You are on page 1of 10

Ocean Engineering 54 (2012) 223–232

Contents lists available at SciVerse ScienceDirect

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

Numerical simulation estimating effects of tree density distribution in


coastal forest on tsunami mitigation
Kosuke Iimura, Norio Tanaka n
Graduate School of Science and Engineering, Saitama University, 255 Shimo-okubo, Sakura-ku, Saitama 338-8570, Japan

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

Article history: After the 2004 Indian Ocean tsunami, several studies quantitatively investigated the effects of coastal
Received 4 April 2011 vegetation on mitigating tsunami, but the effects of the tree density distribution in a forest on the
Accepted 21 July 2012 tsunami-force reduction have not been elucidated yet. This study was conducted to investigate the
tsunami mitigation effects of vegetation by changing the density distribution in a forest model. A one-
Keywords: dimensional numerical model using Boussinesq-type equations including porosity of vegetation and its
Tsunami drag and inertia forces was developed. Laboratory experiments were carried out in a wave channel, and
Coastal vegetation vegetation was modeled simply by wooden cylinders. The applicability of the numerical model was
Tree density confirmed by good agreement with experimental results. In uniform arrangement, even with the same
Solitary wave
dn (where d ¼ diameter of a cylinder, n ¼number of cylinders in the streamwise direction per unit of
Numerical simulation
cross-stream width), the velocity behind the vegetation was reduced by about 17% as the tree density
increased about 25 times. In combined arrangements of cylinders with different densities, the velocity
behind the vegetation was reduced by about 8% as the tree density increased about 8 times. Denser
vegetation increased the reflection, and the resistance by vegetation increased because the water
surface slope in vegetated region increased.
& 2012 Elsevier Ltd. All rights reserved.

1. Introductions
Shuto (1987) defined vegetation thickness as dn, where d is the
Coastal forests have long attracted attention because they diameter of a tree trunk and n is the total number of trees in the
mitigate catastrophic coastal phenomena like tsunamis and storm streamwise direction per unit of cross-stream width, and eval-
surges. Based on previous work (Shuto, 1987; Tanaka et al., 2007), uated the effect of forest damage on tsunami mitigation in
the roles of coastal vegetation in tsunami mitigation include the relation to tsunami water depth using dn.
following: Many field observations have been conducted, particularly since
the Papua New Guinea tsunami in 1998 and Indian Ocean tsunami
(1) trapping, i.e., it stops driftwood, debris (destroyed houses, in 2004. For example, Hiraishi and Harada (2003) investigated the
etc.), and other flotsam (e.g., boats);
Papua New Guinea tsunami and proposed planting a green belt of
(2) energy dissipation, i.e., it reduces water flow velocity, flow mango or coconut trees to reduce the tsunami force. Since the
pressures, and inundation water depth;
Indian Ocean tsunami, several field observations have elucidated the
(3) a soft-landing place, i.e., it provides a means of saving lives by effects of coastal vegetation on tsunami energy reduction (Danielsen
catching persons carried off by tsunamis and enabling them
et al., 2005; Kathiresan and Rajendran, 2005; Tanaka et al., 2007;
to land in tree branches; Mascarenhas and Jayakumar, 2008). Currently, a coastal forest is
(4) an escape route, i.e., it provides a way to escape by climbing
considered to be a comprehensive strategy to mitigate the destruc-
trees from the ground or from the second floor of a tive force of tsunami events. This technique is useful for developing
building; and
countries because it involves less initial capital investment than
(5) collecting wind-blown sands and raising dunes which act as more sophisticated coastal protection measures such as artificial
natural barriers to tsunamis.
structures. In fact, several projects of plant vegetation on the coasts
as a bioshield against tsunamis have been started in South and
n
Southeast Asian countries (Tanaka et al., 2009, 2011; Tanaka, 2009).
Correspondence to: Institute for Environmental Science and Technology,
Saitama University, 255 Shimo-okubo, Sakura-ku, Saitama 338-8570, Japan.
In addition to field investigations, many studies to elucidate the
Tel./fax: þ81 48 858 3564. capacity of coastal vegetation to mitigate tsunami damage have been
E-mail address: tanaka01@mail.saitama-u.ac.jp (N. Tanaka). conducted using laboratory experiments and numerical simulations.

0029-8018/$ - see front matter & 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.oceaneng.2012.07.025
224 K. Iimura, N. Tanaka / Ocean Engineering 54 (2012) 223–232

Harada and Imamura (2005) proposed a numerical model in which dn was calculated as:
the resistance of pines was evaluated by drag force and inertia force. 2 2
The drag and inertia coefficients of pines were estimated on the basis dn ¼ pffiffiffi 2 W f d  105 þ pffiffiffi 2 W b d  105 ð1Þ
3Df 3Db
of field observations and laboratory experiments. Tanaka et al. (2007)
improved the expression of drag force considering real vertical stand where 105 in Eq. (1) adjusts a unit in Shuto’s definition of dn because
structures, and proposed depth-averaged equivalent drag coefficients D and W are in mm units, and d is in m. In the experiments, dn was
for various tropical trees on the basis of field observations in Sri set to a constant 231 in all experiments. For the model scale of 1/100,
Lanka, Thailand, and Indonesia. Tanaka et al. (2007) pointed out that the diameter d was set to 0.005 m considering the depth-averaged
Pandanus odoratissimus grown on beach sand is especially effective in projection width of Rhizophora apiculata. For the distance between
mitigating tsunami damage due to its density and complex aerial cylinders D, Cases 2 and 4 refer to the spacing of trees for 8-m and
root structure in cases of a tsunami less than 5-m high. Tanimoto 5-m high R. apiculata, respectively. D is based on these two values,
et al. (2007) investigated the effects of forest streamwise width on and five cases were set for 50, 40, 30, 20, and 10 mm. Although Case
the run-up height and velocity behind the tropical forest, and 5 is very dense, the conditions express only aerial root layers, like
confirmed that the mitigating effect of a coastal forest on a tsunami those R. apiculata has, which are submerged in water. Fig. 1(d) shows
increased with the increase of the forest width. Further, the flow the dense aerial roots of R. apiculata. In the case of mangrove forests,
velocity and tsunami force were significantly reduced by a coastal the projection width of the aerial root part was 4 to 5 times the breast
forest of Pandanus odoratissimus in proportion to the increase in height diameter of the tree, and the porosity between aerial roots was
forest width (Thuy et al., 2009, 2010). Tanaka et al. (2009) performed 0.8–0.9. The porosity of the densest experimental condition was set to
real-scale simulations combining different tree species in a forest; 0.77 to represent the above condition. Because R. apiculata is a kind of
however, only one combination was simulated in their study. More- mangrove that grows in the tidal zone, the vegetation was set to a
over, in other previous studies, few studies were performed on submerged condition. In addition, the submerged condition was
changing tree density distribution in a forest. determined in this study because the detecting element of the
Thus, the effects of the tree density distribution of a forest on the capacitance-type wave gage used always has to be submerged. The
reduction of flow velocity and tsunami force have not been elucidated run-up height, water level, and force acting on a cylinder were
yet. Iimura et al. (2010) investigated the effects of the density measured in the experiments at points G1 to G9 (Fig. 1(a)). The
distribution of vegetation simply modeled by wooden cylinders on water level was measured at points G1–G9 by a capacitance-type
mitigating a tsunami by wave-channel experiments, and confirmed wave gauge (Tokyo Keisoku Co., model HAT-30, measuring time:
the effects of the density distribution of a coastal forest on the run-up 100 s, measuring time interval: 20 Hz). G1–G6 are fixed points, G7
height. In this study, a numerical model in which resistance due to and G9 are just front of and behind the vegetation, respectively, and
vegetation is added to one-dimensional Boussinesq-type equations G8 is in the middle of the vegetation in the uniform arrangement and
was validated by the experimental results. Moreover, a numerical at the boundary of two different tree densities in the combined
simulation was performed, and the effects of the density distribution arrangements with different densities of tree models.
of a forest under the same dn condition on the water level and Force gage was measured at three points (G7–G9) by a two-
velocity behind the vegetation area are discussed. component force gage (SSK Co., model LB60-1N, measuring time:
100 s, measuring time interval: 50 Hz). The force gage was set as
shown in Fig. 1(f), and it measured the force acting on a cylinder
2. Materials and methods that was located mostly in the center of the channel. In addition,
to measure the drag force, a gap is required between a cylinder
2.1. Experimental setup and conditions and the channel bottom. The experiments were conducted with
the smallest possible gap, 0.001 m, similar to the measurement by
Fig. 1(a) shows the experimental wave channel setup. Labora- Takemura and Tanaka (2007). The run-up distance was visually
tory experiments were carried out using a wave channel 0.4 m measured from the shoreline to the run-up point. Run-up height
wide and a channel bed with two slope phases, 1:4.7 and 1:20.5. was calculated from run-up distance and the bed slope ( ¼1/20.5).
The slope condition selected did not represent the ground condi-
tion at a specific location but was considered a typical case, and 2.2. Numerical model
tsunami characteristics in general are discussed. The vegetation
was modeled by wooden cylinders with a diameter of 0.005 m set 2.2.1. Governing equations
in a staggered arrangement from 10.36 m to 11.36 m. The deep- Numerical simulations were performed using Boussinesq-type
water depth was 0.4 m, and a solitary wave with a height of equations. In the calculation of a solitary wave, the numerical
3.14 cm was generated at the start boundary. The scale of this results using the shallow water equation, which does not consider
experiment was set as 1/100, so the wave height corresponds to a the dispersibility of waves, were compared with experimental
tsunami height of approximately 3 m offshore. This wave height results. As a result, the wave shape was inclined too far forward in
was the greatest value for a long wave that can be generated with the shallow zone in which the vegetation model was arranged.
this equipment. In the experiments, the cylinders in the vegeta- Therefore, in this study, the dispersibility of waves was consid-
tion model were first arranged uniformly, and the effects of tree ered. Bed resistance, porosity of vegetation, drag, and inertia are
density on mitigating waves were investigated (see Fig. 1(b)). included in the one-dimensional Boussinesq-type equations of
Second, the vegetation model was arranged with two different Madsen and Sorensen (1992). The porosity of the vegetation is
tree densities, and the mitigation effects (see Fig. 1(c)) were given according to Nandasena et al. (2008). y is the porosity of
investigated. Fig. 1(e) and Table 1 show the details of the vegetation and is calculated as:
vegetation arrangement where D is the distance between cylin- 2
ders and W is the width of the vegetation model. Subscripts f and nt pd
y ¼ 1 ð2Þ
b mean the front and back of the vegetation model, respectively, 4
d ( ¼0.005 m in this study) is the diameter of each cylinder. Cases where nt is the vegetation density.
1 to 5 and Cases 6 to 15 were uniform and combined arrange- Because a dense vegetation layer restricts the flow pass, the
ments, respectively (Table 1). D and W were determined under porosity effect should be considered. The 2-D continuity equation
the same vegetation thickness: dn (Shuto, 1987). In this study, and momentum equations are constructed by carrying out the
K. Iimura, N. Tanaka / Ocean Engineering 54 (2012) 223–232 225

G1–G9 are measurement points

front part back part

Fig. 1. Experimental set-up: (a) wave channel condition, (b) arrangement of vegetation model in uniform arrangement, (c) arrangement of vegetation model in combined
arrangement, (d) aerial roots of Rhizophora apiculata, (e) details of vegetation arrangement and its definition, and (f) set-up of force gage.

depth integration of the 3-D continuity equation and momentum resistance per unit area, fD is the drag force per unit area, fM is the
equations including the porosity effect. In this study, 1-D equa- inertia force per unit area, and b is the correction coefficient of the
tions were considered and the governing equations are as follows: dispersion term ( ¼1/15). The bed resistance in Eq. (5), drag force
in Eq. (6), and inertia force in Eq. (7) are reformulated to be
@z @Q
y þ ¼0 ð3Þ compatible with the governing equations:
@t @x
Q 9Q 9
! tb ¼ rgm2 2
ð5Þ
@Q 1 @ Q2 @z t f f y ðh þ zÞ7=3
þ þ ygðh þ zÞ þ y b þ y D þ y M
@t y @x h þ z @x r r r
  1 Q 9Q 9
1 2 @3 Q h @h @2 Q 3
3@ z
2
2 @h @ z fD ¼ rC D dnt 2 ð6Þ
¼ bþ h þ þ ybgh þ 2ybgh ð4Þ 2
3 @t@x2 3 @x @t@x @x3 @x @x2 y ðh þ zÞ

Eq. (3) is the continuity equation and Eq. (4) is the momentum  
equation, where z is the water surface elevation, Q is the flow rate pd2 @ Q
f M ¼ rC M nt ðh þ zÞ ð7Þ
(discharge per unit width), h is the still water depth, tb is the bed 4y @t h þ z
226 K. Iimura, N. Tanaka / Ocean Engineering 54 (2012) 223–232

Table 1 zR ¼ aR cosðkxst þ eR Þ ð9Þ


Conditions of vegetation model arrangement (symbols are defined in Fig. 1(e)).
where Eq. (8) is the water level of an incident wave, Eq. (9) is the
Case no. Front part Back part Case Front part Back part water level of a reflected wave, a is the amplitude, k is the wave
no. number, s is the angular frequency, and e is the phase difference.
Df Wf Db Wb Df Wf Db Wb
Subscript I means incident wave, and subscript R means reflected
(mm) (mm) (mm) (mm) (mm) (mm) (mm) (mm)
wave. If the distance from P1 to P2 is set to DL, the water levels in
Case 1 0 0 50 1000 Case 9 20 80 30 180 P1 and P2 are written respectively as:
Case 2 0 0 40 640 Case 10 10 20 20 80
Case 3 0 0 30 360 Case 11 50 500 40 320
zP1 ¼ ðzI þ zR ÞP1 ¼ AP1 cosst þ BP1 sinst ð10Þ
Case 4 0 0 20 160 Case 12 50 500 20 80
Case 5 0 0 10 40 Case 13 40 320 20 80 zP2 ¼ ðzI þ zR ÞP2 ¼ AP2 cosst þ BP2 sinst ð11Þ
Case 6 40 320 50 500 Case 14 30 180 20 80
where coefficients of AP1, BP1, AP2, and BP2 are written as:
Case 7 20 80 50 500 Case 15 20 80 10 20
Case 8 20 80 40 320 AP1 ¼ aI cosjI þ aR cosjR ð12Þ

BP1 ¼ aI sinjI aR sinjR ð13Þ


where m is the Manning’s roughness coefficient, CD is the drag
coefficient, d is the diameter of the cylinders, nt is the vegetation AP2 ¼ aI cosðkDL þ jI Þ þaR cosðkDL þ jR Þ ð14Þ
density (number of cylinders per unit area), and CM is the mass
coefficient. Moreover, the inertia force in this study was suffi- BP2 ¼ aI sinðkDL þ jI ÞaR sinðkDLþ jR Þ ð15Þ
ciently small compared with the drag force as in previous studies
(Nandasena et al., 2008; Tanaka et al., 2009) so that it hardly jI ¼ kxP1 þ eI ð16Þ
affected the simulation results (inertia force is 10% or less of the
jR ¼ kxP2 þ eR ð17Þ
drag force because the effect of inertia force on the water level
and velocity would be less than 0.5% even if the mass coefficient where each coefficient of AP1, BP1, AP2, and BP2 is a known value
was changed from 1 to 5). Therefore, CM was fixed at 2.0. calculated by the Fourier transform. Unknown values are aI, aR, fI,
Manning’s roughness coefficient and drag coefficient were deter- and fR in Eqs. (12)–(15), so unknown values are solved by four
mined by calibration with experimental results. allied formulas, aI, aR, fI, and fR, and written as follows:
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ðAP2 AP1 coskDLBP1 sinkDLÞ2 þ ðBP2 þ AP1 sinkDLBP1 coskDLÞ2
aI ¼
2.2.2. Numerical method 29sinkDL9
A set of the above equations was solved by the finite- ð18Þ
difference method of a staggered leap-frog scheme. A semi-
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
Crank–Nicholson scheme was used for the terms of bed friction ðAP2 AP1 coskDL þ BP1 sinkDLÞ2 þ ðBP2 AP1 sinkDLBP1 coskDLÞ2
and drag force. The momentum equation was solved implicitly by aR ¼
29sinkDL9
the Thomas method. For a moving boundary treatment, a number
ð19Þ
of algorithms are necessary for the flow occurring when the water
surface elevation is high enough to flow to the neighboring dry  
AP2 þ AP1 cos kDL þ BP1 sin kDL
cells (Iwasaki and Mano, 1979). Because the still water depth h is jI ¼ tan1 ð20Þ
BP2 þAP1 sin kDL þ BP1 cos kDL
included in the dispersion term, in the land area (x412.0 m), the
 
dispersion term is removed from the momentum equation and AP2 þ AP1 cos kDLBP1 sin kDL
solved as a non-linear long wave equation. The initial conditions jR ¼ tan1 ð21Þ
BP2 þ AP1 sin kDLþ BP1 cos kDL
were given for a state without a wave in the computational
Thus, the incident and reflected waves for each frequency
domain including the wave generation zone. The full-reflective
component were solved, and the incident and reflected wave of a
condition was used for the boundary.
solitary wave were calculated using the superposition principle.
However, sin kDL is contained in Eqs. (18) and (19) as the
2.3. Separation of incident wave and reflected wave denominator. Thus, the error will be amplified because the
denominator takes a very small value in the near frequency band
Separation of incident wave and reflected wave was performed used as kDL¼ ap. Then, the ceiling value of frequency fmax is set as
referring to the method of Goda (1985). Here, an incident wave is in the following equation:
a wave that goes in the positive direction of x and a reflected wave 2pf max DL 9
is a wave which goes in the negative direction of x. The reflected kDL ¼ pffiffiffiffiffiffi r p ð22Þ
gh 10
wave is generated by reflection from both the vegetation and
slope. First, a time profile of the water level at two different points In this study, fmax is set to 0.8907 Hz.
was output to a numerical simulation, and a Fourier transform
was performed in order to separate a wave for each frequency
component. A calculation to separate incident and reflected 3. Results
waves was performed for each frequency component wave, and
incident wave height, reflected wave height, and phase angle 3.1. Calibration of Manning’s roughness and drag coefficients
were calculated. Two different points for output water level were
set as P1 (x¼3.0 m) and P2 (x ¼4.0 m) in the fixed depth area Manning’s roughness coefficient is determined by comparing
where the superposition principle of waves can be applied the run-up height of numerical results with the run-up height
because nonlinearity is not strong. Here, the water level of of experimental results without vegetation. The slope model in
incident and reflected waves in a specific frequency component the experiment is made of wood. According to Chow (1959),
wave were written as: Manning’s roughness coefficient of wooden cylinders (staves) is
0.010–0.014. In this study, Manning’s roughness coefficient was
zI ¼ aI cosðkxst þ eI Þ ð8Þ set to 0.0108 by calibration, and it was judged to be appropriate
K. Iimura, N. Tanaka / Ocean Engineering 54 (2012) 223–232 227

Table 2 became large (2.9 times at the maximum) because the water
Drag coefficient in each case (G7–G9 are measurement points). surface slope inside the vegetated zone was large. For example,
the average water surface slopes inside the vegetated zone were
(a) Uniform arrangement
0.054 and 0.602 for D/d ¼10 and 2, respectively.
G7 G8 G9

Case 1 0.71 0.94 0.77 3.2. Validation of numerical model with experimental results
Case 2 0.70 0.74 0.76
Case 3 0.66 0.79 0.94 The numerical simulation was performed using the deter-
Case 4 0.78 0.95 1.18 mined Manning’s roughness and drag coefficients, and the
Case 5 1.73 1.32 2.23
numerical simulation was validated by comparing the calculated
(b) combined arrangement run-up height with/without the porosity model and water level
with the experimental results. Fig. 2(a) and (b) shows the non-
G7 G8f G8b G9 dimensional run-up height in the uniform and combined arrange-
ments, respectively, in relation to the total vegetation width W,
Case 6 0.74 0.81 0.60 0.76
Case 7 0.84 1.39 1.02 0.85 which is given by the sum of Wf and Wb. The vertical axis means
Case 8 0.83 1.07 0.84 0.74 the non-dimensional run-up height (¼ run-up height with vege-
Case 9 0.75 1.10 0.91 0.86 tation/run-up height without vegetation). Both figures show that
Case 10 1.44 0.91 0.63 0.90
the run-up height decreased with decreasing total vegetation
Case 11 0.72 0.67 0.71 0.75
Case 12 0.71 0.69 0.99 1.37
width and increasing vegetation density. Moreover, in a combined
Case 13 0.76 0.76 0.92 1.26 arrangement, especially when the back part was dense, the run-
Case 14 0.69 0.85 1.08 1.08 up height was a little smaller than that in the uniform arrange-
Case 15 0.68 0.68 1.27 1.27 ment. Although the numerical simulation produced a slightly
larger value than the experiments, the run-up height was repro-
duced fairly well. In addition, the accuracy of the calculation was
because this value falls into Chow’s range. Using the roughness
improved when the porosity effect was included.
coefficient of 0.0108, the run-up heights of both experimental and
Fig. 3(a)–(c) shows the distribution of maximum water levels
numerical results without vegetation were 0.099 m.
in the uniform arrangement for Cases 1, 3, and 5, respectively. The
The drag coefficient is determined by calibrating the max-
vertical axis is the maximum water level. Because the water level
imum force of the numerical results with the maximum force
is the displacement from the water surface in a still water
acting on a cylinder in the experiments. Here, ‘‘force’’ in numer-
condition, the water level becomes 0 with the water surface in
ical simulation means the sum of drag and inertia forces.
the initial state, as shown in Fig. 1(a). The horizontal axis is the
Table 2(a) and (b) shows results of the drag coefficient for each
distance from the wave-making plate. The dash-dotted line shows
measurement point in a uniform case and a combined case,
the bottom of the channel. As tree density increases, the width of
respectively. The Froude number of the tsunami current at
vegetation becomes narrower because dn is fixed. As vegetation
Indonesia during the 2004 Indian Ocean tsunami was estimated
density increases, the maximum water level at the front of the
to be 0.65–1.2 (Matsutomi et al., 2006). However, mangrove
vegetation (G7) increases, and the water surface slope inside the
forests were targeted in this study, and many mangrove forests
vegetation becomes large. Moreover, in the densest condition
grow in lagoons or at the end of an inner bay. In such places, the
(Case 5), the difference due to the effect of porosity is large. With
flow velocity of a tsunami could be decreased by the effects of
the porosity effect, the maximum water level at the front face of
topography and the lagoon shape, and the Froude number could
the vegetation increases, and the value approaches the measured
be assumed to be small in many cases. Therefore, the Froude
value.
number (calculated from a numerical simulation) in this study
Fig. 3(d) and (e) shows the distribution of maximum water
was set to about 0.4–0.6 in the vegetation arrangement zone.
levels in the combined arrangement for Cases 8 and 13, respec-
In this table, because G8 is the boundary of two different tree
tively. The arrangement of cylinders inside the vegetated region
densities in the combined arrangement, force is measured in the
for Case 8 is the opposite of the arrangement of Case 13 (the front
back row of the front part of vegetation, and the front row of the
part is dense in Case 8, and the back part is dense in Case 13). In
back part of vegetation, and the drag coefficient is calculated at
both cases, as the maximum water level at the front of the denser
both points. In a numerical simulation, the drag coefficient
part of the model increased, the water surface slope inside the
between different measuring points is given by linear interpola-
denser part of the model became larger.
tion. Because flow velocity was not measured in the experiments,
Fig. 4(a) shows the correlation between run-up heights deter-
the Reynolds number in experiments is not described in detail.
mined by numerical simulations and by experiments. The numer-
However, when we use the diameter of a cylinder and the
ical simulation for run-up height can reproduce well the
maximum depth-averaged velocity for reference length and
experimental results with less than 10% error. Moreover, the
velocity, respectively, the Reynolds number is around 2000 from
error of an experimental value and a numerical value E is defined
the numerical results. The drag coefficient of a single cylinder in
by the following equation:
this Reynolds number range is known from the previous results to
be around 1.0. First, the drag coefficient is lower than 1 in the N
1 X d

sparse arrangement condition (D/d ¼10, 8 and 6). Nepf (1999) E¼ ðRNS REXP Þ2 ð23Þ
Nd 1
pointed out that the drag coefficient changes with the density of
vegetation. In the experiment of Nepf (1999), the drag coefficient where Nd is number of data, and RNS and REXP are numerical and
was 0.65–0.75 when the condition of array density (product of the experimental results of the run-up height, respectively.
density of vegetation and square of the diameter) was 0.01–0.02 The value of E for the model with or without porosity is
and the vegetation was staggered. The drag coefficients in 0.99  10  5 or 2.29  10  5, respectively. The accuracy of the
D/d¼10 and 8 of this experiment were similar to the values of calculation of the model that includes the porosity effect is better
Nepf (1999). On the other hand, in a dense arrangement (D/d ¼4 than that without the porosity effect. Fig. 4(b) shows the correla-
and 2), the force acting on a cylinder and the drag coefficient tion of maximum water levels by numerical simulations and by
228 K. Iimura, N. Tanaka / Ocean Engineering 54 (2012) 223–232

R : Run-up height with vegetation


1 1 RWOV : Run-up height without vegetation
(Cases 1 to 5) (Cases 6 to 15)
0.9 0.9

0.8
R / RWOV

0.8

R / RWOV
Numerical results (with porosity)
Numerical results (front is dense)
0.7 Numerical results (without porosity) 0.7 with porosity
Numerical results (back is dense)
Experimental results Numerical results (front is dense)
without porosity
0.6 0.6 Numerical results (back is dense)
R : Run-up height with vegetation
Experimental results (front is dense)
RWOV : Run-up height without vegetation
Experimental results (back is dense)
0.5 0.5
0 200 400 600 800 1000 0 200 400 600 800 1000
Total vegetation width : W (mm) Total vegetation width : W (mm)

Fig. 2. Comparison of difference between reductions of run-up height in experiments and numerical simulations: (a) uniform arrangement, and (b) combined
arrangement.

0.12
0.12
Maximum water level (m)

Maximum water level (m)


Experimental results Experimental results
0.1 Vegetation area 0.1 Vegetation area
Numerical results (with porosity) Numerical results (with porosity)
0.08 0.08
Numerical results (without porosity) Numerical results (without porosity)
0.06 0.06
0.04 0.04
0.02 0.02
0 0
0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14
Distance from wave-making plate : x (m) Distance from wave-making plate : x (m)

0.12 0.12
Maximum water level (m)
Maximum water level (m)

Experimental results Experimental results


0.1 Vegetation area 0.1 Vegetation area
Numerical results (with porosity) Numerical results (with porosity)
0.08 Numerical results (without porosity) 0.08
Numerical results (without porosity)
0.06 0.06
0.04 0.04
0.02 0.02
0 0
0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14
Distance from wave-making plate : x (m) Distance from wave-making plate : x (m)

0.12
Maximum water level (m)

0.1 Experimental results


Vegetation area
Numerical results (with porosity)
0.08
Numerical results (without porosity)
0.06
0.04
0.02
0
0 2 4 6 8 10 12 14
Distance from wave-making plate : x (m)

Fig. 3. Distribution of maximum water level: (a) Case 1, (b) Case 3, (c) Case 5, (d) Case 8, and (e) Case 13.

experiments. The numerical results of maximum water level can backmost row of the dense vegetation model. A wave that passes
reproduce the experimental results with less than 10% error in through the dense vegetation model is affected by the reflected
general. When the error E for water level estimation is calculated wave from the slope or the other part of the model that has sparse
as well as the run-up height, E is 4.18  10  6 or 5.60  10  6 for density in back. Fig. 5 shows the time profile of the water level at
with or without porosity, respectively. The accuracy of the G9 in Case 3. Compared with the numerical result, the experi-
maximum water level calculation is also improved by including mental result shows smaller values near the peak of the curve.
the porosity effect. The point where the error is largest is the The reproducibility of G6, which is located a few centimeters
K. Iimura, N. Tanaka / Ocean Engineering 54 (2012) 223–232 229

0.1 0.1 y=x

Numerical results of maximum water level (m)


y=x
Numerical results of run-up height (m)

0.08 0.08

0.06 0.06

Uniform
0.04 Combined (front is dense) with porosity 0.04 Uniform
Combined (back is dense) Combined (front is dense) with porosity
Combined (back is dense)
Uniform
0.02 0.02 Uniform
Combined (front is dense) without porosity
Combined (front is dense) without porosity
Combined (back is dense) Combined (back is dense)
0 0
0 0.02 0.04 0.06 0.08 0.1 0 0.02 0.04 0.06 0.08 0.1
Experimental results of run-up height (m) Experimental results of maximum water level (m)

Fig. 4. Correlation of numerical results and experimental results: (a) run-up height, and (b) maximum water level.

behind the backmost row (G9) on the land side, is comparatively


0.05 Experimental results
good, although improvement of the numerical model regarding
the backmost row of vegetation and the boundary between two Numerical results
0.04
different tree densities will be needed in the future. Therefore, the
reduction effect behind the vegetation is discussed using this 0.03
numerical model which considers the effects of porosity.
Water level (m)

0.02

4. Discussion 0.01

4.1. Energy reduction behind vegetation in uniform arrangement 0


35 40 45 50
-0.01 Time (s)
The energy reduction due to vegetation is discussed based on
the results of numerical simulations. Fig. 6(a) and (b) shows the
-0.02
maximum water level and maximum velocity behind the vegeta-
tion, respectively, in relation to the total vegetation width. Fig. 5. Time profiles of water level at G9 (Case 3).
Measurements are set to two points: x¼11.5 m, the point behind
vegetation (closed plot), and x¼13.0 m, the point on land that is (dn), changing the density produces a difference in the reduction
far from vegetation (open plot). In the uniform arrangement, as effect. The reduction of water level and velocity behind the
the tree density increases (i.e., vegetation width becomes smaller), vegetation in the dense case (case 5) increased due to increases
the maximum water level and maximum velocity behind the in the reflectance by the vegetation model and force acting on
vegetation decrease. The reduction rate is defined as a ratio of the the cylinders by increasing the water surface slope inside the
maximum water level and velocity with vegetation to that with- vegetation.
out vegetation. The rates for the reduction of the maximum water With increasing density of the model trees, the water level was
level were 0.90 in the densest case (case 5, W¼40 mm) and 0.96 increased in front of the vegetation by reflection and damming
in the sparsest case (case 1, W ¼1000 mm), so the difference is up, and the water surface slope inside the vegetation increased, as
only 0.06. However, the reduction rates of maximum velocity shown in Fig. 3. Thus, the force acting on a cylinder becomes
were 0.70 in the densest case and 0.87 in the sparsest case, so the large. Therefore, the drag coefficient is affected by the water
difference becomes 0.17 where the maximum water level and surface slope and increases as tree density increases, as shown in
maximum velocity in a case without vegetation are zmax ¼0.042 m Table 2.
(at 11.5 m), zmax ¼0.063 m (at 13.0 m), Vmax ¼0.49 m/s (at 11.5 m), In addition, in order to elucidate the effects of reflected waves
and Vmax ¼0.87 m/s (at 13.0 m). In the experiment of Thuy et al. on vegetation, the components of incident and reflected waves
(2009), the reduction rate of the maximum water level is about were separated. Fig. 7(a) shows the time profile of the water level
0.7, the reduction rate of the maximum velocity is about 0.6, and at measuring point P2 (x ¼4.0 m) for the no-vegetation case.
the reduction rate of velocity is larger than that of the water level. The continuous line shows the numerical results, and the dash-
In that case, the diameter of the cylinder was 0.005 m, and the double dotted line shows the re-compounded wave after separation
distance between the cylinders was 0.023 m (Thuy et al., 2009). of the incident and reflected waves. The water level seems to shift
Although this density condition is similar to Case 4 in our slightly because in the re-compounded wave set-up, the ceiling
experiment, a very big difference in the reduction effect is shown values of the frequency and the high frequency component were
compared with this study. This is because the vegetation width excluded. However, the reproducibility is comparatively good, and
and dn were six and five times larger, respectively, in Thuy the incident wave and reflected wave can be separated with good
et al.(2009) than in this study. Moreover, in their experiment, accuracy because the error between the numerical results and the
tree density was the only parameter, and vegetation was arranged re-compounded wave of the maximum water level of the first wave
uniformly. In this study, under the same vegetation thickness is very small (less than 5%). The dash-dotted line shows the time
230 K. Iimura, N. Tanaka / Ocean Engineering 54 (2012) 223–232

0.8

Maximum water level (m) 0.06

Maximum velocity (m / s)
0.6

0.04
0.4

0.02 0.2
Measuring point x=13.00m Measuring point x=13.00m
(Cases 1 to 5) Measuring point x=11.50m (Cases 1 to 5) Measuring point x=11.50m
0 0
0 200 400 600 800 1000 0 200 400 600 800 1000
Total vegetation width : W (mm) Total vegetation width : W (mm)

Fig. 6. Reduction effect behind vegetation model (uniform arrangement): (a) maximum water level, and (b) maximum velocity.

0.04 Numerical results 0.04 Numerical results


Incident wave Incident wave
0.03 Reflected wave 0.03 Reflected wave
Water level (m)

Water level (m)


Incident + Reflected Incident + Reflected
0.02 0.02

0.01 0.01

0 0
30 35 40 45 50 55 30 35 40 45 50 55
-0.01 -0.01
Time (s) Time (s)
-0.02 -0.02

Without vegetation
0.02
Case 1
Case 2
Case 3
Water level (m)

0.01 Case 4
Case 5

0
30 35 40 45 50 55
Time (s)
-0.01

Fig. 7. Time profile of water level for estimation of incident and reflected waves at measuring point P2 (x ¼4.0 m): (a) without vegetation model, (b) Case 3,
and (c) reflected wave in uniform arrangement.

profile of the reflected wave. Although the small reflected wave 4.2. Energy reduction behind vegetation in combined arrangement
came at t¼40 s, this is a reflected wave from the 1/4.7 slope, and a
reflected wave from 1/20.5 slope came after that. Fig. 7(b) shows Fig. 8(a) and (b) shows the maximum water level and max-
the time profile of the water level at P2 in Case 3. The calculation imum velocity behind the vegetation, respectively, in relation
to separate the incident and reflected waves also showed good to the total vegetation width in the combined arrangement. In
accuracy, because the error of the maximum water level of the the combined arrangement, as total vegetation width decreased
first wave is 5% or less in this case. Moreover, one crest of the (i.e., vegetation was arranged densely), the maximum water level
reflected wave increased in size at t¼47 s. This peak was a wave and maximum velocity behind vegetation also decreased.
reflected by vegetation. Fig. 7(c) shows the time profile of a Although the difference was not very large compared with the
reflected wave in the uniform arrangement. As the tree density uniform arrangement, the difference between the densest case
increased, the reflected wave from vegetation became larger, but (case 15) and the sparsest case (case 6) was about 8%. Moreover,
the reflected wave from 1/20.5 slope became smaller. This is under the same total vegetation width, the reductions of water
because (1) the reflection in front of the vegetation zone was level and velocity were slightly larger when the back part was
increased, (2) the height of the wave that reached behind the dense.
vegetation was decreasing, and (3) the wave reflected by the 1/ For this reason, the force acting on the cylinders is discussed
20.5 slope was obstructed by vegetation. Therefore, under the first. Fig. 9 shows the distribution of maximum force acting on a
same vegetation thickness (dn), the reduction of wave energy cylinder in combined arrangements (Cases 8 and 13). Both cases
behind the vegetation in the dense case has become larger than have the same total vegetation width, but the front part is dense
that in the sparse case due to increasing reflection from the in Case 8, while the back part is dense in Case 13. In both cases,
vegetation and force acting on the cylinders. the force was larger with the denser vegetation, especially at the
K. Iimura, N. Tanaka / Ocean Engineering 54 (2012) 223–232 231

0.8

Maximum water level (m)


0.06

Maximum velocity (m/s)


0.6

0.04
0.4

Measuring point x=13.00m, front is dense Measuring point x=13.00m, front is dense
0.02 Measuring point x=13.00m, back is dense 0.2 Measuring point x=13.00m, back is dense
Measuring point x=11.50m, front is dense Measuring point x=11.50m, front is dense
(Cases 6 to 15) Measuring point x=11.50m, back is dense (Cases 6 to 15) Measuring point x=11.50m, back is dense
0 0
0 200 400 600 800 1000 0 200 400 600 800 1000
Total vegetation width : W (mm) Total vegetation width : W (mm)

Fig. 8. Reduction effect behind vegetation model (combined arrangement): (a) maximum water level, and (b) maximum velocity.

Case 8 (front is dense) 0.02


0.05 Case 13 (back is dense) Without vegetation
Case 8 (front is dense)
Dense vegetation
Maximum applied force (N)

0.04 Case 13 (back is dense)

Water level (m)


Sparse vegetation 0.01

0.03

0.02 0

Sparse vegetation Dense vegetation 30 35 40 45 50 55


0.01
Time (s)
0 -0.01
10.900 11.000 11.100 11.200 11.300 11.400
Distance from wave making plate : x (m) Fig. 10. Time profile of water level in reflected wave at measuring point P2
(x¼ 4.0 m) (comparison between Case 8 and Case 13).
Fig. 9. Distribution of maximum force in combined arrangement (comparison
between Case 8 and Case 13).
case with a denser back part was slightly larger than in the case
with a denser front part. Therefore, the reductions of water level
backmost row of the vegetation. As in the uniform arrangement, and velocity increased with decreasing total vegetation width W
the water level was increased in front of the denser vegetation by under the same dn, and the reductions of water level and velocity
reflection and damming up, and the water surface slope increased slightly increased in the case with the denser back part.
inside the denser vegetation, as in Fig. 3. Thus, the force acting on
a cylinder became larger. Although the acting force in Case 13 was
large locally (the backmost row of vegetation) in comparison with 5. Conclusions
that in Case 8, the acting force in Case 8 was slightly larger or
almost the same on the whole. Laboratory experiments using a wave channel and numerical
Then, a calculation to separate incident and reflected waves simulations were carried out to investigate the effects of the
was performed for the combined arrangement to confirm the density distribution of vegetation in mitigating a tsunami. The
effect of the reflected wave. Fig. 10 shows the time profile of effects of tree density and arrangements of different tree densities
reflected waves in Cases 8 and 13 at measuring point P2 can be summarized as follows:
(x¼4.0 m). The arrangement of the cylinders in Case 8 was the
opposite of Case 13. Although the total vegetation width was the 1. A one-dimensional numerical model of Boussinesq-type equa-
same in both cases, the times to reach the peak of the wave tions including resistance by vegetation was developed. The
reflected from the vegetation at P2 were different. The peak of the applicability was validated by experimental results within a
reflected wave was delayed in Case 13 compared with Case 8. This reasonable limit. In the numerical model, resistance from
shows that the reflection due to vegetation was greatly affected vegetation is given by drag and inertia. The mass coefficient
by density. However, because the arrival time of the reflected CM was fixed at 2.0 because the inertia was sufficiently small
wave was t ¼45 s in both cases (the water level of the reflected compared with the drag as in previous studies, and thus the
wave began to separate from that in the case without vegetation), inertia hardly affected the simulation results, and the drag
reflection also occurred in the sparser vegetation part (when the coefficient was given by calibration with the experimental
back is dense). On the other hand, the duration of the reflected results.
wave from the vegetation was shorter in Case 8 than in Case 13 2. In the uniform arrangement, the reduction of water level and
(Fig. 10). Because the front part of the vegetation was dense, it velocity behind the vegetation increased with increasing tree
obstructed the reflection from the back part of the sparser density, even under the same dn. One reason is that the wave
vegetation. Fig. 11 shows the maximum water level of the wave height of the reflected wave in front of the vegetation was
reflected by model vegetation in the combined arrangement. increased by arranging the vegetation more densely. The
When the model vegetation was arranged more densely (i.e., second reason is that the water surface slope inside the
W was small), the maximum water level of the reflected wave vegetation increased with increasing tree density, and the
increased. The maximum water level of the reflected wave in the resistance due to the vegetation thus increased.
232 K. Iimura, N. Tanaka / Ocean Engineering 54 (2012) 223–232

0.015 R-max: Maximum water level of reflected wave


Goda, Y., 1985. Random Seas and Design of Maritime Structures. University of
Tokyo Press, Japan.
by vegetation model Harada, K., Imamura, F., 2005. Effects of coastal forest on tsunami hazard
mitigation—a preliminary investigation, tsunamis: case studies and recent
development. In: Satake, K. (Ed.), Advances in Natural and Technological
0.01 Hazards Research. Springer, pp. 279–292.
R-max (m)

Hiraishi, T., Harada, K., 2003. Greenbelt tsunami prevention in South-Pacific


region. Rep. Port Airport Res. Inst. 42 (2), 23.
Iimura, K., Tanaka, N., Harada, K., Tanimoto, K., 2010. Tsunami mitigation effects
by the combination of vegetation with different tree density. J. Jpn. Soc. Civ.
0.005 Eng. Ser. B2 (Coastal Eng.) 66 (1), 281–285 (in Japanese).
Combined (front is dense) Iwasaki, T., Mano, A., 1979. Numerical computations of two dimensional run-up of
tsunamis due to the Eulerian description. In: Proceedings of the 26th Japanese
Combined (back is dense) Conference on Coastal Engineering, JSCE, pp. 70–74, in Japanese.
(Cases 6 to 15)
Kathiresan, K., Rajendran, N., 2005. Coastal mangrove forests mitigated tsunami,
0 short note. Estuarine Coastal Shelf Sci. 65, 601–606.
0 200 400 600 800 1000 Madsen, P.A., Sorensen, O.R., 1992. A new form of Boussinesq equations with
Total vegetation width : W (mm) improved linear dispersion characteristics. Part 2. A slowly-varying bathymetry.
Coastal Eng. 18, 183–204.
Fig. 11. Maximum water level of reflected wave in vegetation model (combined Mascarenhas, A., Jayakumar, S., 2008. An environmental perspective of the
arrangement). post-tsunami scenario along the coast of Tamil Nadu, India: role of sand
dunes and forests. J. Environ. Manage. 89, 24–34.
Matsutomi, H., Sakakiyama, T., Nugroho, S., Matsuyama, M., 2006. Aspects of
inundated flow due to the 2004 Indian Ocean Tsunami. Coastal Eng. J. 48,
3. In the combined arrangement, the reduction of water level and 167–195.
velocity behind vegetation also increased with increasing tree Nandasena, N.A.K., Tanaka, N., Tanimoto, K., 2008. Tsunami current inundation of
density (i.e., total vegetation width decreased). The reduction ground with coastal vegetation effects; an initial step towards a natural
solution for tsunami amelioration. J. Earthquake Tsunami 2 (2), 157–171.
was slightly larger in the case in which the back part was Nepf, H.M., 1999. Drag, turbulence, and diffusion in flow through emergent
denser than in the case in which the front part was denser. The vegetation. Water Resour. Res. 35 (2), 479–489.
reflection by vegetation is mainly affected by the denser Shuto, N., 1987. The effectiveness and limit of tsunami control forests. Coastal Eng.
Jpn. 30 (1), 143–153.
vegetation part, and if the front part of the vegetation is
Takemura, T., Tanaka, N., 2007. Flow structures and drag characteristics of a
denser, the front part of vegetation obstructs the reflection colony-type emergent roughness model mounted on a flat plate in uniform
from the sparser back-part vegetation. flow. Fluid Dyn. Res. 39, 694–710.
Tanaka, N., Sasaki, Y., Mowjood, M.I.M., Jinadasa, K.B.S.N., 2007. Coastal vegetation
structures and their functions in tsunami protection: experience of the recent
In future, experimental and numerical studies considering the Indian Ocean tsunami. Landscape Ecol. Eng. 3, 33–45.
vertical stand structure of trees on a real scale need to be performed. Tanaka, N., Nandasena, N.A.K., Jinadasa, K.S.B.N., Sasaki, Y., Tanimoto, K., Mowjood,
M.I.M., 2009. Developing effective vegetation bioshield for tsunami protection.
J. Civ. Environ. Eng. Syst. 26, 163–180. (Taylor & Francis).
Tanaka, N., 2009. Vegetation bioshields for tsunami mitigation: review of
Acknowledgments the effectiveness, limitations, construction, and sustainable management.
Landscape Ecol. Eng. 5, 71–79.
This study was supported in part by Grant-in-Aid for JSPS Tanaka, N., Jinadasa, K.B.S.N., Mowjood, M.I.M., Fasly, M.S.M., 2011. Coastal
vegetation planting projects for tsunami disaster mitigation—effectiveness
Fellows (22 7964) and the JSPS AA Science Platform Program. evaluation of new establishments. Landscape Ecol. Eng. 7, 127–135.
Tanimoto, K., Tanaka, N., Nandasena, N.A.K., Iimura, K., Shimizu, T., 2007.
References Numerical simulation of tsunami prevention by coastal forest with several
species of tropical tree. Annu. J. Coastal Eng. JSCE 54 (2), 1381–1385
(in Japanese).
Chow, V.T., 1959. Open-channel Hydraulics. McGraw-Hill Publishing Co., Thuy, N.B., Tanimoto, K., Tanaka, N., Harada, K., Iimura, K., 2009. Effect of open gap
New York. in coastal forest on tsunami run-up—investigations by experiment and
Danielsen, F., Sorensen, M.K., Olwig, M.F., Selvam, V., Parish, F., Burgess, N.D., numerical simulation. Ocean Eng. 36, 1258–1269 Elsevier.
Hiraishi, T., Karunagaran, V.M., Rasmussen, M.S., Hansen, L.B., Quarto, A., Thuy, N.B., Tanimoto, K., Tanaka, N., 2010. Force due to tsunami runup around a
Suryadiputra, N., 2005. The Asian tsunami: a protective role for coastal coastal forest with a gap—experiments and numerical simulations. Sci.
vegetation. Science 310, 643. Tsunami Hazards 29 (2), 44–69.

You might also like