetation model between 1.20 and 2.80
between 0.7 and 1 (Spiske et al. 2010). As discussed (to represent actual investigation sites) and also due
in our previous studies (Tanaka, Yagisawa, and to the restricted channel length. This study defined
COASTAL ENGINEERING JOURNAL 11

Figure 6. (a) Relationship between overflow depth/step height (hz/hs) and loss of energy head, and (b) relationship between
index of resistance and loss of energy head.

the initial Froude number (F0) when the reference model, wooden cylinders with a diameter of 0.004 m
velocity and water depth were used without a vege- were used as tree models in a staggered arrange-
tation model in a water flume. For creating sub-critical ment. Figure 5(b) and Table 2 show details of the
conditions of an inundating tsunami, the water vegetation arrangement, where D is the distance
depths without a vegetation model (h0 in Figure 4 between cylinders and W is the width of the vegeta-
(a)) in the experiment were from 3 to 7 cm in incre- tion model.
ments of 0.5 cm, setting the initial Froude number According to Takemura and Tanaka (2007), flow
approximately equal to 0.55–0.75. Against the water structures differ depending upon the G/d arrange-
depth-to-step height ratio of 1.0–3.5, the backward- ment of the vegetation model (vegetation density),
facing step height selected for the experiment was where G represents the spacing between each cylin-
2.5 cm for a model scale of 1/100 (Figure 5(c,d)). der in a cross-stream direction and d is the diameter
of a cylinder (Figure 5(b)). The G/d ratio indicates
2.3.4. Vegetation conditions whether vegetation is sparse or dense. In order to
The tree species selected for the vegetation model investigate the energy reduction through vegetation,
were the Japanese red pine (P. densiflora) and black vegetation models with three different G/d values
pine (P. thunbergii) (average tree height = 15 m and (0.25, 1.09, and 2.13) were made. A model with a G/
trunk diameter = 0.4 m) found on the Sendai Plain; d of 0.25 represents dense vegetation, while G/d of
their tree crown was high relative to the tsunami 1.09 and 2.13 characterize intermediate and sparse
height, and the trees can be considered to be circu- vegetation, respectively (Figure 5(e–g)). In the field,
lar cylinders (Tanaka et al. 2014). For a 1/100 scale the sparse vegetation (G/d = 2.13) can be planted
12 G. A. PASHA ET AL.

using Japanese red pine (P. densiflora) and Japanese cases of sparse vegetation (G/d = 2.13) include 3, 6,
black pine (P. thunbergii) trees. However, small G/d and 9 with the three values of vegetation thickness,
values like intermediate (G/d = 1.09) and dense (G/ i.e. dn = 180, 380 and 580 No. cm.
d = 0.25) vegetation do not correspond to real forests
in the area of applicability. When the resistance by
tree crowns is included, the drag force becomes quite 3. Results and discussion
high. Tropical trees that have aerial roots, like 3.1. Importance of backward-facing step in
Rhizophora spp. and Pandanus odoratissimus, can pro- energy reduction revealed by field investigation
vide a large resistance because the density is quite
high. Young trees less than 2–3 m tall have a low Table 1 shows the number of houses that existed on
crown height and can provide a large drag force. all 15 lines of the investigation sites before and after
The crown height of the pine trees in Sendai is quite the tsunami, while the location of houses relative to
high, so we cannot expect resistance by the tree other structures such as the embankment, step, vege-
crown. Thus, when a reference diameter that includes tation, canal, and step is shown in Figure 2(a). Lines 1,
the drag effect by the crown part of trees having a 2, and 15 had no sea embankment but had a step of
low crown height and dense branches and leaves is 1.8, 1.8, and 1.5 m height, respectively. The number of
considered, the G/d value may be small. In addition, it surviving houses on these lines after the tsunami was
is possible to create conditions for a small G/d by around 50–60%. The existence of a step may be one
selecting trees that have aerial roots like Rizophora of the possible reasons for their survival. Moreover,
or P. odoratissimus in tropical regions, although tropi- the houses at lines 1 and 2 were located behind large
cal regions are limited in range. Moreover, if vegeta- structures and in a harbor area, whereas the houses
tion is combined with an artificial structure like on line 15 were located on a relatively higher ground
timber, it becomes possible to make a dense enough than surrounding houses, and this may have mini-
structure. mized the damage too.
D and W were determined under the same vegeta- Lines 3, 4, 5, 11, 13, and 14 had a sea embankment,
tion thickness dn (No. cm), which is expressed as a but it was washed out by the tsunami flow. However,
function of the summed tree diameter and is defined the percentage of houses remaining after the tsunami
as a product of the diameter of a tree (d) at breast was far higher on lines that had a step (3, 4, and 5)
height and the number of trees (n) in a rectangle with than lines that had no step (11, 13, and 14). Lines 3
a frontage of unit length along the shoreline and and 4 had one structure, which is the South Gamo
depth equal to the width of the forest (W) (Figure 5 sewage treatment plant, and the building that
(b); Shuto 1987). In this study, dn was calculated using remained is a large reinforced concrete structure
the following equation: along with the presence of a creek (a narrow water
channel) besides the Teizan Canal (Figure 2(b)). The
2
dn ¼ pffiffiffi Wd  102 (15) strength of the structure and presence of the creek
3D2 might also be reasons for the survival of the houses.
where 102 in the above equation adjusts a unit in Lines 11, 13, and 14 had no step but had a wooden
Shuto’s definition of dn (No. cm) on a 1:100 scale, structure typology, flat area, open coastline, etc. These
because d, D, and W are in centimeters. In the experi- may also be possible reasons for the damage to
ment, dn was set to three values, i.e. 180, 380, and houses along with the absence of a step. Rahman
580 (No. cm). Shuto (1987) classified the effective- et al. (2016) investigated tsunami mitigation by canals
ness of a forest and the degree of damage to the and confirmed that the tsunami lost energy due to
forest and land in terms of tsunami inundation the turbulence caused by the splashing in the canal
depth, vegetation thickness, and the presence of and the resultant reflected wave. Dao, Adityawan, and
undergrowth. Referring to his study, dn = 180 Tanaka (2013) also showed that the tsunami flow may
No. cm represents areas where some trees on weak lose energy when it enters a canal. In cases of over-
soil or at the forest fringe may be damaged and soil topping flow from the embankment, the scour depth
around the trees may be scoured, dn = 380 No. cm and length were greatly increased by an increase in
corresponds to an area where soil in the forest may the energy head at the embankment (Tanaka and
be scoured and damaged to some extent, and Sato 2015). Similar to a canal and overflow from an
dn = 580 No. cm represents an area where neither embankment, the tsunami flow loses energy when it
damage to trees nor damage to soil occurs. Cases 1, collides with the backward facing step.
4, and 7 represent cases of dense vegetation (G/
d = 0.25) with variable thickness (dn = 180, 380, 3.1.1. Impact of water depth-to-step height ratio
and 580 No. cm), while cases 2, 5, and 8 are cases on loss of energy head
of intermediate vegetation (G/d = 1.09) against Figure 6(a) shows the relationship between the water
dn = 180, 380 and 580 No. cm, respectively. The depth-to-step height ratio (hz/hs) and loss of energy
COASTAL ENGINEERING JOURNAL 13

head for the field investigation lines 1, 2, 3, 4, 5, 7, 8, at the center of a channel for vegetation plus step con-
and 15, where hz is the water depth measured in the ditions (VS), whereas Figure 8 shows a comparison of the
field immediately before the step and hs is the mea- water profiles at the center of a channel for only vegeta-
sured step height in the field (Figure 3(b)). The loss of tion (OV) and vegetation and step (VS) conditions. When
energy head was calculated by subtracting E2 (land water flows through vegetation, it faces great resistance,
side) from E1 (sea side) using Equation (1). Lines 3 and resulting in decreased water depth as it progresses. As a
4 had small hz/hs, 1.4–1.6, and the energy reduction result, a large difference in water level can be seen in VS
was due to the effect of striking the bottom, i.e. and OV arrangements, especially seaward of the vegeta-
collision. The loss due to collision was reduced in tion and inside the vegetation (Figures 7 and 8). However,
lines 1 and 2 due to a higher water depth-to-step landward of the vegetation and after the jump, the water
height ratio, i.e. hz/hs = 2.8, and thus lowering loss levels are the same for VS and OV.
of the energy head, as shown in Figure 6(a). In the VS arrangement, when the water depth-to-
step height ratio was lower, i.e. with a low initial
3.1.2. Impact of index of resistance on loss of Froude number, the stream of water landward of the
energy head step became supercritical, and during its transforma-
Both the trees and houses offered sufficient resistance tion to a subcritical state, it formed a weak undular
to the tsunami flow, which was calculated in terms of jump. Although the intensity of the undular jump was
index of resistance (Cd dn) using Equation (12). low compared to the only vegetation conditions (OV)
Figure 6(b) shows the relationship between the observed by Pasha and Tanaka (2016c, 2017) and land-
index of resistance (Cd dn) and the loss of energy ward of vegetation, the difference in water depths
head (m). The overall trend shows that increasing before and after the undular hydraulic jump was very
the Cd dn increased the energy reduction. However, small (Figure 7). Figure 8 shows that the length of an
lines 3 and 4 (marked by section A in Figure 6(b)) undular hydraulic jump increases with the provision of
indicate a large loss of energy head even with a a step. However, at higher ratios of water depth-to-step
small Cd dn due to the presence of a large structure height, no undular hydraulic jump was formed, and the
and creek. Line 3 also has a river nearby (Figure 1(b)), flow remained in a supercritical state in the channel. In
which contributed to the energy loss due to lateral such a case, most of the energy was dissipated by
mixing of water from the river and sea side. Moreover, vegetation resistance and the collision with the lower
due to a smaller water depth-to-step height ratio in elevation ground (step). Increasing the initial Froude
lines 3 and 4, greater energy reduction was observed number also increased the length of the undular
due to the collision effect (Figure 6(a)). Section B in hydraulic jump, and for most of the cases, the length
Figure 6(b) (lines 1, 2, 5, 7, 8, and 15), which has a stretched beyond the available length of the channel,
dropping structure, showed a larger loss of energy as can be seen in Figure 7.
head compared to only houses plus forest trees with-
out any dropping structure (section C, Figure 6(b)). A
3.3. Seaward side of vegetation model –
flow that is perpendicular to the coast line is assumed
backwater rise
to have more energy and produce more damage. The
damage observed perpendicular to line 11 was not 3.3.1. Effect of initial Froude number on backwater
much changed throughout the line, and the energy rise
reduction showed a similar trend to that with no large The initial flow conditions without vegetation models
step (Figure 6(b)). Therefore, the effect of lateral flow placed in the channel varied between F0 = 0.55 and
was not considered in this study. The results from the F0 = 0.75. Figure 9(a-i–iii) shows the relative backwater
field survey suggest that a dropping structure like a rise calculated using Equation (13) for different vege-
step resulted in the energy reduction and contributed tation densities [dense (G/d = 0.25), intermediate (G/
to survival of more houses. d = 1.09), and sparse (G/d = 2.13)] vegetation thick-
ness (dn = 180, 380, and 580 No. cm), with a step (VS)
and without a step (OV). The relative backwater rise is
3.2. Flow structure around vegetation model
defined as the ratio of water elevation (Δh) to initial
Increasing the density of vegetation increases the max- flow depth (h0). Figure 9(a) suggested that the relative
imum water level in front of the vegetation and the water backwater rise increased slightly with different Froude
surface slope inside the vegetation (Iimura and Tanaka conditions. The increased velocity for higher Froude
2012). However, on the landward side, an undular numbers, i.e. higher energy head for higher velocities,
hydraulic jump is produced if the water depth landward consequently increased the backwater rise. The results
of the vegetation is less than the critical depth (Pasha and shown in Figure 9(a) suggest that the backwater rise
Tanaka, 2016c, 2017). The characteristics of the undular (Δh) increased relative to the initial water depth (h0);
hydraulic jump mainly depend upon the flow and vege- thus, the difference (Δh/h0) with the increase in the
tation conditions. Figure 7 shows detailed water profiles initial Froude number is much less.
14 G. A. PASHA ET AL.

Figure 7. Distribution of water level for vegetation thickness 380 No. cm with step: (a) dense vegetation (G/d = 0.25), (b)
intermediate vegetation (G/d = 1.09), and (c) sparse vegetation (G/d = 2.13).

3.3.2. Effect of vegetation density and thickness on (Δh/h0) and vegetation densities (G/d = 0.25, 1.09,
backwater rise 2.13) for three values of vegetation thickness: (i)
Other than the Froude number, the backwater rise dn = 180, (ii) dn = 380, and (iii) dn = 580 No. cm.
greatly depends on the vegetation arrangement, i.e. Figure 9(b) includes a comparison of the relative back-
vegetation density and thickness. In the VS arrange- water rise for conditions of both vegetation with a
ment, while keeping the spacing between cylinders step and vegetation without a step. To improve the
(G/d) constant and increasing the vegetation thick- readability of Figure 9(b), the trend line and R2 value
ness (dn), the backwater rise became almost double were included for the single flow condition with
when the dn was raised from 180 to 580 No. cm F0 = 0.68.
(Figure 9(a)). Similarly, the backwater rise was also It can be seen in Figure 9(b) that the backwater rise
increased by reducing the distance between cylinders, was lower for the sparse vegetation conditions as
i.e. increasing the vegetation density while keeping compared to the dense condition for a given Froude
the vegetation thickness constant. Figure 9(b) shows number and thickness. A dense arrangement of trees
the relationship between the relative backwater rise caused a higher flow resistance than a sparse
COASTAL ENGINEERING JOURNAL 15

Figure 8. Distribution of water levels from bed for vegetation thickness 380 No. cm with step (VS) and without step (OV) (Pasha
and Tanaka, 2016c, 2017): (a) dense vegetation (G/d = 0.25), (b) intermediate vegetation (G/d = 1.09), and (c) sparse vegetation
(G/d = 2.13).

arrangement and consequently resulted in a higher 3.4. Energy reduction behind vegetation model
backwater rise. However, the relative backwater rise
Hydraulic resistance and reflection of water from trees
seaward of vegetation slightly decreased in the com-
can reduce the energy of flowing water, inundation
bined arrangement (VS) as compared to only vegeta-
depth, inundation area, and the hydraulic force
tion (OV) (Figure 9(a,b)). This is because, due to inertial
behind the vegetation. The force of the water passing
forces and the decrease of bed resistance just behind
through vegetation becomes weaker, which reduces
vegetation, the step drags the stream of water slightly
the damage behind the vegetation. As the density of
down just at the landward of vegetation, which also
vegetation increases (i.e. vegetation width becomes
affected the seaward flow depth and thus contributed
smaller), the maximum water level and maximum
to a slightly lower backwater rise in the combined
velocity behind the vegetation decrease (Iimura and
arrangement.
16 G. A. PASHA ET AL.

Figure 9. Relationship between relative backwater rise (Δh/h0) and (a) initial Froude number (F0): (i) G/d = 0.25, (ii) G/d = 1.09,
and (iii) G/d = 2.13; (b) vegetation density: (G/d) (i) dn = 180 No. cm, (ii) dn = 380 No. cm, and (iii) dn = 580 No. cm.

Tanaka 2012), resulting in significant energy reduc- and Tanaka (2012), and landward of the vegetation,
tion. Figure 7 shows a significant difference between the depth became less than the critical depth that
the water depths seaward and landward of vegeta- resulted in the formation of an undular hydraulic
tion, which resulted in a certain reduction of energy. jump. Pasha and Tanaka (2016c, 2017) found that
The energy reduction (ΔE) through a vegetation the undular hydraulic jump landward of the vegeta-
model is the difference between specific energy sea- tion (Figure 8) also contributed to energy reduction to
ward (E1) and landward of vegetation some extent. However, adding a step to vegetation in
(E2), which was calculated by using Equation (1) the current study reduced the intensity of the jump
(Figure 4). and also increased the length of the undular hydraulic
With an increase in vegetation density, the water jump even beyond the available length of the channel
level was increased in front of the vegetation by in some cases (Figure 7). Thus, it was impossible to
reflection and damming, and the water surface slope figure out how energy reduction contributed to the
inside the vegetation was increased (Iimura and jump in the combined arrangement (VS). Therefore,
Tanaka, 2012; Pasha and Tanaka, 2016a). During the only the total energy reduction (vegetation + hydrau-
experiments, it was observed that as the water flow lic jump) was calculated, i.e. ΔE = E1 − E2 where E1 is
passed through the vegetation, the depth started to the specific energy at the forest front and E2 is the
decrease, as was previously demonstrated by Iimura mean specific energy after formation of a hydraulic
COASTAL ENGINEERING JOURNAL 17

jump where the flow is subcritical (Figure 4). However, between the relative total energy reduction (ΔE/E1)
in the VS condition, the difference between water and vegetation densities (G/d = 0.25, 1.09, 2.13) for
depths before and after the jump was much less three values of vegetation thickness: (i) dn = 180,
(Figure 7), and for a higher Froude number flow, the (ii) dn = 380, and (iii) dn = 580 No. cm. Figure 10(b)
mean value of E2 was calculated with the water depth includes a comparison of relative total energy
prevailing in the last part of the channel, i.e. from 350 reductions for both conditions of vegetation with
to 480 cm from the channel start. The mean values of a step and vegetation without a step. To improve
E2 were considered due to fluctuations in the water the readability of Figure 10(b), the trend line and R2
surface after the hydraulic jump. value are included for the single flow condition
with F0 = 0.68.
3.4.1. Effects of initial Froude number, vegetation Figure 10(b) shows that in OV conditions, while
density, and thickness on total energy loss keeping the vegetation thickness constant, the total
Figure 10(a) shows the relationship between the relative energy reduction was greatest for dense
relative total energy reduction (ΔE/E1) and the initial vegetation (G/d = 0.25) and least for sparse vegetation
Froude number (F0), whereas the standard deviation (G/d = 2.13) due to higher resistance offered by the
values of all the plots are shown in Tables 3 (VS) dense vegetation arrangement. However, the relative
and 4 (OV). Figure 10(b) shows the relationship energy reduction remained almost constant when the

Figure 10. Relationship between relative total energy reduction (ΔE1/E1) for (a) initial Froude number (F0), (i) G/d = 0.25, (ii) G/
d = 1.09, and (iii) G/d = 2.13; (b) vegetation density (G/d): (i) dn = 180 No. cm, (ii) dn = 380 No. cm, and (iii) dn = 580 No. cm.
18 G. A. PASHA ET AL.

Table 3. Standard deviation values for relative energy reduction in vegetation + step (VS) (Figure 10(a)), where dn is in No. cm.
VS
Dense vegetation Intermediate vegetation Sparse vegetation
F0 OS dn = 180 dn = 380 dn = 580 dn = 180 dn = 380 dn = 580 dn = 180 dn = 380 dn = 580
0.58 0.03 0.02 0.01 0.01 0.02 0.02 0.01 0.01 0.03 0.01
0.62 0.02 0.02 0.01 0.01 0.06 0.01 0.01 0.01 0.02 0.01
0.64 0.01 0.01 0.00 0.01 0.01 0.01 0.01 0.01 0.01 0.01
0.66 0.01 0.01 0.01 0.01 0.00 0.02 0.01 0.02 0.01 0.02
0.68 0.01 0.01 0.01 0.01 0.01 0.02 0.01 0.02 0.01 0.02
0.69 0.02 0.01 0.01 0.03 0.02 0.02 0.01 0.01 0.02 0.01
0.70 0.02 0.02 0.01 0.03 0.01 0.02 0.02 0.01 0.02 0.01
0.71 0.01 0.01 0.01 0.03 0.01 0.02 0.02 0.01 0.02 0.02
0.73 0.01 0.01 0.00 0.04 0.01 0.02 0.02 0.01 0.02 0.02
OS: Only step; VS: vegetation + step.

Table 4. Standard deviation values for relative energy reduction in only vegetation (OV) (Pasha and Tanaka 2016c, 2017)
(Figure 10(a)), where dn is in No. cm.
OV
Dense vegetation Intermediate vegetation Sparse vegetation
F0 dn = 180 dn = 380 dn = 580 dn = 180 dn = 380 dn = 580 dn = 180 dn = 380 dn = 580
0.58 0.01 0.01 0.03 0.03 0.01 0.01 0.02 0.02 0.02
0.62 0.01 0.01 0.01 0.02 0.01 0.01 0.02 0.03 0.02
0.64 0.01 0.00 0.01 0.02 0.01 0.01 0.02 0.03 0.02
0.66 0.02 0.01 0.01 0.01 0.01 0.01 0.01 0.03 0.02
0.68 0.02 0.01 0.01 0.02 0.01 0.01 0.01 0.02 0.02
0.69 0.01 0.00 0.01 0.01 0.01 0.01 0.01 0.02 0.02
0.70 0.03 0.03 0.03 0.01 0.01 0.01 0.01 0.02 0.02
0.71 0.03 0.03 0.04 0.01 0.01 0.01 0.01 0.02 0.02
0.73 0.03 0.04 0.04 0.01 0.02 0.01 0.01 0.02 0.01
OV: Only vegetation.

initial Froude number was increased from 0.55 to 0.75 However, the effect of vegetation thickness
(Figure 10(a)). Against a constant vegetation density, became very small in the VS arrangement, and the
the relative energy reduction was slightly increased energy was reduced only 3–5% when dn was
by increasing the thickness of the vegetation (in the increased from 180 to 380 compared to an energy
OV arrangement). Wider vegetation produced a larger reduction of 8–10% in the OV arrangement.
resistance and reflection of water and thus resulted in Conversely, the total energy reduction through the
higher energy reduction to some extent. VS arrangement was 13–18%, 15–20%, and 20–25%
With a no-vegetation arrangement, i.e. only step for dense, intermediate, and sparse vegetation,
condition (OS), the energy reduction was solely respectively, higher than in the OV arrangement
due to the collision with the bed surface that (Figure 10(a,b)). Although among the three spacing
occurred at the step, i.e. at the point of change conditions, the dense VS arrangement dissipated
in elevation. This energy reduction started to greater energy as compared to intermediate and
decrease with the increase in water depth over a sparse VS arrangements (Figure 10(b)), the above
step of a constant height (points of OS in Figure 10 results suggest that the loss due to collision was
(a)). Since the energy reduction due to a collision greater in the sparse VS arrangement because the
starts to decrease with an increase in water depth resistance offered by trees in sparse vegetation was
or with an increase in the initial Froude number, smaller compared to that by dense vegetation and
the total relative energy reduction in the VS the water flowed with high velocity through vegeta-
arrangement also showed a decreasing trend with tion and collided with the lower elevation ground
an increase in the Froude number. However, the with greater impact. The results showed that a com-
experimental results of total energy reduction in a bination of more closely spaced vegetation (dense/
VS arrangement were less than a summation of the intermediate) and a step can be a more favorable
separate energy reductions due to OV and OS arrangement for planning tsunami mitigation in the
arrangements. This is due to the fact that the loss future.
due to collision in a VS arrangement was less than
that in OS, because the water depth at the step
location was higher in the VS arrangement than in 3.4.2. Possible damage to trees
the OS arrangement. Therefore, an additional coef- Damage to trees is greatly dependent on tsunami
ficient needs to be added in numerical calculations conditions, and the diameter of trees and root ancho-
to find the combined energy reductions. rage greatly affect the possible damage (Tanaka,
COASTAL ENGINEERING JOURNAL 19

Yagisawa, and Yasuda 2013). However, the possible prevented by protecting the earth using a suitable
damage to trees with the provision of a backward- energy dissipation structure or concrete protection,
facing step can be judged by knowing the moment as proposed in Figure 11. Another disadvantage of a
acting on a tree trunk at ground height (Equation step is that the length of the hydraulic jump landward
(14)). The moment index (IM = u2h2) is a function of of vegetation is increased, which is confirmed by
unit discharge. The water depth and velocity at a step Figures 7 (with step) and 8 (without step). Due to
location are changed in with and without step condi- the increased length of the hydraulic jump, the super-
tions, although the unit discharge remains the same. critical flow prevails for a longer length of ground and
However, the drag coefficient changes with the flow can be harmful for inland residential areas. In order to
structure. Thus, it is assumed that the step has very prevent this, a grass field following the energy dissi-
little effect on bending of trees. pation structure and a small embankment is pro-
posed, which can help reduce the velocity of water
due to additional resistance and ultimately reduce the
3.5. Tsunami mitigation proposed plan for area length of the hydraulic jump. Although the provision
near port of an embankment will reduce the energy due to
collision, having the effect of a water cushion, the
During the extensive damage of the 2011 tsunami, overall energy reduction is expected to be greater.
coastal forests played a vital role in tsunami mitigation. Tanaka and Igarashi (2016) investigated the effects
If the Froude number is relatively lower on the inland of a double embankment on energy dissipation and
side, then the trees can withstand the pressure of float- found greater energy reduction with the provision of
ing debris and can trap debris (Tanaka, Yagisawa, and a second embankment. However, a quantitative inves-
Yasuda, 2013; Pasha and Tanaka, 2016a, 2016b), and tigation of the effects of an embankment (down-
this was also confirmed at Sendai, Natori, Yotsukura stream of a step) on energy reduction is proposed
Beach, and Shinmaiko Beach during the field investiga- for future work. The combination of a forest and a
tion. Similarly, near the port area, the Froude number step is one promising option under a compound
was low and the forest can be considered as an inland defense system for dissipating tsunami energy.
forest. The Reconstruction Agency (2011) has recom- However, more study is needed to investigate the
mended taking advantage of coastal forests when scour length and area of the grass field.
reconstructing coastal areas in the future. In the
Sendai Plain, a multilayer countermeasure is being
implemented to mitigate the tsunami impact by com-
4. Conclusions
bining a seawall with a coastal forest and elevated land
or roadways, and Iwanuma City in Miyagi Prefecture is Field investigations and experimental studies were
one example of this strategy (Suppasri et al. 2016). conducted to optimize a compound defense system
Likewise, to design and establish a vegetation bioshield composed of vegetation and a backward-facing step
that dissipates the energy of the flow inland and espe- for tsunami mitigation. Experiments were designed
cially in the port area, the findings of the current with variations in inflow, vegetation density, and
research can be utilized. Figure 11 shows a proposed thickness against a constant step height. The follow-
plan for tsunami mitigation by a combination of vege- ing conclusions were drawn from the current study:
tation and a step.
However, adding a step to only vegetation has a (1) The results of the field investigation showed that
few disadvantages as well. The biggest disadvantage the survey lines which had a dropping structure
is local scouring just after the step as a result of the like a step behind vegetation resulted in greater
water collision with impact. Excessive scouring can energy loss from tsunami flow. The energy loss
also undermine the roots of vegetation, which can occurred due to vegetation resistance, collision
lead to possible failure. However, scouring can be with bed at step location, and by the formation

Figure 11. Proposed plan for tsunami mitigation at a port area by a combination of vegetation and a step.
20 G. A. PASHA ET AL.

of hydraulic jump at downstream of vegetation Tsunami in Aceh and Southern Thailand: A Review on
and step. It was also seen that with the increase Coastal Ecosystems, Wave Hazards and Vulnerability,
Perspectives in Plant Ecology.” Evolution Systems 10: 3–40.
in the water depth-to-step height ratio, the loss
Danielsen, F., M. K. Sorensen, M. F. Olwig, V. Selvam, F.
of energy head decreased. Parish, N. D. Burgess, T. Hiraishi, et al. 2005. “The Asian
(2) Seaward of the vegetation model without a step Tsunami: A Protective Role for Coastal Vegetation.”
(OV), the backwater rise increases while keeping Science 310: 643. doi:10.1126/science.1118387.
the spacing between cylinders (density) constant Dao, N. X., M. B. Adityawan, and H. Tanaka. 2013. “Sensitivity
and increasing the thickness (width) of the vege- Analysis of Shore-Parallel Canal for Tsunami Wave Energy
Reduction.” Ocean Engineering 69 (2): I_401–I_406.
tation. Similarly, the backwater rise also increases
Dengler, L., and J. Preuss. 2003. “Mitigation Lessons from the
by reducing the spacing between cylinders and July 17, 1998, Papua New Guinea Tsunami.” Pure and
keeping the vegetation thickness constant. Applied Geophysics 160: 2001–2031. doi:10.1007/s00024-
However, adding a step behind the vegetation, 003-2417-x.
i.e. the combined arrangement (VS), lowered the FEMA. 2008. “Guidelines for Design of Structures for Vertical
Evacuation from Tsunamis.” http://www.fema.gov/library/
backwater rise slightly due to inertial forces. A
viewRecord.do?id=3463.
dense arrangement of tree models produces a Fraser, S., A. Raby, A. Pomonis, K. Goda, S. C. Chian, J.
higher flow resistance than a sparse arrange- Macabuag, M. Offord, K. Saito, and P. Sammonds. 2013.
ment, which consequently resulted in a higher “Tsunami Damage to Coastal Defences and Buildings in
back water rise. the March 11th 2011 Mw9.0 Great East Japan Earthquake
(3) On the landward side, the vegetation resistance and Tsunami.” Bulletin Earthquake Engineering 11: 205–
239. doi:10.1007/s10518-012-9348-9.
and water collision with the ground just behind
Gardiner, B., H. Peltola, and S. Kellomäki. 2000. “Comparison
a step result in an energy reduction. In the OV of Two Models for Predicting the Critical Wind Speeds
arrangement, the relative energy reduction Required to Damage Coniferous Trees.” Ecological
remained almost constant when the initial Modelling 129: :1–23. doi:10.1016/S0304-3800(00)00220-9.
Froude number was increased from 0.55 to Harada, K., and F. Imamura. 2001. “Experimental Study on
the Resistance by Mangrove under the Unsteady flow.” In
0.75, while it increased when both the vegeta-
Proceedings of 1st Congress of the Asian and Pacific Coastal
tion density and thickness were increased. On Engineering Conference (APACE), Chinese Ocean
the other hand, in the VS arrangement, the Engineering Society 2: 975–984. Dalian, China, October
relative energy reduction showed a decreasing 18-21, 2001.
trend with increasing initial Froude number Harada, K., and F. Imamura. 2005. “Effects of Coastal Forest
because the loss due to collision becomes less on Tsunami Hazard Mitigation—A Preliminary
Investigation, Tsunamis: Case Studies and Recent
with the increase in the initial Froude number.
Developments.” Advances in Natural and Technological
Hazards Research 23: 279–292.
These findings are important for designing a suitable Hatori, T. 1984. “On the Damage to Houses Due to
tsunami mitigation scheme for an area located near a Tsunamis.” Bulletin Earthq Researcher Institute, University
port or inland area. In the future, more study of a greater of Tokyo 59:433–439. (in Japanese with English abstract).
Hiraishi, T., and K. Harada. 2003. “Greenbelt Tsunami
step height and higher Froude number is required in
Prevention in South-Pacific Region.” Report of the Port
order to further investigate the phenomena. Moreover, and Airport Research Institute 42 (2): 23.
the energy reduction due to collision with some suitable Huang, Z., T.-R. Wu, T.-Y. Chen, and S. Y. Sim. 2013. “A
energy dissipation structure needs to be studied. Possible Mechanism of Destruction of Coastal Trees by
Tsunamis: A Hydrodynamic Study on Effects of Coastal
Steep Hills.” Journal of Hydro-Environment Research 7:
Disclosure statement 113–123. doi:10.1016/j.jher.2012.06.004.
Igarashi, Y., and N. Tanaka. 2016. “Flume Experiments for
No potential conflict of interest was reported by the Clarifying the Effects of the Thickness of Seaside
authors. Coastal Forest and Tree-Overturning Event on the
Overflow Rate from Embankment.” Journal JSCE
Coastal Engineering (B2) 72 (2): I_319–I_324. (in
Funding Japanese with English abstract).
Iimura, K., and N. Tanaka. 2012. “Numerical Simulation
This study was funded by a JSPS Grant-in-Aid for Scientific
Estimating Effects of Tree Density Distribution in Coastal
Research: [Grant Number 15H02987].
Forest on Tsunami Mitigation.” Ocean Engineering 54:
223–232. doi:10.1016/j.oceaneng.2012.07.025.
Iimura, K., and N. Tanaka. 2013. “Dangerous Zone Formation
References
behind Finite-Length Coastal Forest for Tsunami
Anderson, J. D. 1991. Fundamentals of Aerodynamics, 228– Mitigation.” Journal of Earthquake and Tsunami 7 (4):
236. 2nd ed. New York: McGraw-Hill. 1350034. doi:10.1142/S1793431113500346.
Chow, V. T. 1959. Open Channel Hydraulics. New York: Irtem, E., N. Gedik, M. S. Kabdasli, and N. E. Yasa. 2009.
McGraw-Hill. “Coastal Forest Effects on Tsunami Run-Up Heights.”
Cochard, R., S. L. Ranamukhaarachchi, G. P. Shivakoti, O. V. Ocean Engineering 36: 313–320. doi:10.1016/j.
Shipin, P. J. Edwards, and K. T. Seeland. 2008. “The 2004 oceaneng.2008.11.007.
COASTAL ENGINEERING JOURNAL 21

Ishigaki, A., H. Higashi, T. Sakamoto, and S. Shibahara. 2013. Debris.” Journal Earthquake and Tsunami 10 (4): 1650008.
“The Great East-Japan Earthquake and Devastating doi:10.1142/S1793431116500081.
Tsunami: An Update and Lessons from the past Great Pasha, G. A., and N. Tanaka. 2016b. “Optimization of Inland
Earthquakes in Japan since 1923.” The Tohoku Journal of Forest for Trapping Tsunami Driftwood.” In Proceedings of
Experimental Medicine 229 (4): 287–299. doi:10.1620/ 20th Congress of IAHR APD Congress, 1–8. Colombo, Sri
tjem.229.287. Lanka, August 28-31 2016. Asia Pacific Division of the
Kanai, K., and N. Tanaka. 2016. “Multiple Defense Effect by International Association for Hydro Environment
Vegetation and Backward Step in Souma Port Area in Engineering & Research (IAHR APD).
Fukushima Prefecture.” In Proceedings of 71st Annual Pasha, G. A., and N. Tanaka. 2016c. “Energy Loss and Drag in
Conference (JSCE), 329–330. Sendai, Japan, Vol. (165).(in a Steady Flow through Emergent Vegetation.” In
Japanese with English abstract). September 7, 2016. Proceedings of 12th ICHE Conference, 1–6, edited by Pao-
Kathiresan, K., and N. Rajendran. 2005. “Coastal Shan Yu & Wei-Cheng Lo. Tainan, Taiwan, November 6-10
Mangrove Forests Mitigated Tsunami.” Estuarine, 2016. International Conference on Hydroscience &
Coastal and Shelf Science 65 (3): 601–606. Engineering (ICHE).
doi:10.1016/j.ecss.2005.06.022. Pasha, G. A., and N. Tanaka. 2017. “Undular Hydraulic Jump
Leone, F., F. Lavigne, R. Paris, J.-C. Denain, and F. Vinet. 2011. Formation and Energy Loss in a Flow through Emergent
“A Spatial Analysis of the December 26th, 2004 Tsunami- Vegetation of Varying Thickness and Density.” Ocean
Induced Damages: Lessons Learned for A Better Risk Engineering 141: 308–325. doi:10.1016/j.
Assessment Integrating Buildings Vulnerability.” Applied oceaneng.2017.06.049.
Geography 31 (1): 363–375. doi:10.1016/j. Poulin, S., and A. Larsen. 2007. “Drag Loading Of Circular
apgeog.2010.07.009. Cylinders Inclined In The Along-wind Direction.” Journal
Mascarenhas, A., and S. Jayakumar. 2008. “An Environmental Of Wind Engineering And Industrial Aerodynamics 95:
Perspective of the Post-Tsunami Scenario along the Coast 1350-1363. doi: 10.1016/j.jweia.2007.02.011.
of Tamil Nadu, India: Role of Sand Dunes and Forests.” Rahman, M., E. Nakaza, K. Inagaki, S. Tanaka, and C. Schaab.
Journal Environmental Management 89: 24–34. 2016. “Research on Tsunami Damage Reduction Due to
doi:10.1016/j.jenvman.2007.01.053. Canals.” Journal JSCE Coastal Engineering (B2) 72 (1): I_62–
Mikami, T., T. Shibayama, M. Esteban, and R. Matsumaru. I_70. (in Japanese with English abstract).
2012. “Field Survey of the 2011 Tohoku Earthquake and Rahman, M. M., C. Schaab, and E. Nakaza. 2017.
Tsunami in Miyagi and Fukushima Prefectures.” Coastal “Experimental and Numerical Modeling of Tsunami
Engineering Journal 52 (1): 1250011. Mitigation by Canals.” Journal Waterway, Port, Coastal,
Mori, N., and T. Takahashi. 2012. “Nationwide Post Event Ocean Engineering 143 (1): 04016012. doi:10.1061/(ASCE)
Survey and Analysis of the 2011 Tohoku Earthquake WW.1943-5460.0000355.
Tsunami.” Coastal Engineering Journal 52 (1): 1250001. Reese, S., B. A. Bradley, J. Bind, G. Smart, W. Power, and J.
Nandasena, N. A. K., N. Tanaka, and K. Tanimoto. 2008. Sturman. 2011. “Empirical Building Fragilities from
“Perspective of Coastal Vegetation Patches with Observed Damage in the 2009 South Pacific Tsunami.”
Topography Variations for Tsunami Protection in 2d- Earth-Science Reviews 107 (1–2): 156–173. doi:10.1016/j.
Numerical Modeling.” Annual Journal Hydr Engineering earscirev.2011.01.009.
JSCE 52: 133–138. doi:10.2208/prohe.52.133. Reconstruction Agency. 2011. “Basic Guidelines for
Nandasena, N. A. K., R. Paris, and N. Tanaka. 2011. Reconstruction in response to the Great East Japan
“Reassessment of Hydrodynamic Equations: Minimum Earthquake.” 47 pages. Available at: http://www.recon
flow Velocity to Initiate Boulder Transport by High struction.go.jp/english/pdf/Basic_Guidelines_for_
Energy Events (Storms, Tsunamis).” Marine Geology 281: Reconstruction.pdf.
70–84. doi:10.1016/j.margeo.2011.02.005. Shuto, N. 1987. “The Effectiveness and Limit of Tsunami Control
Nandasena, N. A. K., Y. Sasaki, and N. Tanaka. 2012. Forests.” Coastal Engineering Japanese 30 (1): 143–153.
“Modeling Field Observations of the 2011 Great East Spiske, M., R. Weiss, H. Bahlburg., J. Roskosch, and H.
Japan Tsunami: Efficacy of Artificial and Natural Amijaya. 2010. “The TsuSedMod Inversion Model
Structures on Tsunami Mitigation.” Coastal Engineering Applied to the Deposits of the 2004 Sumatra and
67: 1–13. doi:10.1016/j.coastaleng.2012.03.009. 2006 Java Tsunami and Implications for Estimating
Nateghi, R., J. D. Bricker, S. D. Guikema, and A. Bessho. 2016. flow Parameters of Palaeo-Tsunami.” Sedimentary
“Statistical Analysis of the Effectiveness of Seawalls and Geology 224: 29–37. doi:10.1016/j.sedgeo.2009.12.005.
Coastal Forests in Mitigating Tsunami Impacts in Iwate Suppasri, A., P. Latcharote, J. D. Bricker, N. Leelawat, A.
and Miyagi Prefectures.” PLoS ONE 11 (8): e0158375. Hayashi, K. Yamashita, F. Makinoshima, V. Roeber, and F.
doi:10.1371/journal.pone.0158375. Imamura. 2016. “Improvement of Tsunami
Niimi, T., K. Kawasaki, Y. Mabuchi, K. Nagayama, T. Tsuji, T. Countermeasures Based on Lessons from the 2011 Great
Oie, and K. Matsuda. 2013. “Numerical Simulation of East Japan Earthquake and Tsunami — Situation after
Tsunami Damage Reduction Due to the Teizan Canal.” Five Years.” Coastal Engineering Journal 58 (4): 1640011.
Journal JSCE Coastal Engineering (B2) 69 (2): I_211–I_215. doi:10.1142/S0578563416400118.
(in Japanese with English abstract). Suppasri, A., S. Koshimura, K. Imai, E. Mas, H. Gokon, A.
Novak, P., A.I.B. Moffat, C. Nalluri, and R. Narayanan. 1990. Muhari, and F. Imamura. 2012. “Developing Tsunami
“Hydraulic Structures.” Academic division of Unwin Fragility Curves from the Surveyed Data of the 2011
Hyman Ltd, London: 546. Great East Japan Tsunami in Sendai and Ishinomaki
Novak, P., and J. Cabelka. 1981. Models in Hydraulic Plains.” Coastal Engineering Journal 54 (1): 1250008.
Engineering - Physical Principles and Design Applications, doi:10.1142/S0578563412500088.
459. Boston: Pitman. Takahashi, T., H. Nakagawa, and S. Kanou. 1985. “Risk
Pasha, G. A., and N. Tanaka. 2016a. “Effectiveness of Finite Estimation against Washed Away of Wooden Houses
Length Inland Forest in Trapping Tsunami-Borne Wood by a Flooding.” Bulletin Disaster Prevention Research
22 G. A. PASHA ET AL.

Institute 28:455–470. (in Japanese with English Tanaka, N., and M. Sato. 2015. “Scoured Depth and Length
abstract). of Pools and Ditches Generated by Overtopping Flow
Takemura, T., and N. Tanaka. 2007. “Flow Structures and from Embankments during the 2011 Great East Japan
Drag Characteristics of a Colony–Type Emergent Tsunami.” Ocean Engineering 109: 72–82. doi:10.1016/j.
Roughness Model Mounted on a Flat Plate in Uniform oceaneng.2015.08.053.
Flow.” Fluid Dynamics Research 39: 694–710. doi:10.1016/j. Tanaka, N., S. Yasuda, K. Iimura, and J. Yagisawa. 2014.
fluiddyn.2007.06.001. “Combined Effects of Coastal Forest and Sea
Tanaka, N. 2009. “Vegetation Bioshields for Tsunami Mitigation: Embankment on Reducing the Washout Region of
Review of Effectiveness, Limitations, Construction, and Houses in the Great East Japan Tsunami.” Journal Hydro-
Sustainable Management.” Landscape and Ecological Environment Research 8: 270–280. doi:10.1016/j.
Engineering 5: 71–79. doi:10.1007/s11355-008-0058-z. jher.2013.10.001.
Tanaka, N. 2012. “Effectiveness and Limitations Of Coastal Tanaka, N., and Y. Igarashi. 2016. “Multiple Defense for
Forest In Large Tsunami: Conditions Of Japanese Pine Tsunami Inundation by Two Embankment System and
Trees On Coastal Sand Dunes In Tsunami Caused By Prevention of Oscillation by Trees on Embankment.” In
Great East Japan earthquake.”.” 68 (4): II_7-II_15. . Proceedings of 20th Congress of IAHR APD Congress, 1–
Tanaka, N., and A. Onai. 2017. “Mitigation of Destructive 8. Colombo, Sri Lanka, August 28-31 2016. Asia Pacific
Fluid Force on Buildings Due to Trapping of Floating Division of the International Association for Hydro
Debris by Coastal Forest during the Great East Japan Environment Engineering & Research (IAHR APD).
Tsunami.” Landscape Ecology Engineering 13: 131-144. Tanaka, N., Y. Sasaki, M. I. M. Mowjood, and K. B. S. N.
doi:10.1007/s11355-016-0308-4. Jinadasa. 2007. “Coastal Vegetation Structures and Their
Tanaka, N., A. Onai, and K. Kondo. 2015. “Fragility Curve of Functions in Tsunami Protection: Experience of the
Different Damage of Wooden Building Due to Tsunami Recent Indian Ocean Tsunami.” Landscape and Ecological
Based on Tsunami Fluid Force and Its Moment.” Journal Engineering 3: 33–45. doi:10.1007/s11355-006-0013-9.
JSCE Coastal Engineering (B2) 71 (1): 1–11. (in Japanese Tappin, D. R., H. M. Evans, C. J. Jordan, B. Richmond, D.
with English abstract). Sugawara, and K. Goto. 2012. “Coastal Changes in the
Tanaka, N., and J. Yagisawa. 2009. “Effects of Tree Characteristics Sendai Area from the Impact of the 2011 Tohoku-Oki
and Substrate Condition on Critical Breaking Moment of Tsunami: Interpretations of Time Series Satellite Images,
Trees Due to Heavy flooding.” Landscape and Ecological Helicopter-Borne Video Footage and Field Observations.”
Engineering 5 (1): 59–70. doi:10.1007/s11355-008-0060-5. Sedimentary Geology 282: 151–174. doi:10.1016/j.
Tanaka, N., J. Yagisawa, and S. Yasuda. 2013. “Breaking sedgeo.2012.09.011.
Pattern and Critical Breaking Condition of Japanese Pine Thuy, N. B., K. Tanimoto, N. Tanaka, K. Harada, and K. Iimura.
Trees on Coastal Sand Dunes in Huge Tsunami Caused by 2009. “Effect of Open Gap in Coastal Forest on Tsunami
Great East Japan Earthquake.” Natural Hazards 65: 423– Run-Up – Investigations by Experiment and Numerical
442. doi:10.1007/s11069-012-0373-4. Simulation.” Ocean Engineering 36: 1258–1269.
Tanaka, N., and K. Ogino. 2017. “Comparison of Reduction of doi:10.1016/j.oceaneng.2009.07.006.
Tsunami Fluid Force and Additional Force Due to Impact Thuy, N. B., N. Tanaka, and K. Tanimoto. 2012. “Tsunami
and Accumulation after Collision of Tsunami-Produced Mitigation by Coastal Vegetation considering the Effect
Driftwood from a Coastal Forest with Houses during the of Tree Breaking.” Journal Coastal Conservation 16 (1):
Great East Japan Tsunami.” Landscape and Ecological 111–121. doi:10.1007/s11852-011-0179-7.
Engineering 13: 287–304. doi:10.1007/s11355-016-0321-7. Yanagisawa, H., S. Koshimura, K. Goto, T. Miyagi, F. Imamura,
Tanaka, N., K. B. S. N. Jinadasa, M. I. M. Mowjood, and M. S. A. Ruangrassamee, and C. Tanavud. 2009. “The Reduction
M. Fasly. 2011. “Coastal Vegetation Planting Projects for Effects of Mangrove Forest on a Tsunami Based on Field
Tsunami Disaster Mitigation: Effectiveness Evaluation of Surveys at Pakarang Cape, Thailand and Numerical
New Establishments.” Landscape and Ecological Analysis.” Estuarine, Coastal and Shelf Science 81: 27–37.
Engineering 7: 127–135. doi:10.1007/s11355-010-0122-3. doi:10.1016/j.ecss.2008.10.001.

You might also like